ABSTRACT Title of dissertation: Building and Programming a Universal Ion Trap Quantum Computer Caroline Margaret Figgatt Doctor of Philosophy, 2018 Dissertation directed by: Professor Christopher Monroe Department of Physics Quantum computing represents an exciting frontier in the realm of informa- tion processing; it is a promising technology that may provide future advances in a wide range of fields, from quantum chemistry to optimization problems. This thesis discusses experimental results for several quantum algorithms performed on a programmable quantum computer consisting of a linear chain of five or seven trapped 171Yb+ atomic clock ions with long coherence times and high gate fideli- ties. We execute modular one- and two-qubit computation gates through Raman transitions driven by a beat note between counter-propagating beams from a pulsed laser. The system’s individual addressing capability provides arbitrary single-qubit rotations as well as all possible two-qubit entangling gates, which are implemented using a pulse-segmentation scheme. The quantum computer can be programmed from a high-level interface to execute arbitrary quantum circuits, and comes with a toolbox of many important composite gates and quantum subroutines. We present experimental results for a complete three-qubit Grover quantum search algorithm, a hallmark application of a quantum computer with a well-known speedup over classical searches of an unsorted database, and report better-than- classical performance. The algorithm is performed for all 8 possible single-result oracles and all 28 possible two-result oracles. All quantum solutions are shown to outperform their classical counterparts. Performing parallel operations will be a powerful capability as deeper circuits on larger, more complex quantum computers present new challenges. Here, we perform a pair of 2-qubit gates simultaneously in a single chain of trapped ions. We employ a pre-calculated pulse shaping scheme that modulates the phase and amplitude of the Raman transitions to drive programmable high-fidelity 2-qubit entangling gates in parallel by coupling to the collective modes of motion of the ion chain. Ensuring the operation yields only spin-spin interactions between the desired pairs, with neither residual spin-motion entanglement nor crosstalk spin- spin entanglement, is a nonlinear constraint problem, and pulse solutions are found using optimization techniques. As an application, we demonstrate the quantum full adder using a depth-4 circuit requiring the use of parallel 2-qubit operations. BUILDING AND PROGRAMMING A UNIVERSAL ION TRAP QUANTUM COMPUTER by Caroline Margaret Figgatt Dissertation submitted to the Faculty of the Graduate School of the University of Maryland, College Park in partial fulfillment of the requirements for the degree of Doctor of Philosophy 2018 Advisory Committee: Professor Chris Monroe, Chair/Advisor Professor Gretchen Campbell Professor Andrew Childs Professor Trey Porto Professor James Williams ©c Copyright by Caroline Figgatt 2018 Dedication to Alan Purring, Admiral Grace Murray Hoppurr, Mr. Grayson, and most of all, Fermi ii Acknowledgments I am incredibly grateful to have received so much support from so many won- derful folks over my years in grad school. I can hardly do them all justice, and hope my humble thanks is enough. First and foremost, my advisor, Chris Monroe, provided me a wonderful op- portunity to work in his lab at UMD. I came to grad school primarily interested in making quantum computers a reality, and my work with Chris has been doing precisely that. (I still sometimes come in to the lab, realize that this is i fact my life, and can hardly believe I’m lucky enough to be doing exactly what I want to be doing.) Chris has provided me with useful guidance throughout grad school and pushed me to do my best work, and I am grateful for his support. I would also like to thank my thesis committee for taking the time to evaluate my candidacy: Gretchen Campbell, Jimmy Williams, Andrew Childs, and Trey Porto. Gretchen and Jimmy in particular have provided mentorship and support during my time at Maryland. I have been fortunate to work some incredibly kind, patient, and brilliant col- leagues throughout my time at Maryland. Norbert Linke, my postdoc for the last several years, has provided me with invaluable mentorship. He has been an incred- ibly patient teacher through endless bouts of questions, encouraged me to take on new challenges, stayed cheerful and pushed me to persevere when morale was low, and always made time to help me. I couldn’t ask for better and I am deeply grate- ful. Kevin Landsman has been a fellow grad student in the lab with me for the iii past few years, and he has been an exemplary colleague. He has made outstanding contributions to the lab, and has been a conscientious team player. I am grateful for his competence and enthusiasm during the good times, and his level-headedness and forebearance during the bad. Daiwei Zhu joined our project as a grad student a few months ago, but is already making excellent progress and has been a pleasure to work with. My immediate predecessor, Shantanu Debnath, taught me every- thing I know about optics, and always had a few moments to cheerfully answer a younger grad student’s questions. Ken Wright has been a friend as well as a great colleague. By bringing the Igor control program with him from Georgia Tech, he saved me from having to work with Labview throughout my graduate career, an act of kindness I deeply appreciate. Andrew Manning and Taeyoung Choi helped introduce me to the Gates project when I joined as a first year grad student, and I appreciate the time they took to help me learn my way around. I’ve also appreciated advice and insight from a number of older students and postdocs during my time in the Monroe lab, and am grateful for my time working with all the members of the group I’ve intersected over the years, including Kristi Beck, Jiehang Zhang, Marty Lichtman, Marko Cetina, Guido Pagano, Michael Goldman, David Wong-Campos, Clay Crocker, Harvey Kaplan, Ksenia Sosnova, Antonis Kyprianidis, Allison Carter, Patrick Becker, Wen Lin Tan, Kate Collins, Laird Egan, Sophia Scarano, Eric Bir- ckelbaw, Micah Hernandez, Nate Dudley, Steven Moses, Kai Hudek, Paul Hess, Hanna Ruth, Jason Amini, Jonathan Mizrahi, Volkan Inlek, Kale Johnson, Jake Smith, Aaron Lee, Geoffrey Ji, David Hucul, Brian Neyenhuis, Phil Richerme, Gra- hame Vittorini, Crystal Senko, Chenglin Cao, Susan Clark, and Dave Hayes. I am iv also grateful for the mentorship and advice I’ve received from Shelby Kimmel, Kristi Beck, Emily Edwards, Qudsia Quraishi, Steve Olmschenk, Kathy-Anne Soderberg, Dave Moehring, John Berberian, and so many other folks. Working on a world-class experiment with flexible capabilities has afforded me incredible opportunities to collaborate with a large number of wonderful and brililant