ABSTRACT Title of Dissertation: IMPROVEMENTS AND STUDIES OF PLANAR TRANSMON QUBITS Yizhou Huang Doctor of Philosophy, 2023 Dissertation Directed by: Professor Christopher J. Lobb Department of Physics Dr. Benjamin S. Palmer Laboratory for Physical Sciences This dissertation describes three main projects focused on characterizing and im- proving superconducting transmon qubits operating nominally at temperature of 20 mK. The first topic I discuss is characterization of ground state fidelity of a passively cooled 3D transmon qubit using two techniques. The first technique was counting the number of false counts when performing single-shot read-out measurements of the weak resonator signal using a nearly quantum limited traveling wave parametric am- plifier. Over about a million shots, only 772 counts were found with the system in the excited state, corresponding to a residual excited state population of Pe = 0.083%. The second technique used was performing Rabi oscillations between the first and second excited state levels of the transmon qubit. By fitting the data, the residual excited state population was shown to be Pe = 0.088%±0.018%. These state of the art low values for the infidelity of the ground state suggest that the effective temperature of the transmon qubit with fundamental transition frequency of 3.6 GHz was T < 25 mK. The second topic I discuss is improvements in the coherence of our planar trans- mon qubits such that I was able to measure energy relaxation times and coherence times up to tens of microseconds. There were two main improvements that I achieved during this research. First, I identified a significant loss mechanism associated with the way that the planar transmon qubits were packaged. I did this by measuring the internal quality factors (Qi) for a series of thin-film Al quarter-wave resonators with fundamental resonant frequencies varying between 4.9 and 5.8 GHz. By utilizing resonators with different widths and gaps, I sampled different electromagnetic energy volumes that affected Qi. When the backside of the sapphire substrate of the res- onator device was adhered to a Cu package with a conducting silver glue, a monotonic decrease in the maximum achievable Qi was found as the electromagnetic sampling volume was increased. Simulations and modeling showed that the observed dissipa- tion was a result of induced currents in large surface resistance regions underneath the substrate. By placing a hole underneath the substrate and using superconducting material for the package, I was able to decrease the Ohmic losses and increase the maximum Qi by an order of magnitude for the larger size resonators. The second improvement I made to achieve improvements to our planar transmon qubits was de- veloping a new fabrication process to improve the quality of the interface between the substrate and the superconducting shunting capacitor of the transmon. For this new process, the large features (> 1 µm) of the thin superconducting film are subtractively defined by etching the film. The small Al/AlOx/Al junction is added in a second step by defining the junction in electron-beam lithography, an ion mill step, and standard double-angle evaporation. Finally, I discuss my study of the coherence recovery time after injecting quasi- particles in several transmon qubits. Quasiparticles in the transmon junction were created by applying a large-amplitude microwave pulse resonant with the readout res- onator. Immediately after generating the quasiparticles, a significant decrease of the energy relaxation time of the transmon qubit was observed. By performing relaxation time T1 and coherence time T2 measurements over the course of several milliseconds, I tracked the recovery T1 and T2, which I then used as metrics of the quasiparticle density at the junction. I fitted the recovery data with a numerical model involving differential equations and extracted the quasiparticle trapping rates around 1ms−1 and recombination rates around 1/25ns−1 at the sites of a few transmon qubits. I measured transmons that were either galvanically connected to ground or isolated, fabricated with aluminum (with superconducting gap ∆Al ≃ 200 µV) or tantalum (with ∆Ta ≃ 600 µV) for the shunting capacitor. Finally, with a larger quasiparti- cle injection power and by measuring two transmons on the same chip, I observed a phenomenon that was consistent with phonon-assisted quasiparticle poisoning. I dis- cussed my quantitative modelling of such data, and how this effect presents challenges to further improving the coherence times of transmon qubits. IMPROVEMENTS AND STUDIES OF PLANAR TRANSMON QUBITS by Yizhou Huang 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 2023 Advisory Committee: Professor Christopher J. Lobb, Chair/Co-Advisor Dr. Benjamin S. Palmer, Primary-Advisor Professor Frederick C. Wellstood Professor Steven Anlage Professor Thomas Murphy, Dean’s Representative ACKNOWLEDGMENTS I wish to express my sincere appreciation to the individuals and institutions who have been instrumental in guiding me through the journey to such degree. Foremost, my heartfelt thanks go to my advisor, Dr. Benjamin S. Palmer, for his unwavering mentorship and dedication to my academic development. His expertise, thoughtful feedback, and patient guidance have significantly influenced the course of my research. I extend my gratitude to the members of my dissertation committee, Profes- sor Frederick C. Wellstood, Professor Christopher J. Lobb, Professor Steven Anlage, and Professor Thomas Murphy, for their valuable insights and constructive critiques, which have enriched the quality of this work. I would also like to acknowledge the camaraderie and support of my colleagues and fellow researchers at LPS and UMD - Jen-Hao Yeh, Haozhi Wang, Shavindra Pre- maratne, Rui Zhang, Tamin Tai, Zachary Steffen, Yi-Hsiang Huang, Jonathan Cripe, Sudeep Dutta, Kungang Li, Neda Forouzani, Chih-Chiao Hung, Martin Ritter. The collaborative environment we’ve shared, the stimulating discussions, and the shared enthusiasm for discovery have been fundamental in fostering creativity showcased in this dissertation. I want to thank the LPS cleanroom, machine shop, research, and administrative staff members, that provided crucial support to ensure the smooth operation of our research facilities. Your tireless efforts often go unnoticed, but they are deeply valued and have significantly contributed to the success of my research. I wish this dissertation shall serve as a testament to the collective effort of those who have been part of this journey with me. Thank you. ii TABLE OF CONTENTS Page LIST OF TABLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii LIST OF FIGURES. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . viii CHAPTER 1 INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 1.1 Quantum Computation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 1.2 Increasing Qubit Lifetime . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3 1.3 Overview of the Dissertation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5 2 THEORY OF TRANSMON AND READOUT RESONATOR . . . . . . . . . . . 7 2.1 Coplanar Waveguide Resonator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7 2.2 Theory of the Transmon . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9 2.2.1 Josephson Junction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9 2.2.2 Transmon Qubit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11 2.2.3 Transmon Coupled to Readout Resonator . . . . . . . . . . . . . . . . . . 15 2.2.4 Measuring the State of the Transmon. . . . . . . . . . . . . . . . . . . . . . . 17 2.2.5 Numerical Simulation of Jaynes Cummings Readout . . . . . . . . 19 2.2.6 Estimating the Voltage across the Junction . . . . . . . . . . . . . . . . . 26 2.2.7 Driving Transmon Transitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30 2.2.8 Bloch Sphere Representation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32 2.2.9 Density Matrix Representation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34 2.2.10 Lindblad-Kossakowski Master Equation and Decoherence . . . 35 2.2.11 Determining Optimal Readout Length . . . . . . . . . . . . . . . . . . . . . . 38 2.3 Loss Mechanisms of Transmon. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40 2.3.1 Quasiparticle Loss . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40 2.3.2 Two-Level System Loss . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42 iii CHAPTER Page 2.3.3 Purcell Loss . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44 2.3.4 Package Loss . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45 2.3.5 Vortex Loss . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45 3 DEVICE DESIGN AND FABRICATION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47 3.1 Device Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47 3.1.1 Finite Element Simulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51 3.1.2 Black Box Quantization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53 3.2 Device Fabrication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55 3.2.1 Photolithography and Wet Etching of Large Superconduct- ing Features . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56 3.2.2 Photolithography of Alignment Marks . . . . . . . . . . . . . . . . . . . . . . 59 3.2.3 Electron-Beam Lithography of the Josephson junction . . . . . . 61 3.3 Tuning Oxidation Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69 4 SYSTEM CHARACTERIZATION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73 4.1 Dilution Refrigerator Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73 4.2 Measurement Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77 4.2.1 Vector Network Analyzer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77 4.2.2 Pulsed Heterodyne Measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . 78 4.2.3 Generating Microwave Pulses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79 4.3 Overview of Measured Devices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82 4.4 Resonator Spectroscopy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86 4.5 Transmon Readout . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89 4.6 Transmon Characterization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92 4.6.1 Transmon Spectroscopy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92 iv CHAPTER Page 4.6.2 Rabi Oscillations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96 4.6.3 Energy Relaxation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97 4.6.4 Ramsey Oscillations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101 4.6.5 Spin Echo Measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104 4.6.6 Photon Number Splitting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110 4.6.7 Residual Excited State Population. . . . . . . . . . . . . . . . . . . . . . . . . . 113 5 REDUCING LOSS FROM THE PACKAGE . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128 5.1 Overview of Package Loss . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128 5.2 Measuring Resonator Loss . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129 5.3 Modelling Resonator Loss . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134 5.4 Finite Element Simulations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138 5.4.1 Surface Simulation with COMSOL . . . . . . . . . . . . . . . . . . . . . . . . . 143 5.4.2 Dielectric Loss . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145 5.5 Extension to Transmon Qubits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145 5.5.1 Improvements in Packaging . