Novel Entangling Gates and Scalable Trap Designs for Trapped-Ion Quantum Computing by ALEXANDER D. QUINN A dissertation accepted and approved in partial fulfillment of the requirements for the degree of Doctor of Philosophy in Physics Dissertation Committee: Hailin Wang Chair David Allcock Advisor & Core member Jens Nöckel Core Member James Prell Institutional Representative University of Oregon Summer 2024 © 2024 Alexander D. Quinn All rights reserved. 2 DISSERTATION ABSTRACT Alexander D. Quinn Doctor of Philosophy in Physics Title: Novel Entangling Gates and Scalable Trap Designs for Trapped-Ion Quantum Computing Trapped ions have received much attention as a platform for quantum computing, a purpose they may be well-suited for on account of their natural features: Ions of the same species are inherently identical; they can be manipulated using electric and magnetic fields via both their net charge, which allows external fields to couple to ions’ center-of-mass motion, and via internal electronic transitions; and when they are isolated from the environment, their states can remain coherent for long time spans, at least by the standards of quantum information experiments. Despite these features, the ability to carry out useful quantum computing in trapped ions is limited, with two major practical constraints being 1) the difficulty of coherently controlling ions’ states and 2) the limited physical scale of existing trapped-ion quantum computers, which typically hold no more than tens of ions. This document present a set of projects meant to help address these limitations through different tracks, including the development and testing of control techniques for trapped ions and the design of new types of traps. Firstly, we present an entangling gate carried out on quantum bits (qubits) encoded in a set of electronic energy levels that have been relatively unexplored until recently and whose viability for quantum information processing may enable more efficient architectures for trapped ion quantum computing. Specifically, we entangled a pair of qubits encoded in two Zeeman sublevels of the D5/2 metastable excited state of a pair of trapped 40Ca+ ions using Raman laser 3 beams 10s of THz detuned from resonance to limit scattering rates (a fundamental error source in Raman gates). We demonstrate that high-fidelity gates (98.6(1)% subtracting state preparation and measurement error, and 99.1(1)% subtracting both SPAM and erasure) can be performed with this encoding scheme, show that the main source of error is technical noise, and employ a leakage detection scheme that allows decay or deshelving from the metastable level to be heralded, potentially making correcting this class of errors easier. After this, we shift focus from control techniques and discuss scalable trap design, focusing on a project to fabricate ion traps using 3D printing, a technique that could potentially enable microfabricated traps with high harmonicity, power efficiency, and depth of confinement relative to the 2D (planar) microfabricated traps widely used in efforts to scale up trapped-ion quantum computing. We design, simulate, fabricate, and carry out preliminary electrical testing on metallized trap prints, demonstrating some minimum viability of the technique, and show computationally that traps produced this way could have trapping characteristics similar to those of other 3D microfabricated designs. Finally, we consider scalable trap design for continuous- variable quantum computing (CVQC), a quantum computing scheme where, in trapped ions, information is encoded in the states of the vibrational modes of an ion crystal. A key requirement for universal CVQC is the ability to perform non- Gaussian operations, which can be difficult to carry out electronically. In this work, we present a 2D ion trap design that can perform non-Gaussian motional operations in an all-electronic way, with the design process accounting for the geometric limitations imposed by 2D traps. We consider some of the motional operations possible with this trap and estimate their associated coupling rates, finding that under a reasonable set of assumptions about operating parameters, coupling rates for non-Gaussian operations could be achieved that are comparable 4 to those previously achieved with all-electronic Gaussian operations could likely be achieved with this trap design. This dissertation includes co-authored, published material. 5 ACKNOWLEDGEMENTS None of the work presented here would have been possible without the independent contributions (e.g. the fabrication/assembly and testing of subsystems of the experiment apparatus as the lab was first being built up), the collaboration (e.g. in the slow-converging feedback loop between simulation and experiment), and the patient guidance of the many members, past and present, of the Oregon Ions lab and those the lab has worked with. 6 TABLE OF CONTENTS Chapter Page I. INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . 27 1.1. A conceptual overview of this document . . . . . . . . . . . . . 27 1.2. Trapped ions - a brief guide . . . . . . . . . . . . . . . . . . . 29 1.2.1. Traps . . . . . . . . . . . . . . . . . . . . . . . . . 29 1.2.2. Ions . . . . . . . . . . . . . . . . . . . . . . . . . . 31 1.3. Quantum computing . . . . . . . . . . . . . . . . . . . . . . 32 1.3.1. Building blocks for QC . . . . . . . . . . . . . . . . . 33 1.3.2. Technical requirements . . . . . . . . . . . . . . . . . 36 1.4. TIQC architecture . . . . . . . . . . . . . . . . . . . . . . . 37 1.5. The omg architecture . . . . . . . . . . . . . . . . . . . . . 40 1.5.1. Optical, metastable, and ground qubits . . . . . . . . . . 40 1.5.2. Using omg qubits together . . . . . . . . . . . . . . . . 41 1.6. Continuous-variable quantum computing . . . . . . . . . . . . . 42 1.6.1. Building blocks for CVQC . . . . . . . . . . . . . . . . 42 1.6.2. Implementing CVQC in trapped ions . . . . . . . . . . . 43 1.7. Scalable ion trap design . . . . . . . . . . . . . . . . . . . . 44 1.7.1. The QCCD architecture and surface electrode traps. . . . . . 45 II. EXPERIMENTAL APPARATUS . . . . . . . . . . . . . . . . . 47 2.1. Ion trap . . . . . . . . . . . . . . . . . . . . . . . . . . . 47 2.1.1. Trap design . . . . . . . . . . . . . . . . . . . . . . 47 2.1.2. Trapping RF drive . . . . . . . . . . . . . . . . . . . 48 2.1.3. Trapping DC potential . . . . . . . . . . . . . . . . . . 50 7 Chapter Page 2.1.4. Shim potentials . . . . . . . . . . . . . . . . . . . . . 51 2.1.5. UHV system . . . . . . . . . . . . . . . . . . . . . . 51 2.2. Magnetic field control . . . . . . . . . . . . . . . . . . . . . 54 2.2.1. Magnetic field coils . . . . . . . . . . . . . . . . . . . 54 2.2.2. Line noise mitigation . . . . . . . . . . . . . . . . . . 55 2.3. Imaging and state detection . . . . . . . . . . . . . . . . . . 57 2.3.1. Imaging system overview . . . . . . . . . . . . . . . . . 58 2.3.2. Imaging system performance . . . . . . . . . . . . . . . 60 2.4. Laser system . . . . . . . . . . . . . . . . . . . . . . . . . 61 2.4.1. Rack-mounted laser boards . . . . . . . . . . . . . . . . 62 2.4.2. Laser frequency locking . . . . . . . . . . . . . . . . . 65 2.4.3. AOM boards . . . . . . . . . . . . . . . . . . . . . . 67 2.5. Far-detuned lasers . . . . . . . . . . . . . . . . . . . . . . . 69 2.5.1. 976 nm Raman lasers . . . . . . . . . . . . . . . . . . 70 2.5.2. 854 nm AC Stark shift laser . . . . . . . . . . . . . . . 72 2.6. Beam delivery . . . . . . . . . . . . . . . . . . . . . . . . 72 2.6.1. Beam pointing . . . . . . . . . . . . . . . . . . . . . 75 2.6.2. Beam focusing . . . . . . . . . . . . . . . . . . . . . 76 2.6.3. Beam power control . . . . . . . . . . . . . . . . . . . 78 2.6.4. Polarization control . . . . . . . . . . . . . . . . . . . 80 2.7. Electronic control fields . . . . . . . . . . . . . . . . . . . . 81 2.7.1. Magnetic Rabi drive . . . . . . . . . . . . . . . . . . . 81 2.7.2. Displacement drive . . . . . . . . . . . . . . . . . . . 82 2.7.3. Squeeze/beamsplitter drives . . . . . . . . . . . . . . . 83 2.7.4. Pulse shaping . . . . . . . . . . . . . . . . . . . . . 83 8 Chapter Page 2.8. Experiment control system . . . . . . . . . . . . . . . . . . . 86 2.8.1. Control hardware . . . . . . . . . . . . . . . . . . . . 86 2.8.2. Clocking . . . . . . . . . . . . . . . . . . . . . . . . 86 2.8.3. DDSes and phase-coherent RF control . . . . . . . . . . . 87 2.8.4. Experimental control codebase . . . . . . . . . . . . . . 89 2.9. Ion loading . . . . . . . . . . . . . . . . . . . . . . . . . . 91 2.9.1. Calcium oven . . . . . . . . . . . . . . . . . . . . . . 92 2.9.2. Neutral calcium detection . . . . . . . . . . . . . . . . 92 2.9.3. Photoionization . . . . . . . . . . . . . . . . . . . . . 94 III. EXPERIMENTAL METHODS . . . . . . . . . . . . . . . . . . 97 3.1. Trapping . . . . . . . . . . . . . . . . . . . . . . . . . . . 98 3.2. Cooling . . . . . . . . . . . . . . . . . . . . . . . . . . . 101 3.2.1. Doppler cooling . . . . . . . . . . . . . . . . . . . . . 102 3.2.2. EIT cooling . . . . . . . . . . . . . . . . . . . . . . 104 3.2.3. Pulsed sideband cooling . . . . . . . . . . . . . . . . . 105 3.2.4. Cooling testing with thermometry . . . . . . . . . . . . . 106 3.2.5. Micromotion compensation . . . . . . . . . . . . . . . . 107 3.3. m qubit preparation . . . . . . . . . . . . . . . . . . . . . . 109 3.3.1. Optical pumping . . . . . . . . . . . . . . . . . . . . 109 3.3.2. Preparation fidelity and limits . . . . . . . . . . . . . . 110 3.3.3. Isolating qubit states . . . . . . . . . . . . . . . . . . 112 3.4. State detection . . . . . . . . . . . . . . . . . . . . . . . . 114 3.4.1. State-selective deshelving . . . . . . . . . . . . . . . . 114 3.4.2. Distinguishing shelved and unshelved ions - Fluorescence . . . 115 3.5. Coherent control - Spin . . . . . . . . . . . . . . . . . . . . 116 9 Chapter Page 3.5.1. RF magnetic Rabi flopping . . . . . . . . . . . . . . . . 117 3.5.2. Raman Rabi flopping . . . . . . . . . . . . . . . . . . 121 3.5.3. Pulse types and coherent errors . . . . . . . . . . . . . . 123 3.5.4. Composite pulses . . . . . . . . . . . . . . . . . . . . 124 3.5.5. AC Stark shifts . . . . . . . . . . . . . . . . . . . . . 127 3.6. Coherent control - Motion . . . . . . . . . . . . . . . . . . . 128 3.6.1. General operations . . . . . . . . . . . . . . . . . . . 130 3.6.2. Displacement . . . . . . . . . . . . . . . . . . . . . . 131 3.6.3. Single-mode squeezing and phase shifting . . . . . . . . . 131 3.6.4. Two-mode squeezing and beamsplitter . . . . . . . . . . . 132 3.7. Coherent control - Spin-motion coupling . . . . . . . . . . . . . 133 3.7.1. Raman sidebands . . . . . . . . . . . . . . . . . . . . 133 3.7.2. Spin-dependent forces . . . . . . . . . . . . . . . . . . 136 3.8. Motional state characterization . . . . . . . . . . . . . . . . . 138 3.9. Coherence time measurement . . . . . . . . . . . . . . . . . . 140 3.9.1. Ramsey experiment . . . . . . . . . . . . . . . . . . . 140 3.9.2. Ramsey with spin echoes . . . . . . . . . . . . . . . . . 141 3.9.3. Fock state Ramsey experiment . . . . . . . . . . . . . . 143 3.10. Routine calibration techniques . . . . . . . . . . . . . . . . . 143 3.10.1. Coherent drive calibrations . . . . . . . . . . . . . . . . 143 3.10.2. Light shift calibrations . . . . . . . . . . . . . . . . . . 145 3.10.3. Optical frequency calibrations . . . . . . . . . . . . . . 146 3.10.4. Two-point calibrations . . . . . . . . . . . . . . . . . . 148 IV. RAMAN GATES IN METASTABLE QUBITS . . . . . . . . . . . 150 4.1. Erasure errors in m qubits . . . . . . . . . . . . . . . . . . . 151 10 Chapter Page 4.1.1. Erasure conversion with fluorescence . . . . . . . . . . . . 152 4.1.2. Leakage pathways in m qubits . . . . . . . . . . . . . . 153 4.2. Geometric phase gates . . . . . . . . . . . . . . . . . . . . . 154 4.2.1. Fidelity estimation . . . . . . . . . . . . . . . . . . . 155 4.3. Gate errors . . . . . . . . . . . . . . . . . . . . . . . . . . 158 4.3.1. Error budget . . . . . . . . . . . . . . . . . . . . . . 158 4.3.2. Addressing dominant error sources . . . . . . . . . . . . 159 4.3.3. Errors from leakage detection . . . . . . . . . . . . . . . 161 4.3.4. Outlook . . . . . . . . . . . . . . . . . . . . . . . . 