theorists. Aaron Ostrander has been a fantastic collaborator on the par- allel gates experiment as gate optimizer extraordinaire, and it has been a pleasure to work with him. Dmitri Maslov has been an important collaborator on several experiments, most particularly the Grover experiment, and I have greatly appreci- ated his inexhaustible knowledge of quantum computing applications and his help in getting our experiment more exposure in computer science circles. Shengtao Wang was a great collaborator on the phonon hopping project. He was also kind enough to share with me a paper draft containing a detailed derivation of the XX entangling gate fidelity that proved incredibly helpful while I was deriving parallel XX gate outcomes, and for that I will forever be grateful. I enjoyed working with James Leung for the frequency modulated gates project, and appreciated his taking time to answer some questions about theory problems while I was working on the parallel gates project. I thank Zhexuan Gong for taking time on several occasions to answer questions about XX gates and pulse shaping, and enjoyed working him him for the 2014 pulse shaping paper. The quantum scrambling project involved some fascinating physics, and I thank Norman Yao for his energy and incisiveness, Beni Yoshida for his thoughtful insights and endless patience with explanations, and Tommy Schuster for his hard work on simulations; working with them has been a v wonderful learning experience. Curtis Volin provided very helpful insights when I was designing the individual control optical system. I have appreciated kind advice and support from Jungsang Kim and Ken Brown, and appreciated Ken’s keen in- sights during collaborations. I have also been grateful to work with a number of other talented theorists, including Sonika Johri, Alireza Seif, Neal Solmeyer, Rad- hakrishnan Balu, Luming Duan, Martin Roetteler, Mauricio Gutierrez, Yunseong Nam, Marcello Benedetti, and Alejandro Perdomo-Ortiz. My undergraduate career at MIT provided an important foundation for my success in grad school, and to that end I’d like to that those who made that possible. Martin Zwierlein, my undergrad research advisor, has a boundless and infectious enthusiasm for all things physics that makes him a wonderful teacher and a patient mentor. I’m incredibly grateful to him for his mentorship in the lab and his support of me through some difficult times, as well as to his group members for teaching me how to be a useful person in a lab. Nergis Mavalvala, Deepto Chakrabarty, and Sean Robinson also provided incredible support and understanding through various hardships, and I am grateful to them for the role all of them played in my eventual success in grad school. I am also incredibly grateful for the support of my friends and family through- out this process. My parents, Mr. Grayson, my brother Thomas and sister-in-law Liz, and the rest of my family have been very supportive. Gina Quan, Megan Marshall, and Lora Price have been incredibly supportive friends and gym buddies during my years here, and I couldn’t ask for a kinder, more sarcastic group with which to share silly lifting videos and commiserate about the day to day frustra- vi tions of being a physics grad student. Zach Skigen, Alan Purring, and Admiral Grace Murray Hoppurr have provided wonderful and fluffy companionship through long days of writing. Alan in particular has been most helpful with data analysis, and Grace has a knack for oversight. My roommates over the past few years have been incredibly supportive and understanding: Richard Friend, Escher, Kearci Car- son, Ella, Tiffany Calvert, and in loving memory of Nola. Dawit Girma has been incredibly supportive as my powerlifting coach, and has been endlessly patient and understanding with providing coaching around an often inconsisten schedule. To all my other friends, of whom there are far to many to list here: thank you for your hugs and understanding while I’ve gone through this process. And of course, my own little Fermi has been a constant companion through this entire process, and has been there for me during good times and bad. I have been privileged to serve as first treasurer and then president of Women in Physics at UMD during my time here, and I’m incredibly grateful for the support my fellow women in physics have given me and each other over these years. I particularly appreciate my fellow officers’ support in stepping in while I focused on my thesis this semester. I’m also grateful to the department for being supportive of WIP. I’m proud of the work the group has done during my leadership tenure and I look forward to seeing where the upcoming leadership takes it. vii Table of Contents Dedication ii Acknowledgments iii List of Tables xi List of Figures xii List of Abbreviations xiv 1 Introduction 1 2 Quantum Computer Hardware 7 2.1 The 171Yb+ Qubit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7 2.1.1 Coherence Times . . . . . . . . . . . . . . . . . . . . . . . . . 8 2.1.2 Cooling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9 2.1.3 Qubit Initialization . . . . . . . . . . . . . . . . . . . . . . . . 11 2.1.4 Qubit Readout . . . . . . . . . . . . . . . . . . . . . . . . . . 12 2.2 The Ion Trap . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16 2.3 Coherent Individual Addressing System . . . . . . . . . . . . . . . . . 21 2.3.1 Raman Beam Geometry . . . . . . . . . . . . . . . . . . . . . 22 2.3.2 32-Channel AOM Characteristics . . . . . . . . . . . . . . . . 25 2.3.3 32-Channel AOM Control . . . . . . . . . . . . . . . . . . . . 26 2.3.4 Optics for Individual Addressing . . . . . . . . . . . . . . . . 30 2.4 Individual Addressing System Upgrades . . . . . . . . . . . . . . . . 38 2.5 Expansion to 7-Ion System . . . . . . . . . . . . . . . . . . . . . . . . 40 3 Quantum Control 43 3.1 Single-Qubit Rotations R . . . . . . . . . . . . . . . . . . . . . . . . . 43 3.1.1 Rabi Flopping on 5 or 7 ions . . . . . . . . . . . . . . . . . . . 45 3.1.2 Z Rotations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47 3.2 Two-Qubit Entangling XX Gates . . . . . . . . . . . . . . . . . . . . 48 viii 3.2.1 Pulse-Shaping Scheme . . . . . . . . . . . . . . . . . . . . . . 49 3.3 Universality of the Gate Set {R,XX} . . . . . . . . . . . . . . . . . 57 4 Control System 59 4.1 System Control Overview . . . . . . . . . . . . . . . . . . . . . . . . 61 4.2 The Igor Control Program . . . . . . . . . . . . . . . . . . . . . . . . 65 4.3 Compiler . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71 4.4 Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75 5 Gate Library 79 5.1 Single-Qubit Gates . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79 5.2 Multi-Qubit Circuit Construction . . . . . . . . . . . . . . . . . . . . 80 5.2.1 Optimization Strategy Adjustments with Experimental Up- grades . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82 5.3 Two-Qubit Composite Gates . . . . . . . . . . . . . . . . . . . . . . . 83 5.4 Toffoli Gates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84 5.4.1 Toffoli-3 Characterization . . . . . . . . . . . . . . . . . . . . 88 5.4.2 Toffoli-4 Gate . . . . . . . . . . . . . . . . . . . . . . . . . . . 92 5.5 The SWAP Gate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93 5.6 Controlled-SWAP Gate . . . . . . . . . . . . . . . . . . . . . . . . . 93 5.7 Using Z-Rotations to Obviate χ Sign in XX Gates . . . . . . . . . . 96 5.8 The Square Root of CNOT . . . . . . . . . . . . . . . . . . . . . . . . 97 5.9 Quantum Scrambling Library . . . . . . . . . . . . . . . . . . . . . . 99 5.9.1 Deterministic Teleportation Protocol . . . . . . . . . . . . . . 101 5.9.2 Bell States and Bell Measurements . . . . . . . . . . . . . . . 102 5.9.3 Scrambling Unitary US . . . . . . . . . . . . . . . . . . . . . . 104 5.9.4 Scrambling Unitary UCZ . . . . . . . . . . . . . . . . . . . . . 106 5.9.5 Classical Scrambler . . . . . . . . . . . . . . . . . . . . . . . . 108 5.9.6 Circuit Optimization . . . . . . . . . . . . . . . . . . . . . . . 110 6 Complete 3-Qubit Grover Search 112 6.1 The Grover Search Algorithm . . . . . . . . . . . . . . . . . . . . . . 112 6.2 Oracles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116 6.3 Circuit Implementation . . . . . . . . . . . . . . . . . . . . . . . . . . 118 6.4 Data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124 6.5 Additional Iterations . . . . . . . . . . . . . . . . . . . . . . . . . . . 127 6.6 Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129 7 Parallel 2-Qubit Operations 130 7.1 Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131 7.1.1 Fidelity of Parallel XX Operations . . . . . . . . . . . . . . . 137 7.2 Implementation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148 7.3 Experimental Results . . . . . . . . . . . . . . . . . . . . . . . . . . . 150 7.3.1 Calculating Experimental Fidelities of 2-Qubit Entangling Gates162 ix 7.3.2 Fidelity of Parallel 2-Qubit Entangling Gates with Different Degrees of Entanglement . . . . . . . . . . . . . . . . . . . . . 164 7.3.3 Independence of Parallel Gate Calibration . . . . . . . . . . . 166 7.4 Simultaneous CNOT Gates . . . . . . . . . . . . . . . . . . . . . . . 168 7.5 The Quantum Full Adder . . . . . . . . . . . . . . . . . . . . . . . . 169 7.6 Toward a Single-Operation GHZ State . . . . . . . . . . . . . . . . . 173 7.7 Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178 8 Outlook 181 8.1 Experimental Error Sources . . . . . . . . . . . . . . . . . . . . . . . 181 8.2 Scalability of Ion Traps . . . . . . . . . . . . . . . . . . . . . . . . . . 182 Bibliography 185 x List of Tables 2.1 Final focal lengths for lenses chosen for the focusing lens system. . . . 35 6.1 Single-Solution Oracles . . . . . . . . . . . . . . . . . . . . . . . . . . 120 6.2 Two-Solution Oracles . . . . . . . . . . . . . . . . . . . . . . . . . . . 124 7.1 Comparison of optical power for parallel and single XX gates. . . . . 160 xi List of Figures 2.1 Energy levels of interest for 171Yb+. . . . . . . . . . . . . . . . . . . . 8 2.2 Qubit initialization. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11 2.3 Qubit readout. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13 2.4 Blade trap pictures. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 2.5 Vacuum chamber and optical access. . . . . . . . . . . . . . . . . . . 19 2.6 Raman transitions on the experiment. . . . . . . . . . . . . . . . . . . 23 2.7 Counterpropagating beam geometry. . . . . . . . . . . . . . . . . . . 24 2.8 Multi-channel AOM. . . . . . . . . . . . . . . . . . . . . . . . . . . . 27 2.9 Efficiency of the 32-channel AOM. . . . . . . . . . . . . . . . . . . . . 28 2.10 RF control of 32-channel AOM. . . . . . . . . . . . . . . . . . . . . . 29 2.11 DOE output. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31 2.12 Individual addressing optics solution. . . . . . . . . . . . . . . . . . . 33 2.13 Individual addressing optical setup on table. . . . . . . . . . . . . . . 37 2.14 Focused beams for individual addressing. . . . . . . . . . . . . . . . . 37 3.1 Rotations on the Bloch sphere. . . . . . . . . . . . . . . . . . . . . . 44 3.2 Rabi flops on 5 ions. . . . . . . . . . . . . . . . . . . . . . . . . . . . 45 3.3 Rabi flops on 7 ions in a 9 ion chain. . . . . . . . . . . . . . . . . . . 46 3.4 Pulse-shaping scheme for XX gate. . . . . . . . . . . . . . . . . . . . 56 4.1 The computational stack. . . . . . . . . . . . . . . . . . . . . . . . . . 61 4.2 Experimental control. . . . . . . . . . . . . . . . . . . . . . . . . . . . 62 4.3 Igor control system. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66 4.4 Example experimental sequence. . . . . . . . . . . . . . . . . . . . . . 69 5.1 Single-qubit composite gates. . . . . . . . . . . . . . . . . . . . . . . 80 5.2 Two-qubit composite gates. . . . . . . . . . . . . . . . . . . . . . . . 83 5.3 Controlled-controlled unitary. . . . . . . . . . . . . . . . . . . . . . . 85 5.4 Three-qubit composite gates. . . . . . . . . . . . . . . . . . . . . . . . 86 5.5 Toffoli-3 gate data. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87 5.6 Toffoli-3 characterization. . . . . . . . . . . . . . . . . . . . . . . . . 89 5.7 Toffoli-4 gate circuit. . . . . . . . . . . . . . . . . . . . . . . . . . . . 90 xii 5.8 Toffoli-4 gate data. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91 5.9 Controlled-SWAP gate implementation . . . . . . . . . . . . . . . . 94 5.10 CSWAP gate data. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95 5.11 XX gate χ-obviation circuit. . . . . . . . . . . . . . . . . . . . . . . . 97 5.12 Component circuits for CNOT and its square root gates . . . . . . . 98 5.13 General circuit to probe scrambling behavior. . . . . . . . . . . . . . 100 5.14 Grover protocol for quantum scrambling measurement. . . . . . . . . 102 5.15 Bell measurement circuit. . . . . . . . . . . . . . . . . . . . . . . . . 104 5.16 Circuits for the scrambling unitary US. . . . . . . . . . . . . . . . . . 106 5.17 Circuits for the scrambling unitary UCZ . . . . . . . . . . . . . . . . . 107 5.18 Circuits for the classically scrambling unitary UCS. . . . . . . . . . . 109 5.19 Unitary equivalence on EPR pair components. . . . . . . . . . . . . . 110 5.20 Optimization of XX gates in a scrambling experiment. . . . . . . . . 110 5.21 Optimization of XX gates in a scrambling experiment. . . . . . . . . 111 6.1 The Grover search algorithm. . . . . . . . . . . . . . . . . . . . . . . 113 6.2 General circuits for a Grover search algorithm. . . . . . . . . . . . . . 117 6.3 Grover search algorithm circuits. . . . . . . . . . . . . . . . . . . . . 119 6.4 Single-solution algorithm. . . . . . . . . . . . . . . . . . . . . . . . . 125 6.5 Two-solution algorithm. . . . . . . . . . . . . . . . . . . . . . . . . . 