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149 5.5.2 Backside Metalization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150 6 MEASURING QUASIPARTICLE LIFETIMES . . . . . . . . . . . . . . . . . . . . . . . . . 154 6.1 Generating Quasiparticles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154 6.2 Probing Quasiparticle Dynamics with Qubit Coherence . . . . . . . . . . . . 157 6.3 Extracting Quasiparticle Dynamics from Experiment Data . . . . . . . . . 160 6.3.1 Simulation Setup. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160 6.3.2 Results and Discussions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164 6.4 Additional Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169 6.5 Quasiparticle-Phonon Effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172 v CHAPTER Page 7 CONCLUSION. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177 7.1 Summary of Main Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177 7.1.1 Residual Excited State Population. . . . . . . . . . . . . . . . . . . . . . . . . . 177 7.1.2 Improvements in Microwave Packaging . . . . . . . . . . . . . . . . . . . . . 178 7.1.3 Quasiparticle Lifetime Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178 7.2 Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 179 A APPENDIX. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181 A.1 Code for Numerical Simulation of Jaynes Cummings Readout . . . . . . 181 REFERENCES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188 vi LIST OF TABLES Table Page 4.1 Parameters of measured qubits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83 4.2 Parameters of measured resonators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84 4.3 Comparison of observed P∣e⟩ and Teff to values reported in the literature 127 5.1 Resonator Design Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129 5.2 Measured resistivity of normal metal present in the packaging . . . . . . . . . 138 5.3 HFSS simulated magnetic field geometric factors γ . . . . . . . . . . . . . . . . . . . . 141 5.4 Package limited T1 of various qubit designs . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147 5.5 Contribution from various PCB surfaces to qubit loss . . . . . . . . . . . . . . . . . 147 5.6 Qubit T1 limit due to a grid backside metalization . . . . . . . . . . . . . . . . . . . . . 152 6.1 QP injection and fitting parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167 vii LIST OF FIGURES Figure Page 1.1 Different types of superconducting qubits distinguished by energy scales 3 1.2 Superconducting qubit coherence time versus year . . . . . . . . . . . . . . . . . . . . . 4 2.1 Cross-sectional schematics of a coplanar waveguide and the associated electric and magnetic fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8 2.2 3D rendering and circuit symbol of a Josephson junction . . . . . . . . . . . . . . 10 2.3 Circuit diagram of transmon qubit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12 2.4 Transmon qubit energy levels for four different Ej/Ec ratios . . . . . . . . . . . 13 2.5 Schematics of angular resonant frequency for a notch resonator coupled to a transmon qubit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 2.6 Relative frequency of resonator vs stored photon number n for qubit in ∣g⟩ and ∣e⟩ states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21 2.7 Stored resonator photon number vs input power for qubit in ∣g⟩ and ∣e⟩ states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23 2.8 Relationship between transmitted voltage and input voltage . . . . . . . . . . . 23 2.9 Simulated ∣S21∣ vs input power for qubit in ∣g⟩ and ∣e⟩ states . . . . . . . . . . . 25 2.10 Simulated transmission ∣S21∣ of the readout resonator vs frequency and power . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26 2.11 Lumped-element circuit model to estimate the voltage across the Joseph- son junction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27 2.12 False color plot of numerically simulated voltage across a Josephson junction during high power Jaynes Cummings readout. . . . . . . . . . . . . . . . . 30 2.13 Bloch sphere representation of qubit state and rotation . . . . . . . . . . . . . . . . 33 2.14 Bloch sphere representation of qubit decoherence . . . . . . . . . . . . . . . . . . . . . . 38 2.15 Figurative representation of SNR with respect to readout length . . . . . . . 40 viii CHAPTER Page 2.16 Double-well potential model of TLS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43 3.1 2D design layout of multiplexed resonators and transmon qubits with annotations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48 3.2 2D design layout of single transmon qubit with annotations . . . . . . . . . . . 49 3.3 2D design layout of structures for e-beam lithography and Josephson junction fabrication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50 3.4 Setup of HFSS simulation to help develop device design . . . . . . . . . . . . . . . 52 3.5 Example of simulation mesh used in HFSS . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53 3.6 Simulated transmission S21 versus frequency as simulated by HFSS . . . . 54 3.7 Results from black-box quantization showing the imaginary part of admittance Y versus frequency. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55 3.8 Photolithography and wet etching process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57 3.9 Mask design for photolithography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58 3.10 Photolithography for alignment marks process. . . . . . . . . . . . . . . . . . . . . . . . . 60 3.11 Micrograph of alignment marks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62 3.12 E-beam lithography for Josephson junction fabrication . . . . . . . . . . . . . . . . 63 3.13 Picture of alignment process on JEOL 6500F SEM screen . . . . . . . . . . . . . 65 3.14 Confocal laser micrograph during different stages of junction processing 66 3.15 SEM image of cross-section of ZEP-MMA bilayer e-beam resist . . . . . . . . 67 3.16 Chip loaded into an e-beam evaporator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68 3.17 SEM image of Josephson junction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69 3.18 2D design layout of structures for testing the resistance of Josephson junction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70 3.19 Measured Josephson junction resistance versus oxidation pressure. . . . . . 71 ix CHAPTER Page 4.1 Photograph of the dilution refrigerators used in this dissertation . . . . . . . 74 4.2 Dilution refrigerator setup diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75 4.3 VNA measurement setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77 4.4 Setup for heterodyne measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79 4.5 Pulse sequences of digital delay generators . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81 4.6 Micrographs of packaged qubit devices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85 4.7 Sample of VNA measured resonator spectroscopy and fit . . . . . . . . . . . . . . 87 4.8 S21 versus power for device G3-i . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88 4.9 Pulse sequence for S21 measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89 4.10 Pulsed heterodyne measurement of ∣S21∣2 vs power and frequency for different qubit states for qubit C-i in high power Jaynes Cummings regime . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90 4.11 Pulsed heterodyne measurement of ∣S21∣2 vs power and frequency for different qubit states for qubit C-i in low power dispersive regime . . . . . . 91 4.12 Pulse sequence for qubit spectroscopy measurements . . . . . . . . . . . . . . . . . . 92 4.13 Transmon spectroscopy for qubit C-i with Lorentzian fits . . . . . . . . . . . . . . 94 4.14 Repeated spectroscopy measurement for qubit C-i . . . . . . . . . . . . . . . . . . . . . 95 4.15 Repeated spectroscopy measurement for qubit A1 . . . . . . . . . . . . . . . . . . . . . 95 4.16 Pulse sequence for Rabi oscillation measurements . . . . . . . . . . . . . . . . . . . . . 96 4.17 Sample Rabi oscillation data for qubit C-i . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98 4.18 Measured Rabi oscillation and fit for qubit C-i . . . . . . . . . . . . . . . . . . . . . . . . 99 4.19 Pulse sequence for T1 measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99 4.20 T1 measurement and fit for qubit C-i . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100 4.21 Pulse sequence for Ramsey oscillation measurements . . . . . . . . . . . . . . . . . . 102 x CHAPTER Page 4.22 Bloch sphere representation of Ramsey oscillation measurements . . . . . . . 102 4.23 Ramsey oscillation measurement and fit for qubit D . . . . . . . . . . . . . . . . . . . 103 4.24 Frequency jumps for qubit H-i as observed from Ramsey measurements 105 4.25 Pulse sequence for spin echo measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106 4.26 Bloch sphere representation of spin echo measurements . . . . . . . . . . . . . . . . 107 4.27 Spin echo measurement and fit for qubit J-i . . . . . . . . . . . . . . . . . . . . . . . . . . . 108 4.28 Spin echo measurement and fit for qubit C-i . . . . . . . . . . . . . . . . . . . . . . . . . . . 109 4.29 Pulse sequence for photon splitting measurement . . . . . . . . . . . . . . . . . . . . . . 111 4.30 Data and fit for photon number splitting measurements of qubit C-iii . . 112 4.31 Pulse sequence for qubit ∣g⟩/∣e⟩/∣f⟩ state single shot measurement . . . . . 115 4.32 False color map of ∣g⟩, ∣e⟩, and ∣f⟩ states from single shot measurement for qubit J-i . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116 4.33 Fit of false color map of ∣g⟩, ∣e⟩, and ∣f⟩ states from single shot mea- surement for qubit J-i . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117 4.34 Pulse sequence for Rabi ∣e⟩↔ ∣f⟩ oscillation measurement . . . . . . . . . . . . . 120 4.35 Illustration of Rabi ∣e⟩↔ ∣f⟩ oscillations in IQ plane. . . . . . . . . . . . . . . . . . . 121 4.36 Data and fit for Rabi ∣e⟩ ↔ ∣f⟩ oscillations at base temperature for qubit J-iii . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122 4.37 Data and fit for Rabi ∣e⟩↔ ∣f⟩ oscillations at elevated temperature for qubit J-iii . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124 4.38 Effective qubit temperature Teff vs mixing chamber temperature Tmix for qubit J-iii’s ∣e⟩↔ ∣f⟩ Rabi oscillation measurements . . . . . . . . . . . . . . . . 125 xi CHAPTER Page 4.39 Data and fit for Rabi ∣e⟩ ↔ ∣f⟩ oscillations at base temperature for qubit E-ii . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126 4.40 Data and fit for Rabi ∣e⟩ ↔ ∣f⟩ oscillations at base temperature for qubit H-ii . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126 5.1 Stitched optical micrograph of the test resonator chip . . . . . . . . . . . . . . . . . 130 5.2 Photos of two device packages . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132 5.3 Log-log plot of the fitted resonator Qi versus stored photon number for all four packages . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133 5.4 Measured maximum Qi,m on a log scale for each resonator and package 134 5.5 Photographs of 3D cavity made of OFHC copper . . . . . . . . . . . . . . . . . . . . . . 136 5.6 Plot of quality factor for 3D cavity and extracted resistivity versus temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136 5.7 Picture of half-wave PCB resonator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137 5.8 Confocal microscope image of silver impregnated glue used for DC resistivity measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138 5.9 Simulation setup of resonator chip and surrounding PCB . . . . . . . . . . . . . . 