164 V. 3D-PRINTED ION TRAPS . . . . . . . . . . . . . . . . . . . . 166 5.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . 167 5.2. Quantifying Trap Performance . . . . . . . . . . . . . . . . . 169 5.3. Trench Trap Geometries . . . . . . . . . . . . . . . . . . . . 172 5.3.1. Simple Trench . . . . . . . . . . . . . . . . . . . . . 174 5.3.2. Stacked Trench . . . . . . . . . . . . . . . . . . . . . 175 5.3.3. Comparison Geometries - SET and Wafer Traps . . . . . . 176 5.3.4. Summary . . . . . . . . . . . . . . . . . . . . . . . 177 5.4. 3D Printing . . . . . . . . . . . . . . . . . . . . . . . . . 178 5.4.1. Dielectric Material . . . . . . . . . . . . . . . . . . . 179 5.4.2. Initial Fabrication . . . . . . . . . . . . . . . . . . . . 180 5.4.2.1. Print . . . . . . . . . . . . . . . . . . . . . 180 5.4.2.2. Metalization . . . . . . . . . . . . . . . . . . 182 5.4.2.3. Integrated Optics . . . . . . . . . . . . . . . . 182 5.5. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . 182 11 Chapter Page VI. SURFACE ELECTRODE TRAPS FOR CVQC . . . . . . . . . . . 189 6.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . 189 6.2. Single ion in a cubic potential . . . . . . . . . . . . . . . . . 190 6.2.1. Hamiltonian . . . . . . . . . . . . . . . . . . . . . . 190 6.2.2. The ωp = 0 case (cubic phase gate) . . . . . . . . . . . . 191 6.2.3. The ωp = ωx case . . . . . . . . . . . . . . . . . . . . 192 6.2.4. The ωp = 3ωx case (trisqueezing) . . . . . . . . . . . . . 192 6.3. Two ions in a cubic potential . . . . . . . . . . . . . . . . . . 193 6.3.1. Hamiltonian . . . . . . . . . . . . . . . . . . . . . . 195 6.3.2. The ωp = 0 case . . . . . . . . . . . . . . . . . . . . 198 6.3.3. The ωp = ωc case . . . . . . . . . . . . . . . . . . . . 199 6.3.4. The ωp = ωc + 2ωr case . . . . . . . . . . . . . . . . . 200 6.3.5. The ωp = ωc − 2ωr case (parametric oscillation) . . . . . . . 200 6.3.6. The ωp = 3ωc and ωp = 3ωr cases (trisqueezing) . . . . . . . 200 6.4. Trap design for generating cubic potentials . . . . . . . . . . . . 201 6.4.1. Multipole expansion - the hexapole potential . . . . . . . . 201 6.4.2. Surface-electrode design . . . . . . . . . . . . . . . . . 203 6.4.2.1. Equivalent cylindrical electrode systems . . . . . . 203 6.4.2.2. Designing electrode systems to produce hexapoles . . 203 6.4.2.3. Integrating hexapole electrodes into surface trap . . . . . . . . . . . . . . . . . . 205 6.4.3. Axial cubic potentials - design limitations . . . . . . . . . 209 6.5. Practical implementation . . . . . . . . . . . . . . . . . . . . 209 6.5.1. Trisqueezing . . . . . . . . . . . . . . . . . . . . . . 209 6.5.2. Parametric oscillation . . . . . . . . . . . . . . . . . . 211 6.6. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . 212 12 Chapter Page VII. CONCLUSION . . . . . . . . . . . . . . . . . . . . . . . . . 213 APPENDIX: GATE ERROR CALCULATIONS . . . . . . . . . . . . . 214 A.1. Error budget . . . . . . . . . . . . . . . . . . . . . . . . . 214 REFERENCES CITED . . . . . . . . . . . . . . . . . . . . . . . . 221 13 LIST OF FIGURES Figure Page 1.1. A schematic illustration of the energy levels of an ion in a trap, showing two independent modes: A harmonic oscillator mode (left) and the internal electronic state of the ion (right). Line thicknesses in the right plot correspond to the number of Zeeman/hyperfine states in a given energy level. Solid arrows denote dipole transitions, while dashed arrows denote quadrupoles. Splittings between energy levels are not to scale. . . . . . . . . . . . . . . . . . . . . . . . . . 32 1.2. Examples of commonplace one- and two-qubit gates written in circuit notation (with each line representing an individual qubit over time) and the operation of the gate on a specific input state. a) The Hadamard gate, which rotates a single qubit. b) The controlled-NOT (CNOT) gate, which carries out a bit flip on the target qubit conditional on the state of the control qubit, generating entanglement. . . . . . . . . . . . . . 35 1.3. An illustration of the conceptual reasoning for using mixed- species ion chains. a) A pair of ions sharing a potential well can be entangled by driving both ions with laser fields. b) For carrying out auxilliary functions (e.g. sympathetic cooling of the ions being used for gates), additional ions need to be added to the chain. If these ions are of the same species as those used for gates, when driven with laser fields, they will emit photons that the other ions can absorb, destroying the information they contain. c) If the auxilliary functions are performed using ions of a different species with different optical transition frequencies, then photons scattered by one species will not be absorbed by the other. . . . . . . 38 1.4. The three qubit types considered by the omg architecture. . . . . . . . 41 14 Figure Page 1.5. A cartoon illustration of a possible QCCD “unit cell,” which can be tiled to produce arbitrarily large trap networks. Dark blue represents DC electrodes, while pink represents trapping RF. Ions are shown in light blue. The illustration on the left shows the unit cell of a 2D/surface electrode implementation of a QCCD architecture, while the one on the right shows how the concept can be extended to three- dimensional traps. . . . . . . . . . . . . . . . . . . . . . . . . 46 2.1. Photographs of our trap, showing (left) the trap attached to a vacuum flange and connected to electrical feedthroughs and (middle) a close-up shot of the unmounted trap, with the rough location of the trapping zone marked with a blue box. The associated inset in a picture taken with a CMOS camera of a five-ion chain, with ion spacing marked. Finally, a COMSOL render (right) shows trapping RF electrodes marked in red. . . . . . . . . . . . . . . . . . . . . . . . . . . 48 2.2. Schematic illustration of the hardware setup used for estimating the thickness of a layer of electroplated metal by measuring the total charge passed through the system during electroplating. . . . . . . . . . . . . . . . . . . . . . . . 49 2.3. Block diagram of the RF chain used to supply the RF trapping voltage. The trapping RF itself is supplied by the chain in red. Quadrupole motional drives (discussed further in Section 2.7.3 are combined with the trapping RF at the resonator. . . 51 2.4. Circuit diagram for the electrical connections to the DC trap electrodes (including AC drives applied to these electrodes), with endcap needles in blue, DC trapping electrodes in white, and compensation electrodes in light yellow. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52 2.5. a) CAD rendering of the vacuum chamber, associated support hardware, and components directly attached to this hardware, including magnetic field coils (Section 2.2.1 and b) mounts for beam delivery optics (Section 2.6). . . . . . . . . . . . 53 15 Figure Page 2.6. a) A schematic of the magnetic field coil arrangement used to set the quantization axis in the trap. Current is marked with white arrows, with arrow thickness corresponding the current magnitude. b) A photograph showing the five magnetic field coils used on the trap. Worth noting are the two smaller coils (white and blue) wrapped around the top left coil, which are used for line noise feedforward (Section 2.2.2). . . . . 55 2.7. Schematic illustrations of our two line noise mitigation schemes, explained in greater detail in text. . . . . . . . . . . . . . 56 2.8. Schematic illustration (and photograph below) of our imaging system, with light emitted by the ion (a) being captured by an objective (b), passed through a slit (c) to reduce stray light and a narrow-band filter (d) to prevent the system from registering light far away from the wavelength of the cycling transition of Ca+, collimated with a lens (f) in an adjustable barrel mount to be redirected by a flip-mounted mirror (g), which can direct light to a focusing lens leading to the PMT (h) and (i) or to a focusing lens (j) and (k) heading to the camera (l). . . . . . . . . . 59 2.9. Schematic of our overall laser system (exluding far-detuned lasers, discussed in Section 2.5) shown in terms of the output of a single laser breadboard. Free-space laser beams are shown with red and blue lines. Fiber optics are shown with solid black lines. Data connections are shown with dashed black lines. . . . . . . . . . . . . . . . . . . . . . . . . 62 2.10. Photograph of our laser rack, showing the drawers containing laser breadboards as well as an additional rack for housing laser controllers and a third for the wavemeter and OSA board. Inset shows a sample laser breadboard mounted in a drawer. . . . . . . . . . . . . . . . . . . . . . . . 64 2.11. Schematic illustrations of the three main AOM configurations we use in the lab. These schematics omit the focusing/recollimating lenses that focus the beams into the AOMs. . . . . . . . . . . . . . . . . . . . . . . . . . . . 69 16 Figure Page 2.12. Schematics of the two 976 nm Raman laser setups we have used in our experiments, showing the diode laser sources (marked in dark red) all the way through to the ion (marked in light blue). These schematics omit details of the beam delivery optics. . . . . . . . . . . . . . . . . . . . . . . . 70 2.13. Schematic of the laser system used to supply our high-power 854 nm beam used for generating AC Stark shifts on the non-qubit levels of the D5/2 manifold. . . . . . . . . . . . . . . . . 72 2.14. Schematic of the optics leading from the output collimators on our optics table to our trap. Definitions of abbreviations are as follows - PD: photodiode, GTP: Glan-Thompson polarizer, PM: piezo motorized mirror, PS: piezo stack mirror, GP: Glan polarizer. . . . . . . . . . . . . . . . . . . . . 74 2.15. A photograph with labels of one of our outcoupler bases, with the front plate of the photodiode box removed to show internal hardware. . . . . . . . . . . . . . . . . . . . . . . . . 79 2.16. Schematic summary of the RF voltages applied to the trap for coherent driving, marking the trap electrodes (shown in cross-section) that they are connected to. . . . . . . . . . . . . . . 82 2.17. Block diagram showing the RF chain for three motional drives used in our experiments. PS: pulse shaper, BS: beamsplitter. . . . 84 2.18. Schematic illustration of our two pulse-shaping schemes. a) Pulse shaping by multiplying a fixed-amplitude RF pulse with a nearly-trapezoidal DC envelope. b) Pulse shaping by running the DDS in RAM mode and generating arbitrarily shaped pulses, with the DC output on the AWG being unchanged. . . . 85 2.19. Schematic illustration of our experiment control system. a) Block diagram showing how our experiment control hardware (blue) interfaces with other blocks of our experiment, including beam control (red), the trap/imagining system (yellow), and the host PC allowing users to interface with the system (purple). b) A schematic of the beam power control system we use (SU-Servo). . . . . . . . . . 87 17 Figure Page 2.20. Class inheritance diagram showing the hierarchy of classes in our in our experiment codebase, down to the level of code for running individual experiments (e.g. an experiment for driving Rabi flopping on a qubit, RabiFlopping). NDScan classes are shown in yellow, in-house support classes are shown in pink, and experiments are shown in blue. The inset in purple shows the sequence for a generic scan point. In both portions, the main block of code that differs between individual experiments is shown in dark blue. . . . . . . . . 90 2.21. a) Illustration of the relevant photoionization geometry, showing the neutral calcium beam coming from the oven and the PI beam relative to the trap (electrodes shown in black). The neutral calcium and PI beams are both in the plane of the page. b) Energy level diagram of the two- photon PI process. . . . . . . . . . . . . . . . . . . . . . . . . 91 2.22. a) Fluorescence spectrum data (black dots) on the 423 nm line for our neutral calcium beam with a manual fit including three calcium isotopes (individual isotopes shown with dashed colored lines, total fluorescence shown with a solid black line). b) Relative probability of loading 43Ca+ at different 423 nm beam detunings and with different amounts of Doppler broadening. . . . . . . . . . . . . . . . . . . . . . . 