126 7.1 Pulse shape for (1,4), (2,5). . . . . . . . . . . . . . . . . . . . . . . . 151 7.2 Pulse shape for (1,2), (3,4). . . . . . . . . . . . . . . . . . . . . . . . 152 7.3 Pulse shape for (1,5), (2,4). . . . . . . . . . . . . . . . . . . . . . . . 153 7.4 Pulse shape for (1,4), (2,3). . . . . . . . . . . . . . . . . . . . . . . . 154 7.5 Pulse shape for (1,3), (2,5). . . . . . . . . . . . . . . . . . . . . . . . 155 7.6 Pulse shape for (1,2), (4,5). . . . . . . . . . . . . . . . . . . . . . . . 156 7.7 Parity curve for parallel 2-qubit gates on (1,4), (2,5). . . . . . . . . . 157 7.8 Parity curve for parallel 2-qubit gates on (1,2), (3,4). . . . . . . . . . 157 7.9 Parity curve for parallel 2-qubit gates on (1,5), (2,4). . . . . . . . . . 158 7.10 Parity curve for parallel 2-qubit gates on (1,4), (2,3). . . . . . . . . . 158 7.11 Parity curve for parallel 2-qubit gates on (1,3), (2,5). . . . . . . . . . 159 7.12 Parity curve for parallel 2-qubit gates on (1,2), (4,5(). ). . . . . . .(. ). 159 7.13 Parity curve for parallel 2-qubit gates on (1,5) XX π , (2,4) XX π .165 4 8 7.14 Independence of parallel gate calibration. . . . . . . . . . . . . . . . . 167 7.15 Independence of parallel gate calibration, continued. . . . . . . . . . . 168 7.16 Simultaneous CNOT gates. . . . . . . . . . . . . . . . . . . . . . . . 169 7.17 Full adder circuits. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170 7.18 Full adder data. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171 xiii List of Abbreviations ASP Algorithm Success Probability AOM Acousto-Optic Modulator AWG Arbitrary Waveform Generator CNOT Controlled-NOT gate DAC Digital to Analog Conveter DDS Direct Digital Synthesizer DOE Diffractive-Optic Element ELU Elementary Logic Unit EOM Electro-Optic Modulator FPGA Field Programmable Gate Array GUI Graphical User Interface H Hadamard gate NA Numerical Aperture OTOC Out of Time Order Correlator PMT Photo-Multiplier Tube QCQP Quadratically Constrained Quadratic Program R Rotation gate RF Radio Frequency SK1 1st order Solovay-Kitaev composite pulse SPAM State Preparation And Measurement SSO Squared Statistical Overlap TTL Transistor-Transistor Logic UHV Ultra-High Vacuum VVA Voltage Variable Attenuator XX Two-qubit σxσx entangling gate Yb Ytterbium xiv Chapter 1: Introduction The potential applications for quantum computers have been of great interest since Manin and Feynman first proposed them in the early 1980s [1, 2]. By taking advantage of properties unique to quantum systems like multi-state superposition and multi-qubit entanglement, a large-scale quantum computer promises advances in humanity’s ability to tackle complex chemistry problems that could revolutionize the energy sector, solve optimization problems with uses across almost every industry, accelerate breakthroughs in biology and medicine, and - depending on your point of view - either wreak havoc on or revolutionize the security and data industries. To simulate a two-level quantum mechanical system on a classical computer (like the one you’re probably reading this thesis on), a system with N particles re- quires keeping track of 2N amplitudes, an exponential growth in problem size that prevents simulating large quantum systems; even the world’s biggest supercomput- ers have not been able to simulate all amplitudes in a quantum system larger than 49 particles [3]. Future advances in computing will likely push past that, but most predictions anticipate only getting to 60-70 particles. Simulating 300 quantum par- ticles would require a classical computer the size of the known universe, which is not practical. However, a quantum computer can simulate an N -particle quantum 1 system with much less than exponential growth, simply because it is a quantum sys- tem as well; in principle, the problem size can be O(N) particles, with perhaps some additional resources required for error correction. This power of “quantum paral- lelism” [4] can be used to directly simulate quantum systems of interest, such as for quantum chemistry applications, but can also be harnessed to solve other kinds of problems, such as optimization problems. One particularly interesting potential application in the field of quantum chemistry is to better understand biological ni- trogen fixation [5], a process that is central to the creation of fertilizer. Our best fertilizer creation methods require a substantial proportion of the human energy budget, but bacteria can create the same chemicals far more efficiently. A quantum computer could help us discover these efficient methods for producing fertilizer, mak- ing it easier to feed humanity while reducing our energy usage, our carbon footprint, and our effects on global climate change. The exact scope of quantum computing technology’s effect on humanity remains tantalizingly unclear, but even pessimistic outlooks envision significant benefits from large quantum computers. Consequently, increasing numbers of government and industry players are investing in this tech- nology - from U.S. government agencies like IARPA, the NSF, the Departments of Defense and Energy, and the NSA, to companies like Google, Intel, Microsoft, and startups like Rigetti and IonQ. Of course, all of this being of any use at all hinges on an important question: how do we build a quantum computer big enough to do something interesting? In this thesis, I leave aside some important abstract issues that arise from this central question (how do we define “big”? What is “interesting” in this context? What 2 constitutes a quantum computer, exactly?) to focus on the pragmatic and practical aspects of the question: “building” and “doing”. Ion trap quantum computers [6, 7] and superconducting quantum comput- ers [8,9] represent the two proposed quantum computing platforms that so far have made the most progress toward a practical and usable technology, and have opti- mistic outlooks for scaling up. Recent advances in both of these platforms portend optimism that a large-scale quantum computer will eventually be built, though chal- lenges certainly remain. Additionally, it is not yet clear which platform will prove itself to be the best candidate for a burgeoning technology, or whether the most successful quantum computers will be some other hardware [10] that is currently still being explored (such as neutral atoms [11, 12], quantum dots [13, 14], nitrogen vacancy centers in diamond [15], or large photonic systems [16,17]), a hybrid of two or more quantum systems [18], or even one that has yet to be proposed at all. In this thesis, I will discuss an ion trap quantum computing machine we have built [19, 20] and describe some results obtained with it. To demonstrate that our machine is indeed a quantum computer, I use the DiVincenzo criteria [21] as a practical measure for a highly programmable and usable device that is similar enough to modern digital computers to be a familiar and user-friendly concept, but still takes advantage of all available resources of quantum computing. The criteria, and our adherence to them, are summarized below. • Qubits: We use well-characterized 171Yb+ ions as qubits (see Section 2.1) in a scalable trapped-ion quantum computing architecture [7]. 