140 5.10 Comparison of simulated surface current density between packages. . . . . 142 5.11 Comparison of estimated and measured resonator loss 1/Qi,m . . . . . . . . . . 143 5.12 2D COMSOL simulation of the ∣H ∣ field of a TEM wave in a CPW . . . . 144 5.13 Common planar qubit designs studied for package loss . . . . . . . . . . . . . . . . . 146 5.14 Cross section view of PCB and substrate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148 5.15 Photographs of device A . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149 5.16 Setup of HFSS simulation to study qubit T1 limit due to backside metalization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151 xii CHAPTER Page 5.17 Magnetic field distribution on back side of substrate . . . . . . . . . . . . . . . . . . . 152 5.18 Sample meshing for HFSS simulation of backside metalization . . . . . . . . . 153 6.1 Semiconductor model energy diagram of a S-I-S Josephson junction at zero temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155 6.2 Pulse sequence for measuring qubit’s energy relaxation time T1 after quasiparticle injection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158 6.3 Sample T1 measurement data with and without quasiparticle injection . 158 6.4 Additional qubit decay rate δΓ plotted against separation time t after quasiparticle injection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160 6.5 Partition of transmon geometry for numerical simulation . . . . . . . . . . . . . . 162 6.6 Measurement of T1 recovery after quasiparticle injection for qubit A1- xmon and simulations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164 6.7 Measurement of T1 recovery after quasiparticle injection for qubit A2- xmon and simulations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165 6.8 Measurement of T1 recovery after quasiparticle injection for qubit A3- float and simulations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165 6.9 Measurement of T1 recovery after quasiparticle injection for qubit A4- float and simulations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166 6.10 Measurement of T1 recovery after quasiparticle injection for qubit B- Ta-float and simulations.. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166 6.11 Simulated T1 recovery after quasiparticle injection for different geomet- rical variations of qubit A1-xmon. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168 6.12 Additional quasiparticle density as calculated from measured fractional frequency shift and change in the decay rate for qubit B-Ta-float . . . . . . 170 xiii CHAPTER Page 6.13 Measured coherence recovery after quasiparticle injection for qubit B- Ta-float . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170 6.14 Measured coherence recovery after quasiparticle injection for qubit C- xmon. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171 6.15 Measurement of T1 recovery after quasiparticle injection for qubit H- xmon. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172 6.16 Measurements and Simulations of T1 recovery of qubit A3-float after quasiparticle injection at qubit A2-xmon. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173 6.17 Measurements and Simulations of T1 recovery of qubit A1-xmon after quasiparticle injection at qubit A2-xmon. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175 A.1 Simulated transmission ∣S21∣ of the readout resonator vs frequency and power, with qubit frequency above resonator . . . . . . . . . . . . . . . . . . . . . . . . . . 187 xiv Chapter 1 INTRODUCTION 1.1 Quantum Computation The exponential growth of information has fueled an unceasing pursuit for more potent computing systems capable of addressing the escalating demands of modern society. Computation, as we have come to understand it, has predominantly relied on classical computers that manipulate discrete bits represented by the binary digits of 0 and 1. However, parallel to these advances in classical computation, our deepest understanding of the physical world has been shaped by the principles of quantum mechanics, a field of study that unveils the astonishing behavior of matter and energy at the quantum level. Quantum mechanics, formulated in the early 20th century, elucidates the funda- mental laws that govern the behavior of particles at microscopic scales. It reveals a reality that is probabilistic, often non-intuitive, and characterized by phenomena such as superposition and entanglement. Quantum mechanics lay the foundation for the concept of a “qubit”, the fundamental unit of quantum information. In contrast to a classical bit, which has only two discrete states (0 or 1), a qubit can be in a simultaneous superposition of both 0 and 1 states. If ∣ψ⟩ represents a qubit state, it can be expressed as ∣ψ⟩ = α∣0⟩ + β∣1⟩, (1.1) with complex numbers α and β satisfying ∣α∣2 + ∣β∣2 = 1. 1 Moreover, two or more qubits can also be entangled, a phenomenon where the quantum states of multiple qubits become inherently interconnected, regardless of their physical separation. As a result, in general it requires ∼ 2n complex numbers on a classical computer to completely store the information of the Hilbert space spanned by n qubits. Exploiting entanglement allows for the execution of parallel computations across multiple qubits, providing an exponential speedup in certain problem-solving scenarios compared to classical computers. Several algorithms for quantum computation have been proposed that would yield a speedup against best-known classical algorithms, including Shor’s algorithm for discrete factoring [125], Grover’s algorithm for unobstructed search [49], and the HHL algorithm for solving linear systems [54]. Experimentally, there has been reports [3, 153, 148] over the past few years of quantum computers achieving “quantum supremacy”, solving problems faster than all existing classical computers, although the problems solved in these demonstrations were of little predicted use. There are many physical implementations for a qubit, including neutral atoms [121], trapped ions [91, 105], nuclear spins [62, 64], quantum dots [76] and various superconducting circuits [93]. In particular, qubits based on superconducting circuits can be made from capacitors, inductors and Josephson junctions. The Josephson junction provides energy levels with nonlinearity, which is critical for constructing a qubit (see Sec. 2.2.1). Depending on the choice of other reactive circuit components, superconducting qubits can be roughly categorized into charge qubit [141], phase qubit [83, 130], flux qubit [38, 138], transmon [68], fluxonium [78, 106], unimon [57], and many more. Fig. 1.1 show different types of superconducting qubits, as distin- guished by the relations between the Josephson energy EJ , the inductive energy EL and the charging energy Ec. Among these, this dissertation in particular is focused on the study of transmon qubits. 2 Figure 1.1: Different types of superconducting qubits distinguished by the ratio of the Josephson energy EJ and inductive energy EL to charging energy Ec. See Sec. 2.2.2 for details of some of the energy definitions. Figure from Ref. [57] under Creative Commons CC by license. 1.2 Increasing Qubit Lifetime Unlike a classical bit, the continuous nature of quantum states means a qubit is inherently vulnerable to the problem of decoherence. This means a qubit may only preserve quantum information for a limited lifetime before such information is lost due to its interaction with the environment. Two common metrics characterizing the coherent behavior of a qubit are its energy relaxation time T1 and coherence time T2 (see Sec. 2.2.10 and 4.6.3-4.6.5 for details). As shown in Fig. 1.2, over the past two decades,T1 and T2 for superconducting qubits has increased significantly. Despite that, lifetime in these systems cannot support arbitrarily long quantum computations, contrary to classical computation with very low bit error rates. To address this issue, quantum error correction, such 3 Figure 1.2: Superconducting qubit coherence time versus year, figure from Ref. [127] under license number 5574400977925. as Shor’s code [126] and the surface code [37], has been proposed. Due to the no- cloning theorem [100], an arbitrary quantum state cannot be “copied” in a classical sense. Instead, quantum error correction works by spreading the information of one (logical) qubit onto a highly entangled state of several (physical) qubits, and multi- qubit measurements are intermittently performed, which do not disturb the quantum information in the encoded state, but retrieve information about potential errors on a subset of the physical qubits (usually one), and then correct for those errors. Recently, a quantum error correction scheme extending the lifetime of a logical qubit beyond the longest lifetime of its constituent physical qubits has been demonstrated [97]. For quantum error correction schemes to work, the error rate per quantum gate op- eration has to be below a certain threshold (typically below 1%) [113, 37]. Lower error 4 rates require a smaller number of auxiliary qubits. Currently, the fidelities for some qubit gates for superconducting qubits are limited by decoherence times [132, 35]. In this case, improving the lifetime of qubits could lead to meaningful expansion of the capabilities of near-term quantum computers. In Chapter 5, I discuss improving the coherence of my superconducting planar resonators and transmon qubits by im- proving the package environment. In most error correction schemes, the error rate is assumed to be Markovian and uncorrelated between different qubits. In contrast, it has been shown that bursts of cosmic-ray muons and γ-rays could generate a large amount of quasiparticles, resulting in correlated errors of multiple qubits on the same chip [145]. In Chapter 6, I present my studies on the quasiparticle dynamics that affects superconducting qubits. 1.3 Overview of the Dissertation In Chapter 2 of this dissertation, I discuss the theory of transmon qubits and the Hamiltonian when it’s coupled to a readout resonator. I lay out the numerical simulations of the energy levels that qualitatively captures the characteristics of the high-power Jaynes Cummings readout [116]. These simulations were needed for the studies of quasiparticle dynamics in Chapter 6, where quasiparticles were inadver- tently generated during readout. In Chapter 3, I discuss how I designed my transmon qubits and coplanar waveguide resonators, and the steps I took to fabricate them. In Chapter 4, I discuss the experimental setup that I used to measure and char- acterize the devices, determining parameters such as resonant transition frequencies, coupling strengths, as well as T1 and T2 of the transmon qubits and Qi of the res- onators. In particular, in Sec. 4.6.7 I discuss how I measured the residual excited 5 state population of some of my qubits, which are among the lowest observed in su- perconducting qubits. In Chapter 5, I discuss my measurement of the internal quality factors Qi of five resonators on a single chip and how such values were related to the packaging environment. Using finite element simulations, I identified a conducting adhesive to be the dominant source of loss for some of the resonators. I extended these simulations to show how the T1 of a transmon qubit could be limited by the packaging environment in the absence of such conducting adhesive. These studies led to the design of a new package that greatly improved the maximum observed T1 for planar transmons in our group. In Chapter 6, I discuss my investigations of quasiparticle diffusion, trapping and recombination rates. Quasiparticles were injected into the qubit junction using a high power microwave pulse at the frequency of the readout resonator. By fitting a numerical model to my result, I found the extracted quasiparticle dynamics are similar to what other groups have reported. In particular, I observed an effect consistent with phonon-assisted quasiparticle poisoning across a long distance on the chip, with a somewhat longer recovery time constant than reported in the literature. Finally, I conclude in Chapter 7 with a summary of my main results together with possible directions for future studies. 