94 3.1. A level diagram showing a Ca+ ion trapped in a harmonic oscillator potential (only one dimension of which is shown). Transitions between internal electronic energy levels are marked with the wavelengths and polarizations that we use to drive them. Energy levels are not to scale. . . . . . . . . . . . . 98 3.2. A schematic illustration of how the RF electrodes of a linear Paul trap generate a harmonic effective potential. . . . . . . . . . . 101 3.3. A set of diagrams showing how a) Doppler cooling, b) EIT cooling, and c) pulsed sideband cooling are implemented in our setup. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102 18 Figure Page 3.4. Schematic illustrations of techniques that we use to minimize micromotion. The offset direction each technique allows us to optimize is marked with a black double-headed arrow. The direction of the micromotion is marked with a double-headed red arrow. These techniques are a) minimization of position change with RF confinement change, b) minimization of Doppler broadening when driving with the 397π beam (blue arrow), and c) minimization of the micromotion sideband associated with our two-beam (red arrows) Raman Rabi drive. . . . . . . . . . . . . 108 3.5. a) The generic sequence for characterizing the effect of some operation (“Expt.”) on some known state, performing repeated experimental “shots” and calculating the probability of different outcomes. b) The specific pulse sequences of each experimental step (not showing cooling, which either happens before or is interleaved with state preparation). . . 110 3.6. Sample Rabi flopping, driven with a resonant RF magnetic field, showing two-level sinusoidal dynamics when the 854LS beam is on (blue) and precession through the D5/2 manifold before returning to the starting state when the 854LS beam is off (red). . . . . . . . . . . . . . . . . . . . . . 113 3.7. A sample count histogram for two-ion fluorescence detection, with the “zero bright” (blue), “one bright” (red), and “two bright” (green) Poisson distributions labelled above. . . . . . 116 3.8. Schematic illustration of our RF magnetic Rabi drive, showing trap electrodes (pink for trap RF, blue for Rabi drive current), the quantization axis magnetic field B⃗ (in the xz plane), and the magnetic Rabi drive B⃗drive (in the xy plane). . . . 120 3.9. Schematic illustration of the dynamics of a Λ system, showing a) the relationship between the three-level dynamics of the actual system and the effective two-level dynamics and b) the three-level system we use in our experiments. . . . 122 3.10. The results of experiments testing composite a) π/2 and b) π pulses. Different colors denote time scans at different ϕBB1, showing that a value of ϕBB1 can be selected such that the output of the pulse sequence is made insensitive to change in pulse times. . . . . . . . . . . . . . . . . . . . . . . . 126 19 Figure Page 3.11. A reference table for the motional operations performed in our experiments, showing in the phase space picture a) the Wigner function of the motional ground state, b the effect of a displacement operation, c) the effect of squeezing, d) the effect of a phase shift on a displaced state, e) the effect of a beamsplitter operation on a pair of modes (with the Wigner function of the second mode shown in red), and f) the effect of two-mode squeezing on the correlations between the quadratures of a pair of modes. . . . . . . . . . . . . . . . . . 130 3.12. Two techniques of achieving spin-motion coupling. a) Raman sideband transitions, showing the ∆n = +1 (BSB) and ∆n = −1 (RSB) transitions, with the spin states being represented as a pair of harmonic oscillator energy ladders separated by the spin state splitting. b) Spin-dependent forces, showing how two interfering beams with the same polarizations will generate periodic intensity gradients of different amplitudes (and different force magnitudes) on the two spin states. . . . . . . . . . . . . . . . . . . . . . . . . . 134 3.13. Pulse sequences for Ramsey sequences for characterizing spin (left) and motion (right) coherence, with (bottom) and without (top) a spin echo pulse. Carrier pulses are shown in dark blue, blue sideband pulses in light blue. To generate fringes at a given t, the phase ϕ of the final pulse can be scanned. . . . 142 3.14. Experiments for measuring the AC Stark shifts caused by a given beam (red). a) Carrying out Rabi spectroscopy while the beam is on and comparing this result to the result when the beam is off. b) Performing a Ramsey sequence with a spin echo, with the beam only being on during one delay so as to cause asymmetrical accumulation of phase. . . . . . . . . . . . 146 3.15. Pulse sequences (with time running from left to right) of our four laser line calibrations, with the state of the ion at the end of the experiment (during a fluorescence check, FC) shown assuming that laser associated with the line being characterized is far off-resonance. . . . . . . . . . . . . . . . . . . 147 3.16. Illustration of the two-point calibration servo loop. . . . . . . . . . . 148 20 Figure Page 4.1. Schematic illustration of our erasure error detection scheme in the context of a gate, showing how checks for 397 fluorescence can be used both to confirm successful state preparation prior to running a gate and check for leakage errors after a gate. . . . . . . . . . . . . . . . . . . . . . . . . 153 4.2. a) Gate pulse sequence showing beams relevant to the gate, with SDF phase (0, π) denoted by color (white, blue). b) A scan over SDF detuning from the LF OOP mode resonance, showing the crossing point in |↓↓⟩ and |↑↑⟩ populations at which the gate was performed. Inset shows high-shot time series data at the gate operating point. c) A sample parity fringe, with insets showing high-shot time-series data at the parity fringe extrema. . . . . . . . . . . . . . . . . . . . . . . . 157 4.3. a) A flowchart illustrating the possible outcomes of a one-ion gate sequence with an FC carried out at the end, considering only leakage errors. Outcomes corresponding to an erasure error are labelled in yellow, while outcomes corresponding to a missed leakage error that is missed are labelled in pink. b) An accounting of the error probabilities in each category with and without leakage checks. . . . . . . . . . . 163 5.1. Cross-section of a trap fabricated using our proposed method of 3D printing an electrodes onto a wafer containing electrode routing and some of the potential integrated ion control elements that have been demonstrated. . . . . . . . . . . . . 168 5.2. Cross-sectional views of of the trap designs simulated. The ions’ location is shown as an orange dot and the trap axis is out of the page. DC electrodes are blue, RF electrodes are red. Numerical apertures are shown in lavender. . . . . . . . . . . . 173 21 Figure Page 5.3. Simulated parameters of surface-electrode traps vs the symmetric and anti-symmetric ‘simple trench’ trench geometries. (a) Trap depth at a fixed radial frequency. (b) The quadrupole component (C2) of the trapping potential at the ion. (c) The hexapole component (C ′ 3) of the trapping potential. (d) The octopole component (C ′ 4) of the trapping potential. (e) Electrode dimensions at a constant ion-electrode separation for the symmetric simple trench trap. (f) Electrode dimensions at a constant ion- electrode separation for the anti-symmetric simple trench trap. (g,h) NA above and below the ion. Parameters are plotted against a generic geometric variable w, which is related to an electrode dimension of each trap as shown in the legend. Simulation results are shown with markers. The representative SET, wafer, and simple trench traps whose characteristics are summarized in Table 4 are highlighted with larger markers. Simulation points are connected with dashed lines only as a guide to eye. . . . . . . . . . . . . . . . . . 184 5.4. Simulated parameters of surface-electrode traps and wafer traps vs the symmetric ‘stacked trench’ trap geometry. (a) Trap depth at a fixed radial frequency. (b) The quadrupole component (C2) of the trapping potential at the ion. (c) The hexapole component (C ′ 3) of the trapping potential. (d) The octopole component (C ′ 4) of the trapping potential. (e) The hexapole component of the trapping potential, excluding the SET data to make the other series more clear. (f) Electrode dimensions at a constant ion-electrode separation for the symmetric stacked trench trap. (g,h) NA above and below the ion. Parameters are plotted against a generic geometric variable w, which is related to an electrode dimension of each trap as shown in the legend. Simulation results are shown with markers. The representative wafer and symmetric stacked trench traps whose characteristics are summarized in Table 4 are highlighted with larger markers. Simulation points are connected with dashed lines only as a guide to eye. . . . . . . . . . . 185 22 Figure Page 5.5. Simulated parameters of surface-electrode traps and wafer traps vs the anti-symmetric ‘stacked trench’ trap geometry. (a) Trap depth at a fixed radial frequency. (b) The quadrupole component (C2) of the trapping potential at the ion. (c) The hexapole component (C ′ 3) of the trapping potential. (d) The octopole component (C ′ 4) of the trapping potential. (e) The hexapole component of the trapping potential, excluding the SET data to make the other series more clear. (f) Electrode dimensions at a constant ion- electrode separation for the anti-symmetric stacked trench trap. (g,h) NA above and below the ion. Parameters are plotted against a generic geometric variable w, which is related to an electrode dimension of each trap as shown in the legend. Simulation results are shown with markers. The representative wafer and anti-symmetric stacked trench traps whose characteristics are summarized in Table 4 are highlighted with larger markers. Simulation points are connected with dashed lines only as a guide to eye. . . . . . . . . . . 186 5.6. SEM images of gold coated ormocer trap structures printed on glass. In this structure, many trenches have been connected up with X-junctions, as is envisaged in a large ‘QCCD’ processor (Blakestad et al. (2009)). (a) An overhead view of a grid of symmetric stacked trenches showing segmentation of the top DC electrodes. (b) An oblique view of the same grid. (c) A close up view of an intersection of trenches. (d) A false-color version of (c) showing DC electrodes in blue and RF electrodes in red. (e) A close-up view of a DC electrode segment. . . . . . . . . . . . . . 187 5.7. Proposed integration of electrode metalization with printing process shown here on a short section of the larger trap shown in Fig 5.6. (a) Silicon wafer with electrode routing and vias up to the surface is fabricated using standard CMOS process (metal shown copper coloured, SiO2 dielectric shown transparent). (b) Aluminum is deposited and patterned to define gaps between electrodes. (c) DLW used to print ormocer trap structures (blue). (d) Gold is deposition from multiple angles to coat all surfaces of the print. (e) Aluminium is etched to allow liftoff of the gold over the gaps between electrodes. Insets shown the liftoff process in more detail. . . . . . . . . . . . . . . . . . . . . . . 188 23 Figure Page 6.1. Plot of the Fock state distribution of the state generated by applying the ωp = ωx operation to the motional ground state for a time t = 0.23 × 2π/g. Components n > 9 not shown. Inset - Associated Wigner function. . . . . . . . . . . . . . 193 6.2. Plot of the Fock state distribution of of the trisqueezed state, obtained by applying the trisqueezing Hamiltonian to the ground state for a time t = 0.035 × 2π/g. Components n > 9 not shown. Inset - Associated Wigner function. . . . . . . . . . 194 6.3. Left - Illustration of a two-ion chain with coordinates labelled. Equilibrium positions are marked in light blue. Right - Table illustrating two-ion vibrational normal modes. . . . . . . 195 6.4. Trapping potential generated by symmetrical surface- electrode trap around ion (marked in yellow). The decomposition of this potential into overlaid quadrupole, hexapole, and octopole terms are represented schematically. . . . . . . 