3 • Qubit initialization: 171Yb+ qubits can be initialized to a known state with high fidelity (see Section 2.1.3). • Long coherence times: These qubits have typical coherence times longer than 1 s, which is many orders of magnitude greater than gate times of 20-200 µs and hence allows for long coherent gate sequences (see Section 2.1.1). • A universal gate set: The native gate set available, single-qubit R rotations and two-qubit entangling XX gates, constitute a universal gate set (see Section 3.3). • Qubit readout: State-dependent fluorescence of 171Yb+ qubits allows for high- fidelity qubit measurement (see Section 2.1.4). Two additional optional DiVincenzo criteria exist that address a system’s capabil- ities for quantum communication; these are not addressed here, as the system is not designed to communicate with other quantum systems and hence does not meet these criteria at present. Other efforts exist to implement robust quantum communi- cation and networking [22] and provide long-distance quantum communication [23] using ion traps and other hardwares. To be most useful, a quantum computer should be easily programmable and flexible to be able to tackle many kinds of problems, as a classical computer can. To this end, we have built an ion trap quantum computer with controls organized into separate levels of abstraction, called a computing stack. Each level of the stack represents a different level of control in the system, and is designed to be a black box to the others; this architecture allows users to implement quantum algorithms 4 without needing to know much, if anything, about what qubits the system uses or the physics and engineering details behind the hardware. This framework is very similar to classical computers; I don’t need to know anything about NAND gates or silicon microprocessors to read the news on the Internet or run data analysis calculations for this thesis. Most previous quantum computing experiments have been built to perform a single algorithm, with the hardware designed and tailored for that purpose, and without flexibility to perform many other experiments; the stack framework of this machine represents a step forward from a systems perspective, as well as the previously-undemonstrated results we have generated on it. The first half of this thesis, Chapters 2-5, will describe the functioning of the ion trap quantum computer by going through the stack, from the low-level hardware up to the high-level user interface. The second half, Chapters 6-7, will describe several new results generated from the system over the course of this thesis work, and can additionally be found in [24, 25]. Additional work performed on this machine over the course of this thesis work but not discussed here in detail can be found in [19,26–32]. In Chapter 2, I present the hardware level of the stack: trapped 171Yb+ ions as qubits, with various lasers of initialization, readout, and coherent control. In Chapter 3, I discuss the quantum control level of the stack, consisting of pulse shaping schemes that effect high-quality quantum gates. In Chapter 4, I explain the compiler level of the stack, which consists of a computer control program that compiles quantum gate commands from the user into the low-level operations necessary to execute the gate sequence on the experiment. 5 In Chapter 5, I enumerate the library of quantum gates we have assembled for use in the compiler to execute various algorithms. In Chapter 6, I report results for a complete 3-qubit Grover search algorithm. In Chapter 7, I introduce a promising new tool for quantum control: parallel 2-qubit operations on pairs of ions in the same ion chain. In Chapter 8, I conclude with a positive outlook on ion traps as quantum computers, and discuss future challenges and suggest areas to improve. 6 Chapter 2: Quantum Computer Hardware In this chapter, I will present an overview of the hardware of the ion trap quantum computer used to perform the work in this thesis. The overview will be brief, as much of this material has already been presented in [20], which concerns the same system. I will also include some upgrades that have since been made to the system, including to the individual addressing system allowing for better individual ion controls crucial to work presented in Chapter 7, and expanding from 5 to 7 ions for use in the work presented in [31] and future work. 2.1 The 171Yb+ Qubit In this system, we have chosen as our qubit the first-order magnetic-field- insensitive pair of clock states in the hyperfine-split 2S1/2 ground level manifold of 171Yb+, with |0〉 ≡ |F = 0;mF = 0〉 and |1〉 ≡ |F = 1;mF = 0〉 (see Figure 2.1) [33]. The two qubit states have a 12.642821 GHz frequency difference, which can be addressed with microwaves or Raman transitions via the 2P levels. 7 Figure 2.1: Energy levels of interest for the 171Yb+ qubit. The qubit is defined as |0〉 ≡ |F = 0;mF = 0〉 and |1〉 ≡ |F = 1;mF = 0〉. 2.1.1 Coherence Times The |F = 0;mF = 0〉 and |F = 1;mF = 0〉 states in 171Yb+ ions are so-called “clock” qubits because they are insensitive to magnetic fields to first order. On our experiment, our measured coherence time is 1.5(5) s, which is more than 3 orders of magnitude greater than typical single-qubit gate times of 20 µs or two-qubit gate times of 200 µs; for short algorithms with at most a few tens of two-qubit gates, this coherence time is therefore sufficient. On this experiment, performance is primarily limited not by qubit coherence, but by pointing noise on the Raman control beams driving coherent transitions; this is further discussed in Chapter 8. 