6 Chapter 2 THEORY OF TRANSMON AND READOUT RESONATOR In this chapter, I discuss the theory of transmon qubits, their physical realization with Josephson junctions, their Hamiltonian when coupled to a readout resonator for state measurement, the microwave manipulation of transmon states, and predominant channels of loss. By numerically solving the eigenenergies of a transmon coupled to a readout resonator, I obtained qualitative insight into the high-power Jaynes Cummings Readout in Sec. 2.2.5. For some of my qubits, I found the quasiparticles were generated during the high-power Jaynes Cummings, which inspired my studies into their dynamics, as discussed in Chapter 6. 2.1 Coplanar Waveguide Resonator A coplanar waveguide (CPW) is a planar transmission line fabricated on a sub- strate, consisting of a center conductor stripe and a pair of ground planes on either side [143]. This geometry supports a quasi-TEM (transverse electromagnetic) mode of transmission. Fig. 2.1 shows a cross-sectional view of the components of a CPW, and the distribution of electric and magnetic fields of quasi-TEM propagation. A CPW can be modelled with a distributed capacitance Cgeo and distributed inductance Lgeo. Let the width of the center trace be w and the gap between the center trace and ground plane be g. I assume w and g are much larger than the thickness t of the metal layers, but much smaller than the thickness h of substrate underneath them, with substrate relative dielectric constant ϵr. These are valid assumptions for 7 air electric field magnetic field metal dielectric w g h t Figure 2.1: Cross-sectional schematics of a coplanar waveguide and the associated electric and magnetic fields. Figure adopted from Spinningspark at Wikipedia under CC BY-SA 3.0 license. the devices studied in this dissertation. Then [128] Cgeo =2ϵ0(ϵr + 1) K(k0) K(k′0) (2.1) Lgeo = µ0 4 K(k′0) K(k0) (2.2) Z0 = √ Lgeo Cgeo = √ µ0 8ϵ0(ϵr + 1) K(k′0) K(k0) , (2.3) where K(k) is a complete elliptic integral with modulus k, k0 = w/(w + 2g) and k′0 = √ 1 − k20. For impedance matching, the values of w and g were designed to give Z0 = 50 Ω. To create a resonant structure out of a CPW, typically two discontinuities along the waveguide are introduced, associated with impedances different from Z0. Typical physical realizations of discontinuities are a gap in the center trace or an inductive short to the ground. The discontinuities prevent travelling wave structures and thus create local storage of energy in the form of oscillating electromagnetic fields. Su- perconducting CPW resonators have been used as optical detectors for astronomical applications [29] and for readout in quantum information systems [45]. All of the CPW resonators that I studied in this dissertation were quarter-wave resonators, created by shorting the center trace to the surrounding ground plane on one end (voltage node, current anti-node), and making the other end open (current 8 node, voltage anti-node), with the distance between the two ends l related to the wavelength of the fundamental resonant mode λ by l = λ/4, hence the name “quarter” wave. 2.2 Theory of the Transmon 2.2.1 Josephson Junction A Josephson junction consists of two superconducting electrodes with a barrier or restriction between them (see Fig. 2.2). For superconductor-insulator-superconductor (S-I-S) Josephson junction, which is the focus of this dissertation, the barrier is a thin layer of insulator in between the two superconductors. In contrast, a thin layer of normal metal can serve as a weak barrier to create a superconductor-normal metal- superconductor (S-N-S) Josephson junction. It was predicted by Brian Josephson in 1962 that a current I can flow through the junction with zero DC voltage drop [63]. This is a result of the macroscopic wavefunction of the superconductor, which extends through the barrier and allows Cooper pairs to tunnel as a coherent quantum mechanical object. In particular, the current-voltage characteristics for an ideal bare Josephson junction can be obtained from the following two relations: I =Ic sin (δϕ) (2.4) d(δϕ) dt =2eV h̵ , (2.5) where V is the voltage across the junction, I is the current that flows through the junction, δϕ is the phase difference of the two superconducting wave functions across the junction, and Ic is the junction’s critical current. From Eq. 2.5, V = 0 as long as δϕ is stationary. According to Eq. 2.4, a non- 9 Figure 2.2: 3D rendering (left) and circuit symbol (right, a cross) of a Josephson junction. zero δϕ results in a constant supercurrent Ic sin (δϕ) flowing across the junction, this is called the DC Josephson Effect. In addition, if V ≠ 0, then δϕ as well as the current through the junction I oscillates at a frequency of 2eV /h, this is called the AC Josephson Effect. Transmon qubits typically operate with small I and V biases. Performing a first order Taylor series expansion (sinx = x +O (x3)) on Eq. 2.4 gives I Ic ≃ δϕ. (2.6) Taking the time derivative of Eq. 2.6 and plugging into Eq. 2.5 leads to V = h̵ 2eIc dI dt = LdI dt , (2.7) where L = h̵ 2eIc is identified as the Josephson inductance. From Eq. 2.7 one can see that to the lowest order, a Josephson junction behaves like an inductor, where the inductance is determined by the tunneling of Cooper pairs and not a magnetic flux. The higher-order terms in the Taylor series expansion give the a non-linear Josephson inductance. As I will discuss later, this nonlinearity is essential for constructing a qubit. 10 2.2.2 Transmon Qubit A transmon qubit, short for transmission line shunted plasma oscillation qubit, consists of a Josephson junction shunted by a capacitor [68]. Fig. 2.3 shows a circuit diagram of a transmon that’s voltage biased via coupling to gate capacitance Cg. The Hamiltonian for this circuit is given by H = 4Ec (n̂ − ng)2 −Ej cos ϕ̂, (2.8) where n̂ is the number operator for excess Cooper pairs that tunnel from one side of the capacitor to the other side, ng = CgVg/e is the reduced gate charge, ϕ̂ is the operator of the difference of superconducting phase across the junction, Ec = e2/2(Cj +Cg) is the electrostatic charging energy, Ej = IcΦ0/2π is the Josephson energy and Φ0 = h/2e is the flux quantum. For transmon qubits, typically Cj ≫ Cg, where Cj contains both the shunting capacitance and the self-capacitance of the junction. In the low temperature limit, the Josephson energy Ej is related to the normal state resistance Rj of the barrier by [136] Ej = RQ Rj ∆ 8 , (2.9) where RQ = h/e2 is the resistance quantum and ∆ is the superconducting gap for the two superconducting leads, which were aluminum in this dissertation. While Eq. 2.8 can be rigorously solved, I provide some insight into the structure of the energy levels for the transmon. Using the conjugate relationship [ϕ̂, n̂] = i (2.10) and assuming ng = 0, the Hamiltonian given by Eq. 2.8 can be rewritten as H = 4Ec ( 1 i ∂ ∂ϕ̂ ) 2 −Ej cos ϕ̂. (2.11) 11 Figure 2.3: Circuit diagram of gate-biased transmon qubit. Taking a first order expansion of the cosine function, one finds H ≈ 4Ec ( 1 i ∂ ∂ϕ̂ ) 2 + 1 2 Ejϕ̂ 2 −Ej. (2.12) This resembles the Hamiltonian for a simple harmonic oscillator H = p̂2 2m + 1 2 kx̂2 = 1 2m h̵2 (1 i ∂ ∂x̂ ) 2 + 1 2 kx̂2 (2.13) with energy eigenvalues given by En = (n + 1 2 ) h̵ √ k m . (2.14) Using this as an analogy, to the lowest order of the cosine function, the energy eigenvalues of Eq. 2.12 are En ≈ (n + 1 2 ) √ 8EjEc −Ej. (2.15) For the harmonic oscillator, in Eq. 2.13, the classical amplitude related to energy E is xmax = √ 2E k . Similarly for Eq. 2.12, one can think of an equivalent “amplitude” of ϕ as ϕmax = √ 2E Ej . The approximation in Eq. 2.12 only includes terms up to O (ϕ̂2), so for the same energy E, the larger the Ej, the smaller the “amplitude” ϕmax, the 12 -2 -1 0 1 2 0 5 10 15 ng E n (n g )/ E 0 1 (0 .2 5 ) E j/Ec=0.1 -2 -1 0 1 2 0 5 10 15 ng E j/Ec=1 -2 -1 0 1 2 0 1 2 3 4 5 ng E j/Ec=10 -2 -1 0 1 2 0 1 2 3 4 ng E j/Ec=100 Figure 2.4: Transmon qubit energy levels for four different Ej/Ec ratios from 0.1 to 100. The six lowest energy levels normalized by E01(0.25) = E1(0.25) −E0(0.25) are plotted. smaller the approximation error with Eq. 2.12. Thus, the transmon energy levels resemble a linear harmonic oscillator, but only to lowest order in n. To conclude, in a qualitative manner, the anharmonicity of the transmon is negatively correlated with the ratio Ej/Ec. The exact solution for the eigenvalues of Eq. 2.8 is given by the Mathieu’s char- acteristic value functions[26] En(ng) = Eca2[ng+k(n,ng)] (− Ej 2Ec ) , (2.16) where av(q) is the v-th Mathieu’s function and k(n,ng) is a sorting function given by k(n,ng) = ∑ l=±1 [int(2ng + l 2 ) mod 2] (int(ng) + l(−1)n [(n + 1) div 2]) . (2.17) where int(x) is a function rounding x to the nearest integer, and (x div y) gives the integer quotient of x and y. The exact transmon’s energies (Eq. 2.16) are plotted in Fig. 2.4 for a variety of Ej/Ec values. As the ratio Ej/Ec increases, the charge dispersion (i.e. the dependence 13 of En on reduced gate charge ng) decreases. Also note that for a fixed Ej/Ec, the charge dispersion of a level increases as the energy level increases. For sufficiently large Ej/Ec and small enough level number n where ng can be neglected, the Hamiltonian (Eq. 2.11) can be Taylor expanded to the ϕ̂4 order in the cosine term. The energy levels are then En ≃ (n + 1 2 ) √ 8EjEc − Ec 4 (2n2 + 2n + 1) −Ej. (2.18) The equation above has an additional term of −Ec (2n2 + 2n + 1) /4 when compared to Eq. 2.15. Eq. 2.18 gives E01 = √ 8EjEc −Ec (2.19) for the transition energy between the ground state (∣g⟩) and the first excited state (∣e⟩) of a transmon. From Eq. 2.18, the transition energy between the first (∣e⟩) and second (∣f⟩) excited state of a transmon is given by E12 = √ 8EjEc − 2Ec. (2.20) As I discuss in Sec. 4.6.7, I used the E12 transition to measure the residual excited state population for some of my qubits. The anharmonicity of a transmon qubit is given by α = E12 −E01 ≃ −Ec. (2.21) From Fig. 2.4 and Eq. 2.19-2.21, an offset-charge insensitive transmon is a par- ticular kind of qubit with sufficiently large Ej/Ec ratio such that its E01 transition does not have a noticeable dependence on ng. I designed my devices to be in this regime because the local voltage and charge environment drifts and fluctuates, result- ing in changes in ng, instabilities of E01, and therefore dephasing of the transmon. [68] At the same time, a transmon has its Ej/Ec ratio not too large such that the relative anharmonicity, α/E01 ≈ √ Ec/8Ej is not too small. This ensures that it can 14 be treated as an effective two-level system with sufficient energy spacing to reduce spurious transitions being driven to the second excited state. 