202 6.5. Cross-section of an asymmetric surface-electrode trap through plane perpendicular to trap axis, with ion position marked by a yellow dot. Cross-section of equivalent cylindrical electrode system (partitioned ring of radius r0) is overlaid on top, with the relationship between surface and cylindrical electrode positions r⃗(θ) = −r0êx + r0 tan ( θ 2 ) êy illustrated geometrically. Figure adapted from Wesenberg (Wesenberg (2008) . . . . . . . . . . . . . . . . . . . . . . . . 204 6.6. a). Quadrupole system. b). Hexapole system. c). Combined electrode system for generating quadrupole trapping potential and hexapole trisqueezing potential. Red - Quadrupole electrodes. Purple - Hexapole electrodes. Green - Inner ground. Blue - Outer ground. Notably, to produce optimal performance under geometric constraints, θ1 + θ2 = 60◦, and the quadrupole electrodes are adjacent to hexapole electrodes. . . . . . . . . . . . . . . . . . . . . . . . 207 6.7. a). C2 for the quadrupole electrodes and C3 for the hexapole electrodes plotted against θ1 when θ1 + θ2 = 60◦. b). Widths of different electrodes in the combined electrode system of Fig. 6.6b, normalized to ion height, under the condition θ1 + θ2 = 60◦. . . . . . . . . . . . . . . . . . . . . . . 208 24 Figure Page A.1. a) Motional Ramsey experiment, plotting the contrast between the |0⟩ and |1⟩ fock states as a function of the delay time between sideband π/2 pulses. b) Numerically simulated Bell-state infidelity due to a white motional noise source as a function of the coherence to gate time ratio. Orange lines represent the infidelity expected for measured coherence and gate times used in this work. Dashed lines represent the 68% confidence interval. . . . . . . . . . . . . . . . 215 A.2. a) Bell-state infidelity corresponding to a normally distributed static offset in gate mode frequency with standard deviation σ for different orders of Walsh modulation. b) Distribution of Bell-state infidelities corresponding to a normal distribution of mode frequency offsets with a standard deviation of 100 Hz for a Walsh 1 4-loop gate operating at 11.6 kHz. Dashed orange lines represent 68% confidence interval. . . . . . . . . . . . . . . . . . 217 A.3. a) Bell-state infidelity corresponding to a normally distributed static offset in gate mode frequency with standard deviation σ for different orders of Walsh modulation. b) Distribution of Bell-state infidelities corresponding to a normal distribution of mode frequency offsets with a standard deviation of 100 Hz for a Walsh 1 4-loop gate operating at 11.6 kHz. Dashed orange lines represent 68% confidence interval. . . . . . . . . . . . . . . . . . 218 A.4. a) Population scattered out of the m = +5/2 and +3/2 in the D5/2 manifold with either the 976 Rpi or Rsig beam on. b) Population scattered out of the qubit states with the 854 beam on. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220 25 LIST OF TABLES Table Page 1. A summary of projects worked on over the course of the last five years and which areas of trapped-ion quantum computing they are relevant to. My contribution to the projects in the top two rows was on the experimental side (hardware design/construction, writing/running/debugging experiments), while my contribution to the projects on the bottom was on the theoretical/computational side. . . . . . . . . . . 29 2. A summary of the components of our control system and what we use them for in our experiments. . . . . . . . . . . . . . . 88 3. Dominant sources of two-qubit gate infidelity as predicted by our error model. The lower half of the table summarizes the expected contributions due to erasure errors. Values in bold represent infidelity as measured in the gate . . . . . . . . . . . . . . 165 4. Summary of results for trap parameter calculations for the SET, simple trench trap (‘Simple’), stacked trench trap (‘Stacked’), and wafer trap. Values shown correspond to a set of geometric parameters selected to balance trap depth and quadrupole strength/purity. . . . . . . . . . . . . . . . . . . 177 5. The dimensions (in microns) of the traps whose parameters are summarized in Table 4. . . . . . . . . . . . . . . . . . . . . . 177 6. Electrical and thermal properties of ormocer. . . . . . . . . . . . . . 179 A.7. Measured scattering rates out of D5/2 (in Hz) when initialized in the m = +3/2 or m = +5/2 qubit states for various beams used in the gate. Measurements for the 854 LS beam were taken with two trapped ions. . . . . . . . . . . . . 219 26 CHAPTER I INTRODUCTION 1.1 A conceptual overview of this document This thesis describes a set of projects, largely independent of each other in terms of the work involved and the specific goals pursued, that may seem disconnected from one another. However, an important common thread runs between them: They are all ultimately focused on developing tools to use trapped ions in technological applications, whether quantum sensing or trapped-ion quantum computing (TIQC). In that light, these projects can be organized according to how they relate to the broader project of refining, expanding the capabilities of, and scaling up TIQC. There are two “axes” along which these projects can be organized: The type of TIQC they work to enable and the type of improvement they seek to make. Specifically, these projects can be in support of discrete-variable quantum computing (i.e. quantum computing using two-level Hilbert spaces or “qubits,” which will typically just be called “quantum computing” in the rest of the text) or of continuous-variable quantum computing (CVQC, i.e. quantum computing using infinite Hilbert spaces of evenly-spaced energy states). They also broadly fall under two parallel paths of development: the development and testing of new techniques for controlling the states of trapped ions (mostly carried out on the experimental side), and the design of novel trap geometries (carried out on the theoretical/computational side). These projects are listed specifically in Table 1. Not listed in the table is a fifth project investigating non- Hermitian quantum systems using our qubit platform. While this specific project was less motivated by technological applications, it demonstrates features of non- Hermitian dynamics with potential applications (e.g. the ability to implement fast and deterministic operations) that are worth discussion. 27 Projects not covered in detail. Of these five projects, two of them (observing of non-Hermitian dynamics in trapped-ion qubits Quinn et al. (2023) and carrying out trapped-ion motional interferometry with phase sensitivity below the standard quantum limit J. Metzner et al. (2024)) are not covered in detail in this thesis. Accounts of these projects can be found in their respective papers and in the thesis of Dr. Jeremy Metzner (see J. M. Metzner (2024)), which discusses them in greater detail (particularly on the theory side). Much of the content laid out in this introduction and in Chapters II and III still lays out the experimental apparatus and methods used to carry out these projects. Currently unpublished projects. Our project to perform entangling gates in metastable qubits, covered in Chapter IV, is soon to be published. My role in this project was on the experimental side, developing and debugging experiments and setting up hardware (though two very important hardware systems enabling this specific project, a high power injection-locked diode laser system for performing gates 2.5.1 and a patched version of our experimental control system 2.8.3, were built and tested by my colleagues Sean Brudney and Gabe Gregory respectively). The work of simulating our gates and calculating an error budget was performed by Gabe Gregory. Important context for this project is provided by a set of photon scattering rate measurements, taken by me and analyzed and presented in the thesis of Dr. Daniel Moore (see Moore et al. (2023)). An entirely separate theoretical/computational project, developing a surface-electrode trap for carrying out non-Gaussian operations on the motional modes of trapped ions for CVQC applications, was carried out by me alone, with guidance from my advisor David Allcock. Projects covered in existing publications. Chapter V is based on a paper currently available as a preprint, ”Geometries and fabrication methods for 3D 28 Discrete-variable quantum computing Continuous-variable quantum computing (CVQC) Control techniques Single- and two- qubit Raman gates in metastable (m) qubits Coherent control of ion motion for motional interferometry Trap design Design of 3D-printable ion traps for achieving scalability and favorable geometries Design of surface- electrode traps (SETs) for carrying out non- Gaussian operations on ion motion Table 1. A summary of projects worked on over the course of the last five years and which areas of trapped-ion quantum computing they are relevant to. My contribution to the projects in the top two rows was on the experimental side (hardware design/construction, writing/running/debugging experiments), while my contribution to the projects on the bottom was on the theoretical/computational side. printing ion traps” (Quinn, Brown, Gardner, and Allcock (2022)). The portions of this paper focusing on theory and simulation were written by me, with David Allcock writing portions of the intro and conclusion and Morgan Brown writing the description of the 3D printing and metallization process. Morgan Brown and David Miller carried out the fabrication and characterization of test prints, and I simulated and analyzed the designs discussed in this chapter. 1.2 Trapped ions - a brief guide Before discussing their technological applications in quantum information, a short discussion is offered below of what trapped ions are, how they can be modelled, and how, in broad strokes, they can be controlled. These topics, particularly as they relate to our own experiments, are discussed more thoroughly in Chapter III. 1.2.1 Traps. Because they have a net electric charge, ion experience a Lorentz force F⃗ = q(E⃗ + v⃗ × B⃗) and this force can be used to confine ions in 29 an entirely classical way. There are two specific types of traps widely in use for quantum information experiments: Penning traps and Paul traps. Penning traps. Penning traps confine ions using a combination of static electric and magnetic fields. A pair of endcaps held at a static voltage offer confinement in one direction while a magnetic field parallel to the endcap axis forces ions to undergo cyclotron motion, providing confinement in the other two directions. The ions’ continuous cyclotron motion (and the large magnetic fields required to confine them) makes them more difficult to work with in certain ways, and Penning traps are less widely used in research on scalable trapped-ion quantum computing than Paul traps (though there has been work done on microfabricated, scalable Penning traps. See Jain et al. (2024) among others). Paul traps. Paul traps confine ions using only electric fields. By Earnshaw’s theorem, it is impossible to achieve confinement in more than two directions at once using static electric fields (an immediate consequence of the fact that ∇·E⃗ = 0 in the absence of charge). Paul traps work around this constraint by using dynamic electric fields instead of static ones. A deeper treatment of this is given in Section 3.1. In short though, if an AC multipole potential is applied to a charged particle, in certain parameter regimes, the time-averaged motion of the particle shows the particle experiencing an effective potential (or “pseudopotential”) ϕeff proportional to the energy density of the AC electric field (i.e. ϕeff ∝ |E⃗|2). For a “linear” Paul trap (shown schematically in Figure 3.2), the trap electrodes are very long relative to the spacing between them, and they generate a 2D potential in the trapping zone. This potential only confines in two directions, with confinement in the third direction being supplied by a DC electric field, as in Penning traps. Most Paul traps used in trapped-ion work generate, to lowest-order, an AC quadrupole potential, which in turn corresponds to a harmonic pseudopotential. An ion 30 trapped in this 2D harmonic pseudopotential and the 1D harmonic potential supplied by the DC electric field (which tends to be anti-trapping in the other two