8 2.1.2 Cooling The qubits are cooled to near their motional ground state in a two-step process. First, the ion is cooled using Doppler cooling [34,35]. Light beams detuned from the {2S 21/2, F = 1} to { P1/2, F = 0} and {2S1/2, F = 0} to {2P1/2, F = 1} transitions is applied to the ions, with both π and σ polarizations added to address all of the Zeeman states in the ground state manifold. The two transitions are applied by supplying 369 nm light resonant with the {2S 21/2, F = 1} to { P1/2, F = 0} transition, and adding a 14.7 GHz sideband to access the {2S1/2, F = 0} to {2P1/2, F = 1} transition using the second-order sideband of an EOM at 7 GHz. The 369 nm light is generated by doubling the output of a 739 nm laser that is locked to an iodine vapor cell [36]. The output of the doubler is 300 MHz detuned from the transition, so frequency shifts and TTL control are provided by one IntraAction AOM at 290 MHz to give the optimal Doppler cooling frequency for a Γ/2π = 19.7 MHz linewidth. A light beam red-detuned from resonance by 300 MHz (the “protection beam”) acts on the ions during trapping only, and two more beams red-detuned by 10 MHz with an AOM cool the ions further during setup before experimental procedures. Red-detuned photons are absorbed by hot ions moving toward the beam source, and then scattered in a random direction, reducing their net momentum (and con- sequently, their kinetic energy) in the direction of the cooling beams. The three beams are therefore at angles to each other to ensure cooling along all axes; the main 10 MHz beam travels at a very slight angle from the ion chain axis in the horizontal plane, the 300 MHz protection beam is applied in the horizontal plane at 9 an angle near perpendicular to the chain axis, and the low-power 10 MHz “oblique” beam is applied with a vertical component at an oblique angle to the other two (see Figure 2.5). On our experiment, we define the z axis as the horizontal direc- tion along the trap axis, the x axis as the horizontal axis perpendicular to the trap axis, and the y axis as the vertical direction, perpendicular to the trap axis and the optical table. Consequently, the beams have components along the x, y, and z directions, and while we don’t reach the Doppler limit, the ions are cooled enough in all directions (n̄ = 10 to 15) that resolved sideband cooling can finish cooling the ions to near their motional ground states. Second, resolved sideband cooling [37] is necessary to further cool the ions to the Lamb-Dicke regime, which is necessary to execute high-fidelity gates. The Lamb-Dicke parameter η can be described as a measure of the coupling strength between the ion’s internal or spin state (so called because we are only using the two qubit states) and its motional state. In the Lamb-Dicke regime, this parameter is small enough that transitions changing the motional quantum number by more than 1 phonon are strongly suppressed. Specifically, η2(2n+ 1) 1 (2.1) where n is the motional quantum number of the ion. Resolved sideband cooling is a process that couples to the motional modes of the ion to progressively remove phonons while cycling between the |0〉 and |1〉 spin states [38]. This cools the ions below the Doppler cooling limit, and on our experiment ends with a typical n̄ = 0.1 10 Figure 2.2: Qubits are initialized to the |0〉 state via optical pumping. The straight blue arrows indicate photon absorption, and the squiggly purple arrows indicate photon emission. quanta. 2.1.3 Qubit Initialization The qubits are initialized to their |0〉 state via optical pumping [33]. The ions are illuminated with light resonant with the transition between the {2S1/2, F = 1} and {2P1/2, F = 1}manifolds, as shown in Figure 2.2. An AOM provides a frequency shift of 300 MHz to make the 369 nm beam on-resonant with the {2S1/2, F = 1} and {2P1/2, F = 0} transition, and an EOM adds 2.1 GHz sidebands to then reach the {2P1/2, F = 1} manifold to clear out the {2S1/2, F = 1,mF = 0} state. The beam contains both π and σ polarization components to ensure successful pumping for ions 11 in any of the Zeeman states in the {2S1/2, F = 1} manifold. (π alone is sufficient, but the beam path is shared with the detection beam discussed in Section 2.1.4, which requires both components.) Once pumped to the {2P1/2, F = 1} manifold, the ions then decay either back to the {2S1/2, F = 1} manifold, where they will be pumped again, or to the {2S1/2, F = 0} target state. Once there, ions in the {2S1/2, F = 0} state stay there, as the applied beam is now 14.7 GHz off-resonant from a transition back to the {2P1/2, F = 1} manifold, so running the pump beam for a short amount of time (5 µs) is sufficient to initialize all qubits to the |0〉 state with high fidelity. 2.1.4 Qubit Readout At the end of a quantum procedure, the qubits are measured and their states read out using state-dependent fluorescence, as shown in Figure 2.3(a). Light res- onant with the transition from {2S1/2, F = 1} to {2P1/2, F = 0} is applied to the ions, using a 300 MHz AOM to shift a beam of 369 nm light to resonance. Qubits in the |1〉 state are then excited to the {2P1/2, F = 0} state, which has a lifetime of 8.12 ns, and then decays back to the {2S1/2, F = 1} manifold, emitting a photon in the process and hence termed “bright.” Polarization elements in π and σ are added to ensure continued cycling from all of the Zeeman levels in the {2S1/2, F = 1} manifold, allowing many photons to be scattered over the course of the detection period. If the qubit is instead in the |0〉 state, the applied light is 2.1 GHz away from the nearest dipole-allowed transition to the {2P1/2, F = 1} manifold, and so 12 (a) (b) (c) Figure 2.3: (a) Qubits are measured using state-dependent fluorescence. Ions in the |1〉 state fluoresce in response to an applied detection beam; ions in the |0〉 state do not. The straight blue arrows indicate photon absorption, and the squiggly purple arrows indicate photon emission. (b) A typical detection histogram showing photon counts for bright (orange) and dark (purple) ions. The measurement fidelity for a qubit in |0〉 is 99.74(3)% and for a qubit in |1〉 is 99.09(5)%. From [20]. (c) Individual ions are detected on separate channels of a linear array of 32 PMTs. This shows photon counts for 5 bright ions; the X-axis is the PMT channel number. 13 the qubit scatters no photons and remains dark. The detection process is run for 150µs to ensure sufficient photons are collected to discriminate between bright and dark qubits. The entire experiment is repeated several hundred or thousand times to build up statistics, and photons are collected with an imaging objective with a numerical aperture (NA) of 0.37. Figure 2.3(b) shows a typical histogram showing the number of photon counts for bright and dark states. The distributions are well- distinguished, and we can discriminate between |0〉 and |1〉 by determining that any qubit with 0 or 1 counts is dark, and any qubit with 2 or more counts is bright. Measurement errors arise when a bright qubit is measured as dark or a dark qubit is measured as bright during the measurement process. The former can occur during the measurement due to off-resonant coupling between the F = 1 states of 2S and 21/2 P1/2, as the {2P1/2, F = 1} manifold is only 2.1 GHz away from the {2P1/2, F = 0} state. Once pumped to the {2P1/2, F = 1} manifold, the ion can decay to {2S1/2, F = 0}, where it ceases scattering photons. Off-resonant coupling between |0〉 and the {2P1/2, F = 1} manifold is also possible, which would scatter a photon upon decay, but with a detuning of 14.7 GHz, this is much less likely. This results in the not-quite-Poissonian histograms in Figure 2.3(b); in particular, the off-resonant coupling of |1〉 to {2P1/2, F = 1} is the largest effect, resulting in visible histogram points at 0 and 1 counts for an ion prepared in the bright state. Background scatter from the trap blades can also result in photon