2.2.3 Transmon Coupled to Readout Resonator A common way to measure the state of a transmon qubit is to couple it to a bosonic field with frequency ωr/2π. For the case of a 3D transmon qubit, the bosonic field is formed from a physical 3D cavity. [90] While for a 2D or planar transmon qubit, the field is typically formed from a CPW resonator. [12] In general, when a two level system (ideal qubit) is coupled to a single-mode bosonic field, the undriven system can be described by the Jaynes-Cummings Hamiltonian [60] HJC = h̵ωr (â†â + 1 2 ) + 1 2 h̵ωqσ̂z + h̵g (âσ̂+ + â†σ̂−) , (2.22) where σ̂± = (σ̂x ± iσ̂y)/2, σ̂x, σ̂y, σ̂z are Pauli operators on the two-level system, ↠and â are creation and annilation operators on the bosonic field, ωq is the transition frequency of the two-level system, and g is the coupling strength between the two- level system and the bosonic field. Since most of my transmon qubits studied in this dissertation (Table 4.1) were 2D transmon qubits, I use the language of a transmon qubit coupled to a resonator for the remainder of this chapter. The Jaynes-Cummings Hamiltonian for a multi-level transmon with nonuniform energy spacing can be generalized to [11] HJC,multi-level = h̵ωr (â†â + 1 2 ) +∑ j=0 h̵ωj ∣j⟩⟨j∣ +∑ j=0 h̵gj,j+1 (â∣j + 1⟩⟨j∣ + â†∣j⟩⟨j + 1∣) , (2.23) where the level coupling strengths are gj,j+1 ≈ √ j + 1g0,1. The detuning ∆j,j+1 between the resonator and the j ↔ j + 1 transition of the multi-level system is defined as ∆j,j+1 = (ωj+1 − ωj) − ωr. (2.24) 15 I typically operated the system in the dispersive limit where ∆j,j+1/gj,j+1 ≫ 1. In this limit, the following unitary transformation T ≈ exp(∑ j=0 gj,j+1 ∆j,j+1 (â†∣j⟩⟨j + 1∣ − â∣j + 1⟩⟨j∣)) (2.25) can be used to diagonalize the Hamiltonian. If I only keep the two lowest levels of the transmon, the unitary transformation of the Hamiltonian in Eq. 2.22 is THJCT † ≈Hdisp JC = h̵ωr (â†â + 1 2 ) + 1 2 h̵ωqσ̂z + h̵χâ†âσ̂z. (2.26) Here, χ = g2/∆ is the dispersive shift between the two-level system (transmon) and the readout resonator. The dispersive shift for the j ↔ j + 1 transition of the multi-level system is χj,j+1 ≃ g2j,j+1 ∆j,j+1 . (2.27) To distinguish from number states in the resonator, in this dissertation, I follow the tradition of the transmon community by using the notation ∣g⟩, ∣e⟩, and ∣f⟩ to denote the lowest three coupled transmon states. Here “g” stands for ground state, “e” stands for (first) excited state, and “f” is for second excited state. Considering multiple transmon levels, applying the unitary transformation on Eq. 2.23 results in THJC,multi-levelT † ≈Hdisp JC = h̵ω̃r (â†â + 1 2 ) + 1 2 h̵ω̃qσ̂z + h̵χâ†âσ̂z, (2.28) where χ = χge−χef/2, ω̃q = ωge+χge, and ω̃r = ωr−χef/2. Here χge and χef are defined by taking j = 0 and j = 1 in Eq. 2.27. Among these parameters, ωr/2π corresponds to the “bare” frequency of the resonator, which is the frequency of the resonator if it weren’t coupled to the qubit. The two frequencies (ω̃r ± χ)/2π are called “dressed” frequencies of the resonator, since they are the frequencies of the resonator coupled to the qubit. Fig. 2.5 shows an illustration of the relationship between these various frequencies, assuming ωq < ωr. 16 ωr ω̃r −χge −1 2χef −χ−χ Angular Frequency |S 2 1 | Figure 2.5: Schematics of angular resonant frequency for a notch resonator coupled to a transmon qubit. The solid black curve is the bare resonance at ωr. The two dressed resonances correspond to the qubit in the ∣g⟩ state (blue curve) and the qubit in the ∣e⟩ state (red curve) at ω̃r ± χ. 2.2.4 Measuring the State of the Transmon Eq. 2.28 shows that when the qubit is in an eigenstate of the σ̂z operator (with eigenvalues ±1), the third term causes the frequency of the resonator to differ by ±χ/2π. Thus by probing the resonator, the state of the qubit can be inferred. For example, Fig. 2.5 shows the transmission across a notch resonator that’s used for qubit state readout, which is the case for most of the qubits studied in this dissertation. If a microwave pulse at (ω̃r + ∣χ∣)/2π frequency is applied and low transmission is measured, that means the pulse is on resonance with the readout resonator and thus the qubit is in the ∣g⟩ state. If instead high transmission is measured, then the readout resonator’s frequency is not at (ω̃r + ∣χ∣)/2π and thus the qubit is not in the ∣g⟩ state. This method is called “low power dispersive readout” [13, 141]. In addition, for each additional photon in the resonator, the frequency of the qubit is shifted additionally by 2χ/2π. Thus by performing qubit spectroscopy, the number of photon stored in 17 the resonator can be accessed (see Sec. 4.6.6 for details). This “Low power dispersive readout” has the advantage of being quantum nonde- molition (QND) [12]. That is, the qubit eigenstates (∣g⟩, ∣e⟩) are preserved after the measurement. Another way to measure the state of the qubit is the “high power Jaynes Cum- mings readout”[15, 116]. To understand it, note the transformation from Eq. 2.23 into Eq. 2.28 is only valid if the number of photons stored in the resonator n satisfies n≪ ncrit = ∆2 ge 4g2ge , (2.29) where ncrit is the “critical photon number”. Otherwise for n ≳ ncrit and Eq. 2.28 breaks down, the frequency of the resonator eventually goes to ωr/2π, which differs from the low-photon case of (ω̃r ± χ)/2π (see Fig. 2.5). In the next subsection, I provide a detailed numerical analysis of the “high power Jaynes Cummings readout”. To understand it in simple terms, according to Fig. 2.5, if a microwave pulse of large enough power is applied at the bare resonator frequency ωr/2π, the dressed resonator frequency for qubit ∣e⟩ state at (ω̃r − ∣χ∣)/2π is closer to ωr/2π than the dressed frequency (ω̃r + ∣χ∣)/2π if the qubit is in ∣g⟩ state. This means during photon buildup in the resonator, the pulse detuning is smaller if the qubit is in ∣e⟩ state and thus it’s easier for photon number to reach ncrit and for the dispersive condition to break down. As a consequence, there is a particular range in power (usually 2-4 dB for my measurements) that a microwave pulse at ωr/2π frequency would create enough photons so that the resonator is at the bare frequency of ωr/2π, if the qubit is in ∣e⟩, in which case there would be low transmission (following the notch resonator example earlier); but the resonator would not be at ωr/2π frequency if the qubit is in ∣g⟩ state, in which case there would be high transmission. The difference in the transmission can be used to distinguish the state of the qubit. 18 The behavior is somewhat different if qubit frequency is higher than the resonator frequency, i.e. ωq > ωr. In this case, the dressed resonator frequency is closer to the bare resonator frequency if the qubit is in the ∣g⟩ state at low photon numbers. Nonetheless, as photon number builds up in the resonator, there would still be a range of photon numbers where this relationship flips, and qubit ∣e⟩ state corresponds to smaller detuning. See the program in the appendix of this dissertation for details. Since a high photon number is used with the “high power Jaynes Cummings readout”, it has the benefit of a relatively large signal-to-noise ratio. This is helpful if the first stage amplifier has a small quantum efficiency. Furthermore, it’s only sensitive to qubit decay during the photon buildup phase, which is helpful if T1 of the qubit is small. The most significant drawback is that it’s not a QND measurement as the qubit state is destroyed by hybridization with a large number of photons. In addition, I found for some of my transmons, this “high power Jaynes Cummings readout” generates quasiparticles in the transmon (see Sec. 2.2.6) and decreases its T1 (see Chapter 6). 2.2.5 Numerical Simulation of Jaynes Cummings Readout If the cosine term in Eq. 2.11 is expanded to order ϕ̂6 and constants are ignored, Eq. 2.23 can be rewritten as H = h̵ωra †a + √ 8EjEcb †b − Ec 12 (6 (b†b)2 + 6 (b†b)) + h̵g (ab† + a†b) + E2 c 45 √ 8EjEc (20 (b†b)3 + 30 (b†b)2 + 40 (b†b)) , (2.30) where a†(a) and b†(b) are the raising (and lowering) operators for the resonator and transmon respectively. The last term on the second line is from the expansion of order 19 ϕ̂6. I will use ∣α,β⟩ to represent a state with α excitations in the resonator and β excitations in the transmon. The n-excitation manifold for the lowest 5 energy levels of the transmon is spanned by the set of basis states (∣n − 4,4⟩, ∣n − 3,3⟩, ∣n − 2,2⟩, ∣n − 1,1⟩, ∣n − 0,0⟩)T . The Hamiltonian in Eq. 2.30 is then expressed in matrix form as ⎛ ⎜⎜⎜⎜⎜⎜⎜⎜ ⎝ (n−4)ωr+4ωj−10Ec+ 128 3 E2 c ωj g √ n−3 √ 4 g √ n−3 √ 4 (n−3)ωr+3ωj−6Ec+ 62 3 E2 c ωj g √ n−2 √ 3 g √ n−2 √ 3 (n−2)ωr+2ωj−3Ec+8 E2 c ωj g √ n−1 √ 2 g √ n−1 √ 2 (n−1)ωr+1ωj−1Ec+2 E2 c ωj g √ n g √ n nωr ⎞ ⎟⎟⎟⎟⎟⎟⎟⎟ ⎠ (2.31) whose eigenvalues can be solved numerically. Here I have defined ωj = √ 8EjEc and assumed h̵ = 1 for simplicity. In this way, the qubit frequency is approximated as ωq = (ωj −Ec)/2π. Considering the common case where the qubit frequency is smaller than the resonator frequency {ωj, ωq} < ωr, let vn,i be the i-th largest eigenvalue of Eq. 2.31 of the n-excitation manifold, then vn,1 would correspond to the energy levels of the transmon-resonator system when the transmon is in the ∣g⟩ state, and vn,2 for ∣e⟩ state. The transition frequency of the resonator between the n-photon and (n + 1)-photon states is f̃r(n) = (vn+i,i − vn+i−1,i) /2π. (2.32) Figure 2.6 shows the difference between the transition frequency of the resonator f̃r and its bare frequency ωr/2π versus photon number n, for the transmon in ∣g⟩ and ∣e⟩ states. The transition frequency f̃r was obtained by numerically solving Eq. 2.32, setting i = 1 for transmon ∣g⟩ state and i = 2 for transmon ∣e⟩ state. The nominal parameters were: bare resonator frequency ωr/2π = 8 GHz, qubit frequency ωq/2π = 5.8 GHz, charging energy Ec/h = 200 MHz, and resonator-qubit coupling g/2π = 60 MHz. At low photon numbers with the qubit in the ∣g⟩ state, the dressed 20 100 101 102 103 104 105 106 107 108 0 0.5 1 1.5 −χge −2χ ncrit n f̃ r − (ω r /2 π ) (M H z) Qubit in |g⟩ Qubit in |e⟩ Figure 2.6: Relative frequency of resonator vs stored photon number n for qubit in ∣g⟩ (blue) and ∣e⟩ (red) states. The dashed line represents critical photon numbers ncrit ≈ 336. The arrows and texts correspond to some of shifts in frequencies defined in Sec. 2.2.3 and shown in Fig. 2.5. resonator frequency is shifted from the bare resonator frequency by an amount ∣χge∣ ≃ g2/∣ωr − ωq ∣ ≈ 1.64 MHz(×2π). The difference in the dressed resonator frequency between the qubit in the ∣g⟩ state and ∣e⟩ state is given by ∣2χ∣, with ∣χ∣ = ∣χge−χef/2∣ ≈ g2 ×Ec/(ωr − ωq)/(ωr − ωq −Ec) ≈ 0.14 MHz(×2π). The quantities of χge and 2χ are represented by arrows in Fig. 2.6. As the photon number n in the resonator increases, the dispersive approximation breaks down, and the resonator frequency f̃r gradually shifts back to the bare frequency ωr/2π. For the high power Jaynes Cummings readout, a microwave pulse of power Pin is applied at the bare resonator frequency ωr/2π. Depending on the qubit state, such a pulse would result in a different average photon number in the resonator and different transmitted power Ptrans. Neglecting the transmon, the relationship between Pin and the average photon number n in a resonator is Pin = 2hQe π × (κ 2 4 + 4π2∆f 2)n, (2.33) 21 where κ is the photon relaxation rate of the resonator, given by κ = 2πf/QL, Q−1L = Q−1e + Q−1i where QL is the loaded quality factor, Qe is the external quality factor and Qi is the internal quality factor, and ∆f is frequency the detuning between the applied microwave pulse and the resonant frequency. Most of the readout resonators that I studied in this dissertation are over-coupled, i.e. the internal quality factor Qi was much larger than the external quality factor Qe, and thus QL ≃ Qe (see Table 4.1). For the following numerical