directions) is, effectively, a 3D harmonic oscillator. This feature is important for much of the work done in trapped ions. 1.2.2 Ions. A neutral atom (or molecule) that has had one (or more) electrons removed (or added) is an ion. For all discussions below, “ion” will refer to a more narrow subset of this category: alkaline earth atoms that are singly-ionized, such that they have a single valence electron in a hydrogen-like potential. This accounts for most of the species used in trapped-ion experiments (though other elements like ytterbium have been used in experiments requiring the control and detection of individual ions (Balzer et al. (2006)), as have molecules (Drewsen, Mortensen, Martinussen, Staanum, and Sørensen (2004))). Electronic level structure. The generic energy level structure of an alkaline earth ion is shown in Figure 1.1. The relevant energy levels are the electronic ground state S1/2, the short-lived excited states P1/2 and P3/2, and the D3/2 and D5/2 excited states (which are long-lived in most ion species). The S and D levels can both couple to the P levels (where other selection rules are not violated) through electric dipole interactions. On the other hand, transitions between the D levels (also called metastable levels) and the S level are dipole-forbidden and can only take place through electric quadrupole interactions. Zeeman/hyperfine splitting. Within an energy level (e.g. within S1/2), the valence electron of the ion can occupy different angular momentum states based on the orbital angular momentum and its own inherent spin. Uncoupled to other systems, these states would be degenerate in energy. However, applying an external magnetic field will lift this degeneracy via the Zeeman effect, defining a quantization axis (i.e. an axis along which the angular momentum state is more 31 Vibrational energy levels Electronic energy levels Figure 1.1. A schematic illustration of the energy levels of an ion in a trap, showing two independent modes: A harmonic oscillator mode (left) and the internal electronic state of the ion (right). Line thicknesses in the right plot correspond to the number of Zeeman/hyperfine states in a given energy level. Solid arrows denote dipole transitions, while dashed arrows denote quadrupoles. Splittings between energy levels are not to scale. or less aligned with the magnetic field) and a set of non-degenerate states. In addition, if the ion nucleus has an inherent magnetic moment, this will also couple with the electron spin and generate hyperfine splitting, lifting the degeneracy of states based on the (anti-)alignment of the electron and nuclear spins. Both of these processes yield a set of energetically distinct magnetic sublevels in each energy level. In all of the work presented here, we use an ion species with no nuclear spin, meaning degeneracy in our system is lifted only by Zeeman splitting. 1.3 Quantum computing A quantum computer is a device that processes information that is encoded in quantum states. Typically when people refer to quantum computing, they implicitly mean computing done using information encoded in discrete two-level systems (a quantum bit or “qubit”). Qubits can be encoded in the states of many quantum systems: the current through a superconducting circuit, the spin state of a nitrogen-vacancy center, or the polarization of an individual photon, among 32 other systems (see Cheng et al. (2023) for a broad overview of the state of field around the time of writing). While as of now, no quantum computing platform has seen the implementation of large-scale “fault-tolerant” quantum computing (described in greater detail later in this section), trapped ions are currently among the leading platforms, with some of the highest reported single-(Löschnauer et al. (2024)) and two-qubit (Clark et al. (2021); Löschnauer et al. (2024)) gate performances reported in the literature, as well as recent reports of better-than- break-even error correction in small-scale experiments (Hong, Durso-Sabina, Hayes, and Lucas (2024)). As such, there is currently much interest in scaling up trapped- ion quantum computing. These next three sections will offer a broad overview of quantum computing. (More detailed discussions can be found in review papers or in books, e.g. Nielsen and Chuang (2010), with this source informing much of the discussion below), Quantum computing’s basic building blocks and the associated technical requirements that must be met for it to be practically useful, its typical implementation in trapped ions, and a recently proposed alternative trapped-ion architecture (the omg qubit scheme Allcock et al. (2021)) that motivates much of the work we do in our lab will be discussed. 1.3.1 Building blocks for QC. A quantum algorithm (a set of calculations run on a quantum computer) can be carried out by initializing a quantum register (a set of qubits) to some known state and carrying out a set of state-changing operations and measurements on it. The output of this sequence when the final state of the register is measured will be some tensor product of measurement basis eigenstates (usually written as |0⟩ and |1⟩ or |↓⟩ and |↑⟩). To determine the output state of the algorithm in this basis, the algorithm must (in 33 general) be repeated many times to build up statistics, though computations can also have deterministic outputs. The state manipulations performed on qubits during a quantum algorithm are often called “gates,” in analogy to classical computing. It can be shown that a set of single-qubit gates together with a single two-qubit gate are “complete” (i.e. that any quantum algorithm can be carried out using a combination of these gates). The meanings of these gate types are clarified here. Single-qubit gates. In classical computing, where an individual bit can only be in a 0 or 1 state, the only meaningful “single-bit gate” is a NOT operation, flipping the state of a bit. In quantum computing however, where information in encoded in a coherent superposition of states |0⟩ and |1⟩ (|Ψ⟩ = C0 |0⟩ + C1e iϕ |0⟩, where |C0|2 + |C1|2 = 1), each qubit has two continuous degrees of freedom to control. In fact, the state of a qubit can be represented as a vector on a unit sphere (the Bloch sphere, discussed in greater detail in Section 3.5), and the general set of single-qubit gates can be regarded as “rotations” of this vector. An example of a single-qubit gate is the Hadamard (or H) gate, illustrated in Figure 1.2a. Multi-qubit (entangling) gates. Entanglement is a key computational resource available to quantum systems. Multi-qubit gates (or entangling gates) generate non-classical correlations between two or more qubits by carrying out an operation that affects the state of one qubit differently depending on the state of the other. A prototypical example of this type of gate, the controlled-NOT (or CNOT) gate, is illustrated in Figure 1.2b. Gate errors. Like in classical computing, errors can occur during algorithms, leading to incorrect output states. In the classical case, the only single-bit error is an unexpected bit flip. In the quantum case, this idea can be extended to a general set of undesired rotations of the state vector, which can be coherent (i.e. consistent 34 }a b Figure 1.2. Examples of commonplace one- and two-qubit gates written in circuit notation (with each line representing an individual qubit over time) and the operation of the gate on a specific input state. a) The Hadamard gate, which rotates a single qubit. b) The controlled-NOT (CNOT) gate, which carries out a bit flip on the target qubit conditional on the state of the control qubit, generating entanglement. between runs of the algorithm or over time scales allowing coherent cancellation) or incoherent (i.e. random over the same time scales). Other classes of error exist as well, with one notable class (discussed in greater detail in Chapter IV) being leakage errors, i.e. the system entering a state not in the two-level qubit manifold. These different error types can be handled in different ways with different levels of ease. For a gate, the error rate can be quantified in terms of a gate fidelity, which is the degree of overlap between the expected and true outputs of the gate, defined as F = ⟨ΨT | ρ |ΨT ⟩, where |ΨT ⟩ is the target output state and ρ is the density operator of the true output. Error correction. Error correcting codes exist for the most common types of errors in quantum computing. Typically, these codes require that each qubit encoding data must be entangled with a set of auxilliary qubits whose states can be measured to check whether an error occurred on the data qubit. Together, these qubits are said to make up one “logical qubit” in an error-corrected quantum algorithm. At a given rate of gate errors, in order for a quantum computer to be fault-tolerant (i.e. able to push rates of logical errors to arbitrarily low levels), each logical qubit must consist of some minimum number of physical qubits. In practice, fault-tolerance is difficult to achieve because the number of physical qubits required 35 for each logical qubit grows quickly with error rate. This problem is discussed further below. 1.3.2 Technical requirements. For quantum computing to be practically useful, two distinct but related technical requirements must be considered: scalability and gate fidelity. Scalability. A useful quantum computer is one that can perform certain calculations too difficult for classical computers (e.g. the factoring of large prime numbers or the simulation of large molecules). By this benchmark, a quantum computer must be large enough that it cannot be simulated classically. A quantum computer of two logical qubits can be “simulated” with pencil and paper. One made up of 10s can be simulated in a reasonable amount of time on a supercomputer. Beyond this, a quantum computer can perform calculations that are classically infeasible. Quantum computers with 10s to 100s of physical qubits have already been implemented (Cheng et al. (2023)). However, as discussed above, error correction requires multiple physical qubits per logical qubit, and in fact, there may need to be orders of magnitude more of the former than the latter, as discussed below. In practice this means that the design of a practical quantum computer must be scalable, i.e. able to generate, store, control, and read out arbitrarily large numbers of qubits with high accuracy. Scalability is the main motivation for the work presented in Chapter V. Gate fidelity. Related to scalability (through the physical overhead required by error correction) is gate fidelity. Given an error rate of 10−4 per gate (and making some assumptions about the distributions of these errors among the different types), fault-tolerance can be acheived using a Steane code for error correction, encoding each logical qubit in seven physical qubits (Nielsen and Chuang (2010)). For higher error rates and different error correcting codes, this 36 ratio of physical qubits needed for one fault-tolerant logical qubit grows rapidly. Depending on the type of error, this number can potentially be much larger or smaller, but in general, 10−4 error rates are regarded as the “threshold” for technically feasible fault-tolerant quantum computing, and much of the current work being done focuses on reducing these rates, whether by improving coherence times (the length of time for which the phase between qubit states remains well defined), improving qubit lifetimes, developing coherent control techniques more robust to decoherence and/or calibration errors, etc. Gate fidelity is a motivating factor and important consideration for the work presented in Chapter IV. Though these requirements may seem to set very stringent standards on the size and performance of quantum computers, there has been much work on how NISQ (noisy intermediate-scale quantum) trapped-ion devices could be optimally designed for scales/performances expected in the near term (Murali, Debroy, Brown, and Martonosi (2020)) and how NISQ devices in general could be used for practical applications (Bharti et al. (2022)). 