counts for dark ions, but has a small effect. Consequently, the measurement fidelity for a single ion in |0〉 is slightly higher than for |1〉: 99.74(3)% and 99.09(5)%, respectively. To distinguish between the fluorescence of different ions, the ions are imaged 14 onto separate channels of a 32-channel linear PMT array1. Figure 2.3(c) shows typ- ical data for N = 5 bright ions on the PMT array, where the X-axis is the channel number. The ions are imaged onto every other channel to reduce electronic crosstalk errors - neighboring channels have about 4% crosstalk error, and next-nearest neigh- bors about 1%. The crosstalk is also asymmetric, as channels to the right see more spillover from channels to their left than the other way around. Dark counts are 1-2 per channel per second and therefore negligible. When data is collected, each chan- nel is analyzed separately, using the discriminator mentioned above to determine which ions are in |0〉 or |1〉 after each experiment. After the experiment is repeated many times, the data can then be assembled into a vector of output populations, which gives the population in each of the N2 N -qubit states; for 5 ions, the state vector gives the populations in each of {00000, 00001, 00010, 00011, . . .}. Bright-to-dark pumping and crosstalk are the two main sources of error for our multi-qubit readout process. We account for this and state preparation errors by measuring our SPAM errors and then renormalizing our data appropriately. To measure our SPAM errors, we trap a single ion, then use the DC voltage controls to move that ion to each of the locations in the trap where an ion would be for an N -ion chain. Two measurements are taken at each location: one where the ion is initialized to the dark state, and another where the ion is initialized and then coherently rotated to the bright state using a microwave pulse resonant with the 12.6 GHz qubit frequency. The measurements are repeated 80,000 times to build robust statistics. The measured data is then assembled into a N2 × N2 matrix of 1Hamamatsu H7260-200 15 each of the possible states. Once data is taken for an algorithm of interest, this SPAM matrix can then be inverted and multiplied with the corresponding N2×N2 state matrix for the data to remove SPAM errors [20, 39]. Additional methods of measuring SPAM errors on our experiment are presented in [40]. 2.2 The Ion Trap The ions are held in an ion trap in a chamber held at room temperature [41]. Our physical qubits are well-isolated from the environment in a stainless steel chamber that has been evacuated to ultra-high vacuum (UHV) with pressure about 10−11 Torr, meaning a given ion will only experience a collision with a background gas particle once every hour or so on average. This is therefore negligible on the order of the coherence time, and primarily limits the lifetime of the ion in the trap. Consequently, we can hold chains of 5 or 7 ions for about an hour at a time. The trap used here is a blade trap, a variant of a linear Paul trap [42]. Four gold-coated electrode substrates shaped like razor blades provide the confining po- tential (see Figure 2.4 for several views of the blade trap). Two blades, diagonally across from each other, provide an oscillating pseudopotential that confines the ions along the radial directions of the trap. The RF signal supplying the pseudopo- tential is generated by a DDS and amplified to ∼300V using a stabilized helical resonator [43,44] at a driving frequency of 23.8167 MHz. The other two blades are divided into 5 electrodes each that then have static voltages applied. These DC electrodes provide an axial confinement of 296 kHz 16 (a) (b) (c) Figure 2.4: Several views of the ion trap blades. (a) A view along the blade, which can be seen here to be divided into 5 electrodes for enhanced ion position control on the DC blades. (b) A view of the 4 blades, looking directly down the trap axis. (c) Another view of the 4 blades, shown at an angle from the trap axis. Image credit S. Debnath. 17 for 5 ions and also allow for finer control of the spacing between the ions. Axial confinement requires, at minimum, 3 electrodes on each blade, where the center electrode on each blade is set to a negative voltage (lower potential for the positively- charged ion), and the outer electrodes are set to a positive voltage (higher potential for the ion). This creates a quadratic potential well confining the ion. Multiple ions in such a trap will not be equally spaced; the inner ions will be pushed closer together than the outer ions, and the spacing disparities increase with increasing ion number [45]. Since our indvidual addressing beams are equally spaced, equal spacing of ions would be ideal. Very nearly equal spacing can be achieved by adding at least 2 more electrodes to each blade, which allow for the addition of a quartic term to the trap potential; appropriately selected trap voltages will create a potential with a “flat” area in the center that will hold ions at approximately equal distances from each other. In practice, however, simulations of our blade trap showed that such a potential would require kilovolts on the outer electrodes.Equal spacing for large numbers of ions will likely require a microfabricated surface electrode trap with many DC electrodes to control the positioning of many ions. However, for 5 ions in our blade trap, the spacing differential is small enough that the individual beams are still fairly well aligned on the ions; the Rabi frequencies seen by the ions are comparable (see Section 3.1.1.) Below the trap, two small stainless steel tubes, or ovens, contain small amounts of solid ytterbium metal that has been isotopically purified to be 95% 171Yb (natural abundance 14%). Niobium wires wrapped around the ovens heat up when current is applied, allowing a small amount of heated ytterbium to sublimate into a vapor 18 355nm Raman + 399nm + 369nm cool 369nm cool, pump, detect + 935nm Microwave Detection Collection 369 cool Figure 2.5: A visualization of the vacuum chamber containing the ion trap, and the optical access for the various necessary control components. The Raman control beams (red) and one of the cooling beams are applied radially, orthogonal to the ion chain, while another cooling beam, the detection and pump beams, and the loading beam (blue) are applied along the chain axis. A third cooling beam is applied through a small window at angles to both of the other cooling beams (turquoise.) Microwaves are applied through a horn from below (purple.) Finally, detection light is collected from above the trap (green.) Chamber CAD modified from S. Debnath [19]. 