simulations, I used Qe = 3 × 104,Qi = 3 × 105 and assumed these values did not depend on the number of photons n. As Eq. 2.33 shows, the average photon number n depends on ∆f and Pin. Fur- thermore, ∆f depends on qubit state and photon number n (see Fig. 2.6). Combining these, Fig. 2.7 shows a plot of photon number n versus microwave input power Pin applied at the bare resonator frequency, for the qubit in the ∣g⟩ and ∣e⟩ states. In Fig. 2.7, at the lowest and highest input power / photon number, there’s an approxi- mately linear relationship between these two parameters, which doesn’t have obvious dependence on the qubit state. However, at intermediate powers, a small increase in input power leads to a huge increase in the average photon number. The location in power where this occurs depends on qubit state. The ratio of the magnitude of the transmitted signal to the magnitude of the input signal is ∣S21∣ = ∣ Vt Vin ∣ = √ Pt Pin , (2.34) where V is the complex voltage, P stands for power. Fig. 2.8 illustrates the relation- ship between Vin and Vt for a notch resonator. The ratio of the minimum transmitted voltage Vt,min (which is achieved when ∆f = 0) to the input voltage Vin is ∣Vt,min/Vin∣ = QL/Qi. (2.35) As Fig. 2.8 shows, the actual transmitted voltage Vt lies on a circle whose diameter 22 10−15 10−14 10−13 10−12 10−11 10−10 10−9 102 104 106 108 Pin (Watt) n Qubit in |g⟩ Qubit in |e⟩ Figure 2.7: Stored resonator photon number n vs input power Pin for qubit in ∣g⟩ (blue) and ∣e⟩ (red) states, when applying a microwave pulse at bare resonator frequency ωr/2π. The same parameters as Fig. 2.6 were used for this simulation. VinVt,min Vt θ in-phase quadrature Figure 2.8: Relationship between the complex transmitted voltage Vt and the com- plex input voltage Vin, see main text for description. 23 connects Vt,min and Vin. In the limit ∆f/f̃r ≪ 1, the angle θ in Fig. 2.8 is given by tan θ = 4QL ∆f f̃r 1 − 4 (QL ∆f f̃r ) 2 , (2.36) where −π < θ < π, and that θ and ∆f have the same sign. From Fig. 2.8, it can be shown that ∣S21∣ = ¿ ÁÁÀ( QL 2Qc sin θ) 2 + (1 − QL 2Qc (cos θ + 1)) 2 . (2.37) θ is given by Eq. 2.36 and it depends on ∆f and f̃r, both of which can be obtained from Fig. 2.6 and 2.7. As a result, the ratio ∣S21∣ can be calculated as a function of input power. Figure 2.9 shows ∣S21∣ versus input power. At low input powers, the applied microwave signal at ωr is not on resonance with the dressed resonator, so a high transmission is observed. As the input power increases, the frequency of the readout resonator shifts to ωr (Fig. 2.6), reducing ∣S21∣ as the detuning between the applied microwave signal and resonator decreases. This transition happens at a lower input power when the qubit is in the ∣e⟩ state (see red curve in Fig. 2.7 and 2.9). The dashed line in Fig. 2.9 represents input power that gives maximum difference in S21 between the two qubit states. At this power, when the qubit is in the ∣e⟩ state, the microwave pulse is in resonance with the (notch) resonator, resulting in a small ∣S21∣, but when the qubit is in ∣g⟩, it is not in resonance, resulting in a larger ∣S21∣. It is this difference between the transmitted power that I used to distinguish the qubit states as high-power Jaynes Cummings readout. It’s also worth noting that the two curves diverge at roughly the same input power in Fig. 2.7 and Fig. 2.9. In Fig. 2.7, the range of photon numbers where the two curves diverge is roughly the same range of photon numbers in Fig.2.6 where the resonator transitions from dressed frequency to bare frequency. In addition, the lower end of 24 10−14 10−13 10−12 10−11 10−10 10−9 0 0.2 0.4 0.6 0.8 1 Pin (Watt) |S 21 | Qubit in |g⟩ Qubit in |e⟩ Figure 2.9: Simulated ∣S21∣ vs input power Pin for qubit in ∣g⟩ (blue) and ∣e⟩ (red) states, when applying a microwave pulse at bare resonator frequency ωr/2π. The same parameters as Fig. 2.6 and Fig. 2.7 were used for this simulation. The dashed line at 2.6 pW represents the power of maximum difference in S21 between the two qubit states, which is used for high-power Jaynes Cummings readout. this photon number range is close to the critical photon number ncrit = ∆2/4g2 ≈ 336 for the parameters I used in these simulations. While Fig. 2.9 shows ∣S21∣ versus input power for a microwave signal applied at the bare resonator frequency ωr/2π, the calculation can be extended to microwave signals applied at any frequency and power. For example, in Fig. 2.10 I show false color plots of ∣S21∣ vs frequency and input power, for the qubit in the ∣g⟩ and ∣e⟩ states. It’s worth noting this simple numerical model does not consider coupling to higher modes of the readout resonator, the transient behavior during photon buildup, the Poisson distribution of photon numbers, higher energy levels of the transmon which may have significant charge dispersion, and higher order expansions of the cosine term in Eq. 2.11. For these reasons, these simulation results appear somewhat differently compared to the experimentally measured spectra presented in later chapters (see 25 |g 7.999 8 8.001 8.002 8.003 Frequency (GHz) 10 -11 10 -12 10 -13 10 -14 10 -15 In p u t P o w e r (W ) |S 21 | |e 7.999 8 8.001 8.002 8.003 Frequency (GHz) 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 Figure 2.10: Simulated transmission ∣S21∣ of the readout resonator vs frequency and power, with the qubit in the ∣g⟩ (left) and the ∣e⟩ (right) state. The same parameters as Fig. 2.6 were used for simulation. Areas in white are not covered by simulation due to small photon numbers. At approximately 2 × 10−12 Watt input power and 8.00016GHz frequency, ∣S21∣ ∼ 1 for the ∣g⟩ state and ∣S21∣≪ 1 for the ∣e⟩ state, which could be used for high power Jaynes Cummings readout. In the transition region (around 10−13 and 10−12), the simulated S21 is very sensitive to input power and frequencies, due to the simplifications of the model discussed in the main text. Fig. 4.10 and 4.11). This is particularly evident in the transition between the low power dispersive region and the high power bare resonator region (around 10−13 and 10−12 W input power in Fig. 2.10). 2.2.6 Estimating the Voltage across the Junction While performing characterization on some of the transmons (e.g. qubit H-i, see Sec. 4.3 for naming nomenclature), I observed a reduced T1 (see Sec. 4.6.3 for mea- surement setup) when measuring the state using the high-power Jaynes Cummings readout, as compared to measuring the state with the low-power dispersive readout. I suspect that an AC voltage V was being generated across the Josephson junction during the readout process, such that V exceeds 2∆/e where ∆ is the the super- 26 Figure 2.11: Lumped-element circuit model to estimate the voltage across the Josephson junction. conducting gap of aluminum. Such a large voltage can generate quasiparticles (see Fig. 6.1) in the superconducting leads and thus reduces the qubit T1. Note that if “multiple Andreev reflection” is present, a voltage of 2∆/ne with integer n may also generate quasiparticles. [5] To estimate the AC voltage across the Josephson junction when a microwave signal was applied at the input port of the package, I used a linear lumped-element circuit model (see Fig. 2.11). In this model, Z0 = 50 Ω is the characteristic impedance of the transmission line and the distributed impedance of the CPW resonator. For this calculation, the Josephson junction is approximated as a linear inductor with inductance Lq, and the transmon qubit includes a shunting capacitance Cq = e2/2Ec. The (angular) resonant frequency of the qubit is ωq = 1/ √ Lq ×Cq. The LRC resonator is coupled to the transmon with a coupling capacitance Cg, with a value obtained either from finite 27 element simulations or the qubit-resonator coupling g by [133] g ≈ √ ωq CqZr Cg Cr = Cg √ ωqZr Cq ωr (2.38) after some approximations. Here ωr = 1/ √ Lr ×Cr is the angular frequency of the read- out resonator, and Zr ≈ √ Lr Cr is the characteristic impedance of the lumped element resonator. For a quarter-wave resonator, which is what I used in my experiments, Zr is related to the distributed impedance by [41] Zr = 4 π Z0. (2.39) The resistance Rr in the parallel RLC circuit of the readout resonator is related to the internal quality factor of the resonator by Rr = QiZr. (2.40) In addition, Cs is the coupling capacitance between the resonator and the transmission line. It is related to the external quality factor Qe of the resonator by Qe = 2 Z0 ×Zr × ω2 r ×C2 s . (2.41) Figure 2.9 shows the (simulated) resonator transmission ∣S21∣ when applying mi- crowave signal at bare resonator frequency ωr/2π. From this plot, I can determine the input power resulting in the maximum difference of S21 between qubit states ∣g⟩ and ∣e⟩ (dashed line in Fig. 2.9); this is the power one would use for readout. At this input power and assuming the qubit is in the ∣e⟩ state, the average number of photons stored in the readout resonator would be n = 1.7 × 105 for the parameters used in Fig. 2.7 and 2.9. In the lumped circuit model of Fig. 2.11, the relationship between the voltage across the readout resonator VRLC and photon number n is given by VRLC = √ n × h̵ × ω2 r ×Zr. (2.42) 28 This voltage is split between the coupling capacitor with impedance 1/iωrCg and the transmon with impedance (iωrCq +1/iωrLq)−1, so the voltage across the junction can be expressed as Vj = VRLC × (ωrCq − 1/ωrLq)−1 (ωrCq − 1/ωrLq)−1 + 1/ωrCg . (2.43) For the parameters used in Fig. 2.9 in Sec. 2.2.5, this corresponds to 55 µV -RMS across the junction for the high power Jaynes Cummings readout. Next, I fixed the parameters of the readout resonator from Sec. 2.2.5 (ωr/2π = 8 GHz, Qe = 3 × 104,Qi = 3 × 105), and Ec/h = 200 MHz of the qubit. I then varied the frequency of the qubit ωq/2π and the coupling g in the simulation. Fig. 2.12 shows how the estimated voltage across the junction during high power Jaynes Cummings readout depends on frequency detuning and coupling strength. As expected, a smaller detuning and larger g both yield a larger voltage across the junction. An important observation from Fig. 2.12 is that, if the frequency detuning between the resonator and the qubit exceeds some threshold (about 2.5 GHz), the voltage across the junction during the readout process only has a small dependence on detuning. To avoid quasiparticle generation when performing the high power Jaynes Cum- mings readout, the peak voltage (a factor of √ 2 of the RMS voltage plotted in Fig. 2.12) should be smaller than two times the superconducting gap. (This ig- nores the possibility of Andreev processes.) Given aluminum’s superconducting gap ∆Al ∼ 200 µeV , this puts an upper bound of g ∼ 140 MHz of coupling strength between the qubit and the readout resonator. I note that this discussion only involved the resonator and transmon lumped circuit model in Fig. 2.11, and not the input/output transmission line. In Sec. 6.1 of this dissertation, I discuss a similar way of estimating Vj by solving the entire circuit in Fig. 2.11 for a ratio of Vj/Vin. In addition to the limitations of the numerical model (outlined at the end of Sec. 2.2.5), quasiparticle generation at a Josephson junction 29 Estimated Voltage across junction ( V-RMS) 50 100 150 200 250 Coupling g/2 (MHz) 1 2 3 4 5 D e tu n in g ( r - q )/ 2 ( G H z ) 200 400 600 800 1000 1200 1400 1600 Figure 2.12: False color plot of numerically simulated (RMS) voltage across a Josephson junction during high power Jaynes Cummings readout, versus detun- ing between readout resonator and qubit (y-axis) and coupling g/2π between read- out resonator and qubit (x-axis), assuming qubit is in the ∣e⟩ state. The solid black line indicates the contour where the estimated voltage across the junction is 2∆Al/ √ 2 ∼ 283 µV . Beyond a certain detuning (∼ 2.5GHz), the voltage across the junction predominantly depends on coupling strength g. is a highly nonlinear process. Thus the model in this section is highly simplified. Nonetheless it sheds some light on the measurement of quasiparticle dynamics, to be discussed further in Chapter 6 of this dissertation. 