1.4 TIQC architecture In trapped-ion quantum computing (TIQC), quantum information is encoded in the Hilbert space available to an ion (or set of ions) in a trap. As discussed above, a single ion has a large set of long-lived electronic energy levels as well as three harmonic oscillator modes and their associated Fock spaces. In principle, any pair of electronic states can be selected as the |0⟩ and |1⟩ states of a qubit (which we will from here on out write as |↓⟩ and |↑⟩ to distinguish them from Fock states in discussions where these are relevant). Typically however, qubits are encoded in Zeeman/hyperfine sublevels of the S1/2 state (ground-state qubits, or just “qubits” for the remainder of this section), which has an effectively infinite lifetime. 37 To explain the standard TIQC architecture and some of its core techniques (techniques for generating entanglement, handling error correction, and reducing unwanted cross-talk between qubits with different functions), we can consider the process of building up a chain of ions in a shared potential well and the operations that need to be performed on them for carrying out a quantum algorithm. This building-up process is represented schematically in Figure 1.3. a b c Figure 1.3. An illustration of the conceptual reasoning for using mixed-species ion chains. a) A pair of ions sharing a potential well can be entangled by driving both ions with laser fields. b) For carrying out auxilliary functions (e.g. sympathetic cooling of the ions being used for gates), additional ions need to be added to the chain. If these ions are of the same species as those used for gates, when driven with laser fields, they will emit photons that the other ions can absorb, destroying the information they contain. c) If the auxilliary functions are performed using ions of a different species with different optical transition frequencies, then photons scattered by one species will not be absorbed by the other. Qubit coupling through shared motion. Two ions sharing a harmonic potential well will be driven together by the well and pushed apart by their electrostatic repulsion. The upshot of this is that the ions effectively make up a two-atom solid. (In general, this solid can be called an ion crystal, or an ion chain if the crystal is one-dimensional). This effective solid has a set of shared vibrational 38 modes. Because both ions couple to the same modes, these modes can serve to mediate entanglement between between the ions if there is a way of driving the modes that depends on the state of each qubit. In other words, if there is a way of entangling the qubit state with the state of the motional mode, then there is a way of entangling two qubit states sharing a motional mode. This entanglement is usually carried out with lasers but can also be performed with strong magnetic field gradients (e.g. Löschnauer et al. (2024)). Ancilla ions. As discussed in Section 1.4 above, fault-tolerant quantum computing requires extra error-correcting qubits. We can imagine adding additional ions to the two-ion chain discussed above to serve this purpose. An issue arises with the fact that, to carry out certain error correction schemes, the ancilla ions need to be read out “mid-circuit” (i.e. before the data qubits are to be measured). This readout process involves driving an optical transition and scattering photons. While this optical drive can be targeted to a single ion (i.e. ions can be individually addressed), some scattered photons will inevitably be absorbed by data qubits if the ancilla and data ions couple to the same optical transition (e.g. if they are both ground-state qubits in the same species of ion). This absorption and spontaneous decay destroys any information in the qubit and induces an error. Similar issues arise when trying to use ancilla ions for other tasks (e.g. sympathetically cooling the chain, as discussed in later chapters). Multi-species ion chains. To work around the problem of cross-coupling between ancilla and data ions, common practice is to use different ion species (e.g. different alkaline earth metals) for different purposes. These elements have ground state transitions at very different optical wavelengths (e.g. the S1/2 → P1/2 transition being 397 nm in Ca+ and 422 nm in Sr+), meaning that photons 39 scattered during ancilla readout or sympathetic cooling will not be absorbed by the data ions. 1.5 The omg architecture While dual-species ion chains resolve the issue of separating the functions of data and ancilla qubits, they have their own set of drawbacks. Dual-species chains require sources for two types of ion species, two separate laser systems for addressing a different set of optical transitions, and the ability to detect different wavelengths of emitted light. In addition to this increased technical overhead, mixed-species chains see less coupling through common motional modes (due to mass differences between the species), making both mixed-species entangling gates and sympathetic cooling less efficient, and the mixed masses makes the chain more vulnerable to distortion by stray electric fields. To work around these issues and achieve separation of ancilla/data functions with a single ion species, the omg qubit scheme was proposed (Allcock et al. (2021)), building off a body of previous work done using optical (Monz et al. (2016); Roos et al. (2004)) and metastable (Sherman et al. (2013)) qubits for multi- and single-qubit operations respectively and proposed in parallel with efforts to implement a similar scheme in ytterbium ions (Yang et al. (2022)). This scheme suggests using the fuller range of electronic states available for encoding for elimination of cross-talk between qubit types. 1.5.1 Optical, metastable, and ground qubits. The omg scheme draws its name from the three qubit types it identifies: optical (o), metastable (m), and ground (g) qubits. These qubit encodings are shown in Figure 1.4. As discussed above, g qubits, encoded in Zeeman/hyperfine splitting in the electronic ground manifold, are the typical choice for encoding information in trapped ions. In contrast, o qubits, encoded in the narrow-linewidth quadrupole transition between ground and metastable states, are sometimes used for information processing, but 40 typically see more use as tools for state detection in g qubits (since this transition can be used to selectively “shelve” one of the g qubit states where it will not couple to the transition used for readout). Finally, m qubits, encoded in the Zeeman/hyperfine splitting of the metastable manifold, were not widely explored at the time that the work described in this document was done, but can be useful in conjunction with these other qubit types as described below. Ground state (g) qubit Metastable state (m) qubit Op�cal (o) qubit Figure 1.4. The three qubit types considered by the omg architecture. 1.5.2 Using omg qubits together. To illustrate how the omg scheme might be used, an example will be given which specifically motivates the need to investigate the prospect of doing gates in m qubits. One can consider two ions (one in g, the other in m) sharing a potential well. The g qubit could be used for some sort of dissipative operation (e.g. sympathetic cooling of the chain or mid-circuit readout of some ancilla qubit previously encoded in the g ion) while information is stored (or is potentially even being actively manipulated) in the m qubit. In a setup like this, the separation of functions discussed in Section 1.4 is achieved in a single ion species, where g qubits are used for dissipative operations and m qubits are used for encoding and processing information. 41 1.6 Continuous-variable quantum computing In typical quantum computing schemes, as discussed above, information is encoded in two-level Hilbert spaces. Often, as in the electronic state of an ion, this Hilbert space is a subspace of some larger but still functionally finite space with varied splitting between levels. In continuous-variable quantum computing (CVQC), information is instead encoded in an infinite Hilbert space of states evenly spaced in energy. Much of the literature on CVQC focuses on its implementation in photons, where photon number states (i.e. the number of excitations of some bosonic mode) are the suggested encoding space (see Fukui and Takeda (2022) for a recent review of the state of photonic CVQC). The harmonic oscillator modes of a trapped ion can be used for much the same purpose, with phonon number state used as a computational basis (Chen, Gan, Zhang, Matuskevich, and Kim (2021); Ortiz-Gutiérrez et al. (2017)), and approaches employing hybrid discrete- variable/continuous-variable approaches using spin and motional modes together have also been proposed (Sutherland and Srinivas (2021)). 1.6.1 Building blocks for CVQC. In the way that a discrete- variable quantum computing scheme requires both single-qubit gates and multi- qubit entangling gates to be computationally complete, CVQC requires a set of Gaussian operations and at least one non-Gaussian operation that can be carried out on a bosonic mode (Braunstein and Van Loock (2005); Lloyd and Braunstein (1999)), as well as operations for entangling modes. The theory behind these operations are discussed in greater detail in Chapter III, but a brief overview of these techniques is given below. Gaussian and non-Gaussian operations. The terms “Gaussian” and “non- Gaussian” refer to the shape of the Wigner function associated with a continuous- variable mode’s state. The Wigner function (described more fully in Section 3.6) 42 is a quasiprobability distribution in the two quadratures of a mode. In a trapped ion, these quadratures may be the position and momentum associated with a motional mode, potentially in some rotating frame. A Gaussian state is one for which the Wigner function is a two-dimensional Gaussian function, and operations which preserve this feature are said to be Gaussian. Notably, the bosonic ground state (the zero-photon/phonon state) and the familiar coherent state (the quantum analog to classical mechanical oscillators / coherent light fields) are both Gaussian, as is the operation for promoting one to the other. Operations which are first- or second-order in the raising and lowering operators of a mode will be Gaussian (since, in the case of mechanical oscillators, they preserve the parabolic shape of the harmonic potential well, just changing its curvature or position). Higher-order operations (corresponding in light to interactions of two fields in e.g. nonlinear optical media) will distort the state’s Gaussian shape and are therefore called non- Gaussian. Non-Gaussian operations can also be implemented by coupling a mode to some non-bosonic system (e.g. through spin-motion entanglement in ions). Entangling operations. Like two-level systems, continuous-variable modes can be entangled with each other in ways useful for quantum information processing. In photonic systems, two modes can be entangled through interference (e.g. combining two modes on a beamsplitter), or entangled modes can be generated through correlated generation of photons in different spatial modes (as in e.g. spontaneous parametric downconversion). Analogues for both types of operations exist for motional modes in trapped ions, as discussed briefly below and in greater detail in Section 3.6. 