19 of neutral ytterbium atoms, some of which will spread into the trapping area. A laser at 399nm, tuned to the 1S0 →1 P1 atomic transition in neutral 171Yb, is directed through the trap area, so some atoms will absorb a photon and be excited to the 1P1 state. An additional photon at ≤369 nm is required to achieve the first photoionization energy and strip off one electron to create 171Yb+ [46]; the various 369nm laser beams already passing through the trap that will next be used for cooling purposes serve to complete this ionization process. Once ionized, the atom is now subject to the trap potential, and can be cooled to the motional ground state by the cooling beams. The ions then form a linear Coulomb crystal along the weakly-confined axial direction in the trap. On this experiment, once the ovens heat up (a 3 minute process), 5-9 ions can typically be trapped within a minute or two. As shown in Figure 2.1, states in the 2P1/2 manifold will decay to the metastable 2D3/2 level with a branching ratio of 200:1. To repump the ion, a 935 nm laser beam is applied constantly to the ions. This repumper induces a transition to the 3[3/2]3/2 level, which then decays to the ground level. The repumper beam has 3.07 GHz sidebands added by an EOM and has both π and σ polarization components to address all states in the 2D3/2 manifold. All of the optical controls discussed so far, as well as the Raman beams de- scribed in the next section, present a challenge in delivering all of these controls to the ions in the trap. Figure 2.5 shows a mockup of the vacuum chamber, which has 8 windows around the trap providing ample optical access for cooling beams (369nm) at 3 different angles, a pump/detection beam (369nm) transported through a shared fiber at different times during the experimental sequence, the loading beam 20 (399nm), the repumper beam (935nm), counterpropagating 355nm Raman beams applied from opposing sides (see Section 2.3), microwaves (12.6 GHz), and detection light collection. 2.3 Coherent Individual Addressing System Coherent transitions between the |0〉 and |1〉 states of the qubit can be driven using direct microwave transitions at 12.642821 GHz, or via the P -levels using Ra- man transitions driven by two beams with a beatnote of 12.642821 GHz. Microwaves have two major drawbacks. First, they are not straightforward to focus in free space to individually address an ion in a trap [47]. Second, microwaves provide very little net momentum transfer to couple to the motional modes of the chain, which is what we use to entangle multiple ions; doing so is possible, but again not straightfor- ward, as it requires a surface trap with waveguides [48]. Individual addressing and implementation of two-qubit gates on non-clock-state qubits can alternatively be accomplished with microwaves by generating a very large magnetic field gradient, which induces different Zeeman splittings in the qubits that can then be individu- ally addressed by changing the microwave frequency and used to couple to motional modes [49]. However, the magnetic field gradient required is not scalable to large numbers of ions in the trap, as the gradient required for more than a handful of ions quickly becomes infeasible. On our experiment, we do provide a microwave horn that emits a 12.642821 GHz signal and globally addresses all ions, as a useful tool for troubleshooting and performing diagnostics. Coherent operations during algorith- 21 mic processes are provided by individually-addressed Raman beams, as described in this section. 2.3.1 Raman Beam Geometry Coherent qubit operations are performed by a pair of counterpropagating Ra- man beams on the ions in our trap (see Figure 2.6(a)). The Raman transition is performed by a mode locked 355nm pulsed laser. The two Raman beams there- fore each consist of a frequency comb [50], with components separated by the laser repetition rate, about 118 MHz. To address the 12.6 GHz qubit splitting, the coun- terpropagating Raman beams must be lined up in frequency space so that there is a 12.6 GHz beat note between each corresponding tooth in the two frequency combs. To account for noise on the laser repetition rate, an AOM beat note lock (similar to that described in [51]) is constructed, in which the repetition rate noise is inverted and added to an AOM on one Raman beam. Consequently, one Raman beam can coherently stimulate a transition from the |0〉 or |1〉 state to a virtual state partway between the 2P 21/2 and P3/2 manifolds in 171Yb+, then the other beam induces a transition to the other qubit state, as shown in Figure 2.6(b). The commercially- available 355 nm laser wavelength is close to minimum spontaneous emission from the 2P 21/2 and P3/2 manifolds, which is 1/3 of the way between the two manifolds since 2P3/2 has more states in the manifold and therefore twice the coupling of the 2P1/2 manifold. Due to the counterpropagating geometry and the large amount of energy carried by 355 nm photons (844 THz), stimulated Raman transitions impart 22 (a) (b) Figure 2.6: (a) An overview of the experimental setup, showing the counterpropagat- ing Raman beams, the ion chain, and the multi-channel PMT detection apparatus. The use of spin-motion couplings to enact 2-qubit entangling gates results in all possible interaction pairs being available for entangling gates. Image from [19]. (b) Raman transitions at 355 nm in 171Yb+. 23 (a) (b) Figure 2.7: (a) Counterpropagating beams overlapping 5 example ions. The wide beam is the global beam that interacts with all ions, while the narrow beam is one individual beam aligned to interact with a single ion. (b) The two arms of the Raman beam setup. a large momentum kick to the ions (see Figure 2.7(a)); this will allow us to couple strongly to the motional modes to implement 2-qubit gates (see Section 3.2.) The beam path is split to provide the two arms of the counterpropagating geometry, as shown in Figure 2.7(b). Each arm is then modulated by an AOM, one of which is controlled by the beat note lock, and the other of which is is controlled by the output of an AWG that modulates the frequency as needed, such as to address the red and blue sidebands necessary to implement a Mølmer-Sørensen gate [52,53]; see Section 2.4 for discussion of AWG control of the AOMs, which was changed partway through this thesis work. The global beam is modulated with an IntraAction AOM at 130 MHz. After the AOM, cylindrical lenses shape the beam to be wide along the trap axis and narrow in the transverse direction, so that light from the beam falls on all of the ions in the chain, while maintaining a narrower focus of ∼20 µm perpendicular to 24 the chain. The individual beam arm of the Raman setup is split horizontally into 10 beams; each individually couples to one of 10 neighboring channels of the 32- channel AOM. The beams are deflected vertically, which is orthogonal to the ion chain axis. Since the zeroth-order beams are not needed, a D-mirror is used after the AOM to redirect them to a beam dump. The first-order beams continue to the ions, where the AOM-provided modulation allows Raman transitions to occur. Since only deflected beams reach the ions, shutting off the RF signal to