2.2.7 Driving Transmon Transitions Eq. 2.28 describes the Hamiltonian of the dressed transmon-resonator system in the absence of an applied time-dependent microwave pulse. In practice, applied mi- crowaves resonant with the transmon are used to stimulate transitions and manipulate 30 its state. Suppose the qubit has an electric dipole moment operator given by [48] ˆ⃗ d = ⟨g∣d⃗∣e⟩ (∣g⟩⟨e∣ + ∣e⟩⟨g∣) = ⟨g∣d⃗∣e⟩ (σ− + σ+) , (2.44) and the electric field of the applied microwave pulse is given by E⃗(t) = ϵ⃗E0 cosωdt = 1 2 ϵ⃗E0 (eiωdt + e−iωdt) . (2.45) Here ωd/2π is the frequency of the microwave pulse, and E0 is the amplitude of the electric field (which could have a time dependent envelope). The interaction between the qubit and the oscillating electric field of microwave pulse can be expressed as Hint,q = − ˆ⃗d ⋅ E⃗(t) = − E0 2 ⟨g∣ϵ⃗ ⋅ d⃗∣e⟩ (σ− + σ+) (eiωdt + e−iωdt) (2.46) ≃ − E0 2 ⟨g∣ϵ⃗ ⋅ d⃗∣e⟩ (σ−eiωdt + σ+e−iωdt) , (2.47) where the rotating wave approximation (RWA) [59] was used to derive Eq. 2.47. Defining Ωq = − E0 h̵ ⟨g∣ϵ⃗ ⋅ d⃗∣e⟩ (2.48) as the Rabi frequency (a metric of the strength of the drive), the interaction Hamil- tonian can be written as Hint,q = h̵Ωq 2 (σ−eiωdt + σ+e−iωdt) . (2.49) Following a unitary transformation U = exp (iωdσz 2 t), one can remove the explicit time dependence in Eq. 2.49 to arrive at Hint,q = h̵Ωq 2 (σ− + σ+) . (2.50) In a similar manner, if a microwave pulse with frequency ωp/2π is applied and interacts with the resonator, one can change the Pauli operators to creation/annihi- lation operators of the resonator and transform Eq. 2.50 into Hint,r = h̵Ωr 2 (a + a†) . (2.51) 31 The combined Hamiltonian is thus H = h̵∆r (â†â + 1 2 ) + 1 2 h̵∆qσz + h̵χâ†âσz + h̵Ωr 2 (a + a†) + h̵Ωq 2 σx, (2.52) where ∆r = ω̃r−ωp is the detuning between the resonator and its drive, and ∆q = ω̃q−ωd is the detuning between the qubit and its drive. 2.2.8 Bloch Sphere Representation The Hamiltonian in Eq. 2.52 includes Pauli operators that act on the qubit. A pure state of an isolated qubit can be expressed in the form of ∣ψ⟩ = ⎛ ⎜⎜ ⎝ β α ⎞ ⎟⎟ ⎠ , ∣α∣2 + ∣β∣2 = 1, (2.53) or in Dirac notation as ∣ψ⟩ = cos θ 2 ∣g⟩ + eiϕ sin θ 2 ∣e⟩, (2.54) with these two expressions related by α = cos θ 2 , β = eiϕ sin θ 2 . (2.55) A Bloch sphere [14] is a powerful way to graphically represent the state of a qubit. The two angles (ϕ and θ in Eq. 2.54) that define the qubit state also define the location of the state on a Bloch sphere surface(see Fig. 2.13a). Assuming σj is a generalized Pauli matrix that operates on a two-level system with Rabi frequency Ω (Hamiltonian is thus H = h̵Ωσj/2), solving Schrödinger equation yields the evolution of the state ∣ψ(t)⟩ = exp (−iσj(Ωt)/2)∣ψ(0)⟩ = (I cos(Ωt/2) − iσj sin(Ωt/2)) ∣ψ(0)⟩. (2.56) The state evolution in Eq. 2.56 can be visually interpreted as rotation of state ∣ψ⟩ on the Bloch sphere, about an axis defined by the components of σj, and with rotational angular velocity Ω (see Fig. 2.13b and the following discussions). 32 Φ θ (a) (b) Figure 2.13: Bloch sphere representation of qubit state and rotation. (a) Bloch sphere with states labelled. I label the south pole of the sphere as ∣g⟩ state, the north pole as ∣e⟩ state, and the states on the equator are equal superposition of them with different phases. The green arrow is a sample state ( ∣g⟩√ 2 + (1+i)∣e⟩2 ) uniquely identified by two angles θ = π/2 and ϕ = π/4. (b) Evolution of qubit state starting from ∣g⟩ under Pauli operator σx (green) and operator (σx + 0.5σz) / √ 1.25 (red). The x, y and z axes are labelled. The sphere is rotated 90 degrees along z axis from (a) to better present the rotation of qubit states. The qubit states go through different rotations (see text) with the arrows indicating the axes of rotation. To illustrate the state rotation of a qubit on the Bloch sphere, assume the system described by Eq. 2.52 starts in the ground state, i.e. there are no photons in the resonator (⟨â†â⟩ = 0) and the qubit is in the ∣g⟩ state, and there is no resonator drive (Ωr = 0), but there’s a microwave drive on resonance with the qubit frequency (∆q = 0). The remaining (qubit) Hamiltonian in the rotating wave approximation is simply H = h̵Ωqσx/2. (2.57) The evolution of the qubit state is a rotation about the x-axis on the Bloch sphere (i.e. around the (∣g⟩ + ∣e⟩)/ √ 2 state), indicated by the green points in Fig. 2.13b. 33 Starting from ∣g⟩ state, after time π/Ωq the qubit is “flipped” and reaches the ∣e⟩ state. This is called a “π-pulse” as it rotates the state of the qubit on a Bloch Sphere by an angle of π about the x-axis. If the qubit drive is not on resonance with the qubit (∆q ≠ 0), the (qubit) Hamil- tonian would be H = h̵Ωqσx/2+ h̵∆qσz/2. The red points in Fig. 2.13b show the qubit state evolution starting from ∣g⟩ for the case ∆q/Ωq = 0.5. Here, the qubit state ro- tates about an axis (red arrow) that’s an angle of (arctan (∆q/Ωq))) from the x-axis. As a result of this detuning, the qubit never reaches the ∣e⟩ state. As a result, when driving qubit transitions I typically sent microwave pulses on resonance with the qubit frequency. In rare occasions when a small detuning was needed (such as Ramsey oscillation measurements), I made sure ∆q ≪ Ωq so that the transitions had good fidelity. On the other hand, due to the limited anharmonic- ity of the transmon, the microwave pulse could also couple to unwanted higher-level transitions of the transmon (e.g. ∣e⟩ ↔ ∣f⟩) which drive the transmon out of the computational basis. The ∣e⟩↔ ∣f⟩ transition differs from the desired ∣g⟩↔ ∣e⟩ tran- sition by the qubit anharmonicity ∣α∣. To decrease the spurious ∣e⟩↔ ∣f⟩ transition, I typically required Ωq ≪ ∣α∣. For example, for typical ∣α∣ ∼ 200 MHz, the maximum drive strength Ωq I used in my experiments was on the order of a few MHz. 2.2.9 Density Matrix Representation In addition to pure quantum states, which can be represented by Eq. 2.53, a more general quantum state can be represented by a density matrix operator ρ̂ as [94] ρ̂ =∑ i pi∣ψi⟩⟨ψi∣, (2.58) where ∣ψi⟩ are pure quantum states, and pi indicates the probability that the system is in state ∣ψi⟩. For a two level system such as a qubit, ρ̂ in general can be expressed 34 by a 2 × 2 matrix ρ̂ = ⎛ ⎜⎜ ⎝ ∣β∣2 η η∗ ∣α∣2 ⎞ ⎟⎟ ⎠ , (2.59) where both ∣α∣2 and ∣β∣2 are non-negative real numbers satisfying ∣α∣2 + ∣β∣2 = 1. In addition, tr (ρ̂ ⋅ ρ̂) ≤ 1, (2.60) where “tr” stands for trace as the summation of all diagonal terms of a matrix. Eq. 2.60 takes ‘=’ sign if ρ̂ is a pure state that can be expressed without the summation in Eq. 2.58. Eq. 2.60 takes ‘<’ sign if ρ̂ is a mixed state. [94] If Ô is an observable of the system, then its expected value can be expressed as ⟨Ô⟩ = tr (Ôρ̂) . (2.61) In particular, all Pauli matrices are physical observables, and thus can have expec- tation values ⟨σi⟩, i ∈ {x, y, z}. Since the Bloch sphere in Fig. 2.13 has unit radius, a general qubit state ρ̂ can be represented by coordinates (⟨σx⟩, ⟨σy⟩, ⟨σz⟩). This point lies on the surface of the Bloch sphere if it corresponds to a pure state, and inside the Bloch sphere if it’s a mixed state. For an isolated system described by Hamiltonian H, the density matrix evolves in time according to ih̵ ∂ρ̂ ∂t = −[ρ̂,H] =Hρ̂ − ρ̂H. (2.62) 2.2.10 Lindblad-Kossakowski Master Equation and Decoherence There are many ways for a quantum system to be in a mixed state, such as being the result of a partial trace from a larger Hilbert Space. For my qubits, a mixed 35 quantum state often stems from decoherence. While Eq. 2.62 describes the density matrix evolution of an isolated quantum system, a typical quantum system would interact with the surrounding environment. For example, a qubit may be coupled to and exchange excitation with a heat bath. In general, it’s not possible to keep track of coherent evolution of the environment. Instead, the effects of the intersection are treated as decoherence of the qubit system. Assuming the environment can be treated as Markovian [79], that is, it retains no memory of its interaction with the smaller system, then the evolution of the small system can be described by the Lindblad-Kossakowski master equation [69, 73] ih̵ ∂ρ̂ ∂t = −[ρ̂,H] +L[ρ̂]. (2.63) Compared to Eq. 2.62, there is an additional term L[ρ̂], which is called a Liouvillian and is given by L[ρ̂]/h̵ =∑ i ΓiD[Âi]ρ̂ =∑ i Γi (Âiρ̂† i − 1 2 † i Âiρ̂ − 1 2 ρ̂† i Âi) . (2.64) In Eq. 2.64, the summation is over the decoherence channels, each with decoherence rate Γi, and Âi is the “Lindblad operator” of each decoherence channel [89]. One decoherence channel for a qubit is energy relaxation, with a Lindblad operator given by σ− = ⎛ ⎜⎜ ⎝ 0 0 1 0 ⎞ ⎟⎟ ⎠ . (2.65) Using the matrix notation of Eq. 2.59, one can see D[σ−]ρ̂ = ⎛ ⎜⎜ ⎝ −∣β∣2 −0.5η −0.5η∗ ∣β∣2 ⎞ ⎟⎟ ⎠ , (2.66) which can be interpreted as transferring population from the upper-left cell of ρ̂ (i.e. ∣e⟩ state) to the lower-right cell (i.e. ∣g⟩ state), while proportionally reducing the diagonal terms. This is an incoherent process. 36 Figure 2.14a illustrates qubit energy relaxation in a graphical way on a Bloch sphere, where the qubit decays from ∣e⟩ to ∣g⟩. The blue points represent sampling of density matrix ρ̂ with time interval 0.1/Γ−, where Γ− is the decay rate for the Liouvillian. It can be shown that the qubit’s population in ∣e⟩ (P∣e⟩) decays in an exponential fashion P∣e⟩ ∼ exp (−Γ−t). In addition, a qubit may be excited by its environment, with Lindblad operator σ+ and rate Γ+, although usually Γ+ ≪ Γ−. Adding these two rates gives the energy relaxation rate, the inverse of which is a qubit’s energy relaxation time constant T1, T1 = 1 Γ+ + Γ− ≈ 1 Γ− . (2.67) Another typical decoherence channel for a qubit is pure dephasing, with a Lindblad operator given by σz = ⎛ ⎜⎜ ⎝ 1 0 0 −1 ⎞ ⎟⎟ ⎠ (2.68) such that D[σz]ρ̂ = ⎛ ⎜⎜ ⎝ 0 −2η −2η∗ 0 ⎞ ⎟⎟ ⎠ . (2.69) This can be interpreted as removing phase information, which is encoded in the off- diagonal terms of ρ̂, while keeping the “population” of ∣g⟩ and ∣e⟩ constant. Figure 2.14b illustrates qubit dephasing in a graphical way on the Bloch sphere, where qubit state evolves from (∣g⟩ + ∣e⟩) / √ 2 to an evenly mixed state of ∣g⟩ and ∣e⟩. The blue points represent sampling of density matrix ρ̂ with time interval 0.1/Γz, where Γz/2 is the dephasing rate for the Liouvillian (due to the extra factor of 2 in Eq. 2.69). It can be shown that the qubit’s population evolves in an exponential faction towards the central z-axis in the Bloch Sphere, whose time constant is defined as the pure dephasing time Tϕ = Γ−1z . (2.70) 37 (a) (b) Figure 2.14: Bloch sphere representation of qubit decoherence, (a) qubit energy relaxation from ∣e⟩ to ∣g⟩ and (b) qubit pure phase relaxation from (∣g⟩ + ∣e⟩) / √ 2 to an evenly mixed state of ∣g⟩ and ∣e⟩. 