1.6.2 Implementing CVQC in trapped ions. A trapped ion chain has an associated set of harmonic oscillator modes. Because ions are charged, their motion will couple to electric fields, and this presents a direct way of controlling 43 these modes. This means that the operations described above can be achieved in ions in a deterministic, all-electronic fashion. Gaussian operations are routinely performed on ion motion for mode/trap characterization and are sometimes used for sensing applications (Burd et al. (2019)), and electronic operations that generate entanglement between modes by exchanging energy between them have also been demonstrated (Gorman, Schindler, Selvarajan, Daniilidis, and Häffner (2014); Hou et al. (2024); Toyoda, Hiji, Noguchi, and Urabe (2015)). (The ways that we implement these operations in our setup are described in Section 2.7). The final piece required for all-electronic trapped-ion CVQC, the ability to perform non-Gaussian motional operations efficiently, is relatively difficult to achieve, as it requires strong electric potentials that are third-order or higher in ion position, which are generally not desirable in a trap. Overcoming this difficulty and implementing non-Gaussian operations motivates the work presented in Chapter VI. 1.7 Scalable ion trap design As discussed in Section 1.3, if a quantum computer carried out gates perfectly, even a small one (e.g. 100 qubits) could be used to solve useful problems. However, error rates are likely to remain high (e.g. > 10−4 for two-qubit gates) for the near future, meaning that to carry out useful computations, much larger qubit numbers will be needed for performing error correction. Because of this, a practical hardware architecture is one that is scalable, e.g. one that can be expanded from controlling hundreds of physical qubits to hundreds of thousands of them with minimal increase in technical overhead and minimal reduction in quality of control. One key consideration in scalable quantum computer design is all-to-all connectivity, or the ability to perform entangling gates between arbitrary pairs of qubits. In trapped ions, there are two widely-investigated systems for achieving this 44 all-to-all connectivity. In architectures relying on quantum networking (e.g. the modular universal scalable ion trap quantum-computer or MUSIQ system, Monroe et al. (2014)), ions in entirely separate traps are entangled via single photons. Most proposed architectures however rely on the shuttling of ions through a trap network using electric fields. The discussions in the rest of this document will be focused on this second type of trap architecture, though many of the same design considerations apply for both. Ideally, a scalable trap architecture is one that can be microfabricated (e.g. with CMOS or MEMS techniques) and integrated easily with components for ion control and state detection (which themselves could be microfabricated, potentially in ways such that integration is inherent to the fabrication process) and that can be driven with off-trap resources (e.g. laser power, trapping power, etc.) for which demand scales in a manageable way with the number of ions trapped. These considerations are discussed more thoroughly in Chapter V, but a brief account is given below of the trap architecture most widely in use and the trap designs most commonly used in research seeking to implement scalable TIQC. 1.7.1 The QCCD architecture and surface electrode traps.. A quantum charge-coupled device (QCCD, Kielpinski, Monroe, and Wineland (2002), first described conceptually in D. Wineland et al. (1998)) is a set of traps, individually controllable segments (generally sharing a trapping RF voltage), making up interconnected “channels” in which ions can be confined and through which they can be transported by adjusting the DC voltage on each segment. Using these DC voltages, pairs of ions can be brought together in short crystals, made to interact, and pulled apart for further operations. A schematic view of this type of trap is shown in Figure 1.5. The scalability of this architecture comes from the fact that a block like the one shown in Figure 1.5 represents a sort of “unit 45 3D implementation QCCD “unit cell” concept Figure 1.5. A cartoon illustration of a possible QCCD “unit cell,” which can be tiled to produce arbitrarily large trap networks. Dark blue represents DC electrodes, while pink represents trapping RF. Ions are shown in light blue. The illustration on the left shows the unit cell of a 2D/surface electrode implementation of a QCCD architecture, while the one on the right shows how the concept can be extended to three-dimensional traps. cell” which can be repeated endlessly. Each of these unit cells introduces some fixed overhead in terms of power requirements and number of control lines, but this overhead can be reduced by engineering e.g. new approaches for ion shutting and swapping (Malinowski, Allcock, and Ballance (2023)). Implementing the QCCD architecture, or something like it, motivates much of the work carried out in Chapter V, where we also discuss the typical hardware approach for seeking to implement this architecture, the surface electrode trap (or SET Chiaverini et al. (2005)), which adapts the three-dimensional Paul trap to a set of electrodes in a two-dimensional plane. 46 CHAPTER II EXPERIMENTAL APPARATUS 2.1 Ion trap To trap ions, we need to subject them to electric fields, and to generate these electric fields, we use an arrangement of electrodes to which appropriate voltages can be applied. This arrangement, shown in photographs before and after mounting on a vacuum flange (Figure 2.1), is our trap. The details of this trap’s design, geometric features, electrical drives, and vacuum environment are presented in greater detail below. 2.1.1 Trap design. We run experiments using a macroscopic Paul trap. Our trap electrodes are 0.5mm dia. rods, with the end cap electrodes (American Probe 72BE-E3/1000x1.25”) having conical tips with ≈ 11◦ internal angle and 200µm dia. spherical ends. These ten electrodes are arranged in an X-configuration, with four RF trapping electrodes surrounding a pair end cap electrodes, with a set of DC compensation electrodes ringing the outside, with uniform spaces to maximize numerical aperture (NA) for both imaging and laser access. This trap has an 0.75mm ion-electrode separation, and separation between end cap needle tips is ≈3mm. Simulation. The design process for this trap was informed by simulations performed using the AC/DC module of COMSOL Multiphysics to generate electrostatic potentials associated with voltages applied to individual sets of rods. These electrostatic potentials/fields were further analyzed in the Electrode simulation suite from NIST (NIST (2017)), allowing calculation of expected trap parameters. From simulations, the trap design used should have a geometric efficiency of 0.90 (defined as the ratio of the mode frequency of a trapped ion at a 47 10mm 7μm Figure 2.1. Photographs of our trap, showing (left) the trap attached to a vacuum flange and connected to electrical feedthroughs and (middle) a close-up shot of the unmounted trap, with the rough location of the trapping zone marked with a blue box. The associated inset in a picture taken with a CMOS camera of a five-ion chain, with ion spacing marked. Finally, a COMSOL render (right) shows trapping RF electrodes marked in red. given set of parameters to the mode frequency at the same parameters given ideal, i.e. hyperbolic, electrodes). Fabrication. The electrode rods and needles are made of beryllium-copper (98% Cu 2% Be with a hard temper) electroplated (to prevent oxidation of the electrode surface) with a ∼100 nm thick layer of gold plated on top of a palladium diffusion barrier (which is non-magnetic). To ensure that the electroplated gold was sufficiently thick, we used the scheme shown in Figure 2.2, measuring the number of gold atoms deposited onto the rod by measuring the total charge passed through the electroplating circuit. To electrically insulate the trap, the rods are held in a pair of end plates machined from Shapal Hi-M Soft, a machinable ceramic. 2.1.2 Trapping RF drive. To generate an effective potential (see Section 3.1) to confine the ions along the trap axis we need an RF voltage that can be applied to the two RF trap electrodes. This voltage is supplied by four- stage RF chain shown schematically in Figure 2.3 which can be summarized as follows: 1) an AWG supplies a sinusoidal voltage with high amplitude noise to 2) a Squareatron, which generates a stable-amplitude signal by turning the AWG input 48 DC power supply Multimeter + _ I Ethernet PC Figure 2.2. Schematic illustration of the hardware setup used for estimating the thickness of a layer of electroplated metal by measuring the total charge passed through the system during electroplating. into an effectively digital signal and stripping off all higher harmonics, with this signal in turn being fed into 3) an amplifier (Minicircuits ZX60-100VH+) which drives 4) our resonator cavity. These components are discussed in greater detail below, but for a full accounting, see J. M. Metzner (2024). AWG. A Rigol DG1022Z arbitrary waveform generator (AWG) generates an RF sinusoid voltage. In the context of the rest of the chain, this signal acts as a “frequency reference,” and can be controlled over ethernet to change the trap RF frequency when needed (e.g. for automatically reducing radial confinement for ion chain recrystallization). The output of this AWG has a large amount of amplitude noise, so we do not use it to drive our resonator directly, since fluctuations in the trap RF voltage translate to fluctuations in the mode frequencies of the ions. Instead, we take a 0 dBm AWG output and feed it into a second stage of RF conditioning. Squareatron. The Squareatron (the design for which originates at the ion trapping group at Oxford with modifications developed in our own lab) is a device that takes in a frequency reference with large amounts of amplitude noise and outputs a sinusoid with a stable amplitude. The basic operating principle is that 49 the input signal saturates an amplifier, generating an effectively digital signal, from which the higher harmonics are stripped with filters. More complete accounts of how this device works are offered in the theses of Tom Harty (Harty (2013)) and Jeremy Metzner (J. M. Metzner (2024)). The Squareatron allows ≈1 dB of digital tuning of its output amplitude, which is fed into an amplifier before being sent to our resonator. The Squareatron and amplifier are both attached to a temperature- stabilized aluminum plate in a box intended to reduce airflow. Helical resonator. To get a factor of ≈80 step up in the RF voltage going to the trap, we use a helical resonator with a resonant frequency of ≈14.5MHz with Q-factor of up to 150. Coils are mechanically stabilized with low-loss polyethylene foam supports to prevent pickup of acoustic-frequency vibrational noise. The step- up coil includes a connection for applying other signals to the trap electrodes, discussed in greater detail in Section 2.7. The output of the resonator is connected to the trap electrodes through a single wire in a vacuum feedthrough to which the RF trap electrodes are connected. Under typical operating conditions, we see radial center-of-mass (COM) mode frequencies of ∼1.8MHz, corresponding to RF voltages at the trap of ≈0.8 kV. 2.1.3 Trapping DC potential. To generate confinement along the trap axis, an equal DC voltage is supplied to the two endcap needles using a DC source (Stanford Research Systems DC205), which can output up to ±100V. The output of this voltage source is connected to the trap through a pair of wires on a vacuum feedthrough, with these connections being grounded at RF through 0.68 nF capacitors, as shown in Figure 2.4. Under typical operating conditions (e.g. ≈60V DC), we see ≈0.55MHz axial mode frequencies. 50 Trapping RF Quadrupole mo�onal drives AWG Squareatron Amplifier Resonator DDS AWGFilter box Trap Figure 2.3. Block diagram of the RF chain used to supply the RF trapping voltage. The trapping RF itself is supplied by the chain in red. Quadrupole motional drives (discussed further in Section 2.7.3 are combined with the trapping RF at the resonator. 2.1.4 Shim potentials. Unwanted electric fields in the trap (typically called “stray fields” and coming from e.g. charging of nearby dielectric materials by stray UV laser light) will tend to move the ion off of the trap center (the “RF null” discussed further in Section 3.2.5). To counteract these fields, we use our compensation electrodes to provide “shim potentials,” small DC potentials that shift the trap equilibrium point. We drive these electrodes using digital- to-analog converters (DACs) incorporated into our ARTIQ control system, described in Section 2.8 below. These DACs can supply up to ±10V DC, and each compensation electrode can generate an electric field of magnitude 19.9 (V/m)/V at the ion. Connections are through four wires on a vacuum feedthrough, each connection being grounded at RF through a 2.2 nF capacitor, as shown in Figure 2.4. 2.1.5 UHV system. The trap is kept in an ultrahigh vacuum (UHV) chamber held at 10−11 Torr. Low pressure is required in order to minimize rates of collision between trapped ions and background gasses (mainly hydrogen) that can cause decrystallization, dephasing, etc. The system is designed to be held under UHV continuously and never brought up to atmosphere, and the hardware used to 51 DAC DAC DAC D AC Bias Tee DC RF Tickle 71 dB att RF Rabi 3 dB att RF RF 2.2 nF 0.68 nF 2.2 nF 2.2 nF 2.2 nF Needle DC 0.68 nF Trap Figure 2.4. Circuit diagram for the electrical connections to the DC trap electrodes (including AC drives applied to these electrodes), with endcap needles in blue, DC trapping electrodes in white, and compensation electrodes in light yellow. maintain vacuum is not connected to air. A render of this system, with components labelled, is presented in Figure 2.5. The chamber. The ion trap sits in an octagonal chamber with six viewports on 2-3/4” flanges for laser entry, one viewport (used for imaging/state detection) on a 6” flange with electrical feedthroughs for the trap, and one custom made high-NA viewport for future ion-photon entanglement experiments. The top of the chamber is connected to a custom-fabricated spherical manifold attaching to hardware (described below) for maintaining/measuring UHV as well as a valve for pumping the system down with mechanical pumps. The bottom of chamber is covered with a 2-3/4” blank flange. The chamber is mounted in stainless steel bracketing designed in-house. Chamber components are held together with silver- 52 a b Figure 2.5. a) CAD rendering of the vacuum chamber, associated support hardware, and components directly attached to this hardware, including magnetic field coils (Section 2.2.1 and b) mounts for beam delivery optics (Section 2.6). coated screws to prevent parts from seizing during bakeout. Conflat (CF) gaskets used at the interfaces of flanges are silver-coated for the same reason. Maintaining and measuring vacuum. Vacuum is maintained using a passive non-evaporable getter pump (a NEG, SAES CapaciTorr Z 100) and an active ion pump (Agilent StarCell, 20). These parts hold the system at UHV without being connected to any mechanical pumps. Low pressures can be measured using an ion gauge (Agilent UHV-24P), which measures the current generated by ionized gasses. 53 Preparing chamber hardware. Before assembly and pumpdown of the vacuum system, most of the stainless steel components were baked at 450◦C under atmosphere for several days. This baking process causes the formation of a bronze- colored oxide layer that reduces the amount of hydrogen outgassing in the interior of the system (Sefa, Fedchak, and Scherschligt (2017)). Hardware was washed in isopropanol and acetone before and after pre-assembly baking. 2.2 Magnetic field control Qubits are encoded in the magnetic moments of our trapped ions. In 40Ca+, which has no nuclear spin, the quantization axis and energies associated with these moments are set by an external magnetic field, which induces Zeeman splitting and breaks the degeneracy of spin states in each energy level. In our system, this magnetic field is at a 45◦ angle to the trap axis and is a combination of the magnetic field generated by a set of wire coils and the local magnetic field of the earth. All in all, the magnitude of our magnetic field is ≈1.56G, producing a Zeeman splitting (between spin states, e.g. m=+3/2 and m=+5/2) of 2.63MHz in the D5/2 metastable level of 40Ca+. Noise in this magnetic field is the limiting factor in the coherence of spins encoded in these Zeeman sublevels (since their splitting is first-order sensitive to magnetic field strength). Techniques for characterizing the strength and (to some extent) spectrum of this noise will be discussed in Section 3.9, but an important noise type to discuss here is line noise (noise coming from the 60Hz mains), which we cancel using a feed-forward system. This system is described in greater detail below in Section 2.2.2. 2.2.1 Magnetic field coils. To control the magnitude and direction of the magnetic field at the ion, we use a set of five wire coils, wound with 20 gauge copper wire. Opposite pairs (spaced about 8.6” apart) of 3.6”x3.0” rectangular coils with ∼300 windings per coil provide magnetic field control in the xz-plane (in the 54 trap coordinate system defined above), while a single circular 6.1” dia. coil (about 1.4” from the ion) is used for magnetic field cancellation in the y-direction. These coils are driven with a set of three DC currents (with opposing coil pairs wired in series) supplied by a regulated DC power supply (QL355TP), which can provide up to 500mA in its low-noise mode. This maximum current is used to drive the coils along the quantization axis. An image of the coil setup is shown in Figure 2.6. a b Figure 2.6. a) A schematic of the magnetic field coil arrangement used to set the quantization axis in the trap. Current is marked with white arrows, with arrow thickness corresponding the current magnitude. b) A photograph showing the five magnetic field coils used on the trap. Worth noting are the two smaller coils (white and blue) wrapped around the top left coil, which are used for line noise feedforward (Section 2.2.2). 2.2.2 Line noise mitigation. Line noise refers to the magnetic fields produced by the 60Hz (in America) alternating current used to supply electric power to the lab. The noise itself consists of a stable, narrow-band 60Hz component and its harmonics (with the leading harmonic ordinarily being the first odd harmonic, 180Hz, produced by e.g. clipping of a 60Hz sinusoid). Essentially, line noise causes the frequency of our Zeeman qubits to oscillate on a 60Hz cycle. If we are running an experiment that is not itself in-phase with 55 this line noise (i.e. shots of the experiment are being performed at randomly sampled points over a cycle), then the line noise will look like a broadening of our qubit frequency / a reduction in spin coherence time. There are two ways of addressing this issue, discussed below: 1) running experiments in sync with the 60Hz line, and 2) cancelling out 60Hz noise using feed-forward. These techniques are represented schematically in Figure 2.7a and b respectively. (As an important note, both techniques involve the use of a Line Trigger (Sinara (2020)), a component developed as part of the Sinara open source hardware project. This device generates a 5V digital output in-phase with an input from a 60Hz power supply connected to mains). LT Experiment control sys 60 Hz AWG 180 Hz AWG TTL trigger Ion Control pulses out 60 Hz in Digital out LT 60 Hz in Digital out Experiment control sys Ion Control pulses out B-�eld at ion TTL trigger TTL trigger 180 Hz coil 60 Hz coil B-�eld at ion V180, φ180 V60, φ60 Triggering experiment B-field feed-forward Figure 2.7. Schematic illustrations of our two line noise mitigation schemes, explained in greater detail in text. Line triggering. A simple solution to the problem of line noise is to run experiments in sync with the 60Hz line. This can be done by taking a Line Trigger output, reading it in with a TTL channel on the ARTIQ crate (described in more 56 detail below), and using that input to time experiments to run at consistent points on the line cycle, ideally selecting a time where the line noise is at a crest/trough, so that the qubit frequency will be first-order insensitive to experiment run time, for experiments ≪16.7ms. This technique introduces little technical overhead, but it can make experiments substantially slower, since it forces the time between experimental shots to be 16.7ms (or multiples of this). In addition, for experiments lasting an appreciable portion of the line cycle, the qubit frequency will still be shifted over the duration of the experiment. Line noise feed-forward. A technique that requires more technical overhead, but which we generally use in our experiments, is to cancel out line noise with a feed-forward loop. In this setup, we drive small coils (∼5 windings each) with 60Hz and 180Hz voltages generated by a pair of AWG channels triggered using the Line Trigger. (As shown in Figure 2.6b, we only wind feed-forward coils around one of the quantization axis coils. In our case, our line noise magnetic field is small relative to our quantization field, so our qubit frequency will only be first-order sensitive to line noise along the quantization axis). We calibrate the amplitudes and phases of the 60 and 180Hz voltages by selecting values which maximize the contrast of our spin Ramsey fringes at long delay times (the details of this technique are explained further in Section 3.9.1). While this technique requires substantially more calibration than line triggering the experiment, it allows experiments to be run much more quickly and ensures there will no systematic effects from coherent changes in qubit frequency over the course of an experiment. 2.3 Imaging and state detection If we have the ability to generate and trap ions, before we can use them for experiments, we need to be able to 1) confirm when ions are in the trap and how many of them we have and 2) detect the states of these ions during/after 57 experiments. For both of these purposes, we use fluorescence on the 397 nm S1/2 ↔ P1/2 “cycling transition” in 40Ca+. Illuminating an ion with 397 nm laser light, it can scatter photons at a rate proportional to 1/τP1/2, where τP1/2 is the lifetime of the P1/2 excited state. This means that an ion can scatter ∼ 108 photons per second. These photons can be detected and used to infer that ions are in the trap and that they are in the electronic ground state (since if they were e.g. shelved in a metastable state, the cycling transition could not be driven). This being the case, we built a high-NA (numerical aperture) microscope for detecting 397 nm photons, both allowing ions in a chain to be imaged on a camera in a spatially resolved way and for photons to be counted using a photomultiplier tube (PMT). The hardware details and performance specifications of this imaging system are discussed below. 2.3.1 Imaging system overview. This discussion of the imaging system will start with a hardware overview and an accounting of the design process. A schematic of the imaging system, along with a photograph of the system itself in our experimental setup, is presented in Figure 2.8. The design of the objective was based on one used at NIST, with parameters optimized for 397 nm light using Zemax OpticStudio. OpticStudio was used for optimizing the placement and selection of optical elements and for predicting performance factors like magnification and spot size. Objective. The objective consists of, in order of nearest- to farthest-from-ion components, a 2” dia. aspheric convex lens (Asphericon AFL50-60-S-U), a 2” dia. aspheric convex lens (Asphericon AFL50-60-S-U), and a 1” dia. planoconvex lens (Edmund 48-321). At the end of the objective, an iris (not shown in schematic) allows the NA to be restricted. By itself, this objective should provide -3.8x magnification and a 2-3 micron spot size. 58 Figure 2.8. Schematic illustration (and photograph below) of our imaging system, with light emitted by the ion (a) being captured by an objective (b), passed through a slit (c) to reduce stray light and a narrow-band filter (d) to prevent the system from registering light far away from the wavelength of the cycling transition of Ca+, collimated with a lens (f) in an adjustable barrel mount to be redirected by a flip-mounted mirror (g), which can direct light to a focusing lens leading to the PMT (h) and (i) or to a focusing lens (j) and (k) heading to the camera (l). Stray/unwanted light reduction. After the objective, a slit (Owis SP-40) is used to reduce stray light (i.e. light entering the system from sources other than the ions at the objective focus). After this, a narrow-band optical filter (Semrock FF01-390/18-25) is used to screen out light at wavelengths other than 397 nm. Collimation and reimaging. Focused light is collimated using a convex lens (Thorlabs AC127-030-A) and sent along a path with a mirror (Edmund Optics 85-311) mounted in a motorized flippable mount (Owis KSHM 40) which can 59 direct light down two paths: To a PMT (Hamamatsu H10682-210) with just a 1” dia. lens (Thorlabs LA1540-A) used to focus the collimated light onto the PMT’s active area, or to a CMOS camera (Kiralux CS505MU) with a combination con