2.2.11 Determining Optimal Readout Length As discussed in Sec. 2.2.4, the “low power dispersive readout” is a common way to measure the state of the qubit. This technuqie uses the dispersive frequency shift of the readout resonator due to changes in the qubit state. Due to vacuum fluctuations and noise in the amplifier chain, the readout signal from the resonator inherently contains some noise. If the qubit has infinite T1, then a longer readout pulse would give better signal-to-noise ratio: for a readout pulse of length t ≪ T1, the ‘signal’ is proportional to t while the (averaged) ‘noise’ is proportional to √ t. On the other hand, a qubit only has a finite lifetime T1, as a result, at time τ < t during the readout pulse, the qubit only has probability exp (−τ/T1) of being in the original state, scaling the ‘signal’ by a similar factor. Most of the measurements discussed in this dissertation are averaged over many repetitions. As a result, I typically aimed for a readout length t that was a balance between these two aspects, giving me the best 38 signal-to-noise ratio of the averaged data (instead of single-shot readout fidelity). To simplify the readout process, I ignore resonator ringup and ringdown for the moment. In addition, for my measurements, microwave signals throughout the read- out length t were integrated using equal weights. To quantitatively study the optimal readout length t, suppose ν is the ‘signal’ per unit time, characterized by the (voltage) difference in the transmitted signal between the qubit ∣g⟩ and ∣e⟩ state (See Fig. 4.32 for a sample of single-shot readout data in the IQ plane). Considering time τ < t where t is the readout pulse length, during the small time interval from τ to τ + dτ , the qubit only has probability exp (−τ/T1) of being in the original state, thus the contribution to the overall ‘signal’ from this small time interval is exp (−τ/T1)dτ × ν. The overall readout ‘signal’ is an integration of such term signal = ∫ t 0 exp (−τ/T1)dτ × ν = T1ν (1 − exp (−t/T1)) . (2.71) On the other hand, suppose µ is the ‘noise’ per square root of unit time, then the readout ‘noise’ is thus µ √ t. This corresponds to the spread of single-shot readout clusters on the IQ plane in Fig. 4.32. In this way, signal-to-noise ratio (SNR) of the readout can be expressed as SNR = ∣T1ν (1 − exp (−t/T1)) µ √ t ∣ 2 (2.72) The SNR in Eq. 2.72 is maximized respect to readout length t by setting d(SNR)/dt = 0, which is satisfied when exp (−t/T1) − (1 − exp (−t/T1)) × T1 2t = 0. (2.73) One finds at t ≈ 1.26T1. It’s worth noting this optimal readout length does not depend on either µ or ν, but only on the qubit’s T1. It’s worth noting that discussions in this section correspond to my measurements where microwave signals throughout readout length t were integrated using equal 39 readout length t Signal Non-Signal 0 1 P e Figure 2.15: Figurative representation of SNR with respect to readout length t. The readout ‘signal’ is proportional to the area in red, which, due to qubit energy relaxation has an exponential shape with time constant T1. The readout ‘noise’ is proportional to √ t. weights. It’s possible to assign different weights to different portions of the signal, e.g. assign more weights to the initial part of the signal where the qubit is more likely to be in its original state, and less weight to the later part where the qubit likely has experienced decay. In other words, the the ideal length of the “box car” filter function was discussed in this section where the filter function could be of any shape, which could result in higher SNR than discussed in this section. [40] 2.3 Loss Mechanisms of Transmon 2.3.1 Quasiparticle Loss In a superconductor, electrons with opposite spins and momenta form Cooper pairs [22]. A Bogoliubov quasiparticle, or quasiparticle for short, is an elementary excitation above the ground state of a superconductor, which can be formed as a 40 result of broken Cooper pairs. Assuming a symmetric (equal) superconducting gap ∆ on two sides of an ideal Josephson junction, it can be shown that due to single quasiparticle tunneling, the transition rate from qubit state i to (a different) qubit state f can be expressed as [43, 20] Γif = 16Ej h̵π∆ ∫ ∞ ∆ dϵf(ϵ) (1 − f(ϵ + h̵ωif)) ϵ(ϵ + h̵ωif) +∆2 √ ϵ2 −∆2 √ (ϵ + h̵ωif)2 −∆2 ∣⟨f ∣ sin ϕ̂ 2 ∣i⟩∣ 2 (2.74) where ϕ̂ is the operator for the difference of superconducting phase across the junction, f is the quasiparticle distribution function, h̵ωif = Ei − Ef is the energy difference between the initial and final states. This equation assumes the low-energy limit ∣h̵ωif ∣≪∆ and ignores Andreev processes [5]. For fixed-frequency transmons with no external flux bias (ϕ0 = 0), assuming i = ∣e⟩ and f = ∣g⟩, it can be shown that ∣⟨f ∣ sin ϕ̂ 2 ∣i⟩∣ 2 = Ec h̵ωeg 1 + cosϕ0 2 = Ec h̵ωeg . (2.75) Assuming the quasiparticles have a distribution function f that follows Fermi-Dirac distribution at an effective temperature T that’s much smaller than the supercon- ducting gap (T ≪∆/kB), it can be shown that Γeg ≃ √ 2ωeg∆ π2h̵ xqp, (2.76) where xqp is the normalized quasiparticle density, defined by xqp = nqp/ncp, the ratio between absolute quasiparticle volume density nqp and Cooper pair density, which is ncp = 4 × 106µm−3 for aluminum (when T ≪ ∆/kB). At low temperatures and in thermal equilibrium, xqp is related to T by [19] xqp ≃ √ 2πkBT /∆exp (−∆/kBT ) . (2.77) Assuming T ≃ 20 mK for the temperature of the dilution fridge, ∆ = 200 µeV as the superconducting gap of aluminum, one would expect xqp ∼ 10−52. However, ex- 41 perimentally measured quasiparticle densities are in the range of xqp = 10−8 to 10−4 [4, 124, 81, 140, 142, 31]. Possible sources of such excess quasiparticle densities are the absorption of stray infrared radiation [9] and ionizing radiation such as muons and gamma rays [139, 85, 17]. 2.3.2 Two-Level System Loss Dielectric two-level system (TLS) defects are a another source of loss for super- conducting qubits and resonators. A simple model for a TLS defect is a charged ion that can tunnel between two asymmetrical potential wells (see Fig. 2.16) [103, 32]. I assume the minima of energy of the two potential wells differ by ϵ, and the tunneling between the two wells has a strength ∆, which is related to the distance d between the wells, the tunneling barrier height V and the effective mass of the TLS m, such that ∆∝ exp (−d √ 2mV /h̵2). The energy difference between the ground and excited state of the TLS is then E = √ ϵ2 +∆2. If the detuning between a TLS and a qubit is small and the coupling between them is significant, excitations can be exchanged between the TLS and the qubit, leading to the qubit’s energy relaxation and dephasing. Although the exact microscopic identity of TLS causing decoherence in superconducting qubits is not completely understood [92], possible candidates are atomic defects such as O-H bonds [82, 123], chemical residuals [110], oxygen defects [1, 72] among others[92]. Given that TLS loss often occurs in dielectric materials and interfaces, on a macro- scopic level, this yields a loss tangent tan δi. A common characteristic of TLS loss in superconducting resonators is that the loss decreases with increasing applied electric field. On a microscopic level, once a TLS is driven into a mixed state with equal probability, it first has to dissipate its energy before it can absorb more energy from 42 Figure 2.16: Double-well potential model of TLS, reproduced from Ref. [127] under license number 5574400977925. the external field. In this way, the TLS limiting Qi of a resonator can be expressed as [104] 1 Qi,TLS =∑ i pi tan δi tanh ( h̵ωr 2kBT ) √ 1 + ( n nc )α (2.78) where i indicates different loss regions, pi are their participation ratios (ratio of electric field energy in that region compared to total electric field energy), tan δi are loss tangents characterizing TLS loss of each region, ωr is the resonant frequency of the resonator, nc is a critical photon number characterizing the saturation of TLS loss, and n is the photon number stored in the resonator. For a single TLS under uniform field, α = 1 [104]. From Eq. 2.78, one can see that as n increases, the loss from TLS decreases. In Fig. 5.3 of this dissertation, the measured Qi versus photon number is plotted for a series of coplaner waveguide resonators, most of which displayed an increase of Qi with increasing photon number n, a trend consistent with TLS loss. While the loss for superconducting resonators can be reduced by saturating the TLS, transmon qubits typically operate in the single-excitation condition (n ≤ 1 in Eq. 2.78), thus TLS loss can be a significant loss mechanism for transmon qubits. 43 2.3.3 Purcell Loss Discovered by Edward M. Purcell, the Purcell effect [109] refers to enhancement (or suppression) of the spontaneous emission rate of an excited atom or quantum emitter that is coupled to a resonant optical cavity or nanostructure. For the case of a transmon coupled to a readout resonator with resonator decay rate κ, the qubit can decay due to such coupling, with a decay rate given by [114] ΓPurcell ≃ (g/∆)2κ, (2.79) where g is the coupling strength between the resonator and the qubit, and ∆ is the detuning between the two. Eq. 2.79 is only valid in the dispersive limit g ≪∆. Loss from the Purcell effect can be reduced by decreasing κ (i.e. increasing QL) or by reducing the effective resonator-qubit coupling (g/∆). Reducing κ increases the ring-up time of the resonator, preventing a fast measurement of the qubit states while reducing (g/∆) reduces the signal-to-noise ratio of qubit state measurement. In practice, the qubit may also couple to higher modes of a resonator structure, which introduces additional Purcell loss. For transmons, additional modes are most obvious when the qubit frequency is above the fundamental frequency of the readout resonator. For the qubits that I studied (see Table 4.1), the qubit frequencies were sometimes significantly different than the design value, due to variations in the device fabrication. However, the qubit-resonator detuning ∆ was typically sufficiently greater than the coupling g so that most of the qubits’ coherence was not limited by the Purcell effect. Using qubit H-i as an example, κ ∼ 1.3 × 106 s−1, g/2π = 157 MHz, ∆ = 2.05 GHz, the Purcell limited qubit T1 is 130 µs, much larger than the observed value of 18 µs. One way to mitigate Purcell loss without affecting measurement speed is the use of a Purcell filter [122], which is an additional low quality-factor resonator between 44 the transmission line and the readout resonator. However, fast and high-fidelity measurements of the qubit state were not a focus of this dissertation. As a result, I did not incorporate Purcell filters into the designs of my qubits. 2.3.4 Package Loss My 2D superconducting devices were mounted in a package that could introduce additional loss. For example, the qubit could couple to spurious low-Q resonance modes [119, 56] of the package. This can be understood as another case of Purcell loss discussed in the previous subsection. A common way to mitigate such loss is to increase the minimum frequencies of spurious modes of the package, typically above 10 GHz for conventional superconducting transmon qubits. Reducing the dipole moment of transmon qubits and thus reducing coupling to spurious modes is another way to mitigate this loss. In addition, if normal metal is used in the package, it can cause loss due to eddy currents. This loss mechanism is discussed in Chapter 5. 2.3.5 Vortex Loss When a magnetic field is applied to a superconductor, screening currents flow on the surface of the superconductor so as to cancel the magnetic field inside the super- conductor. If a superconducting sample is cooled through its critical temperature Tc, any applied magnetic field will be expelled. This is called the Meissner effect [88]. For type-I superconductors (which include aluminum), such expulsion of magnetic fields happens up to a critical field strength Hc. If H > Hc, the material becomes normal. For type-II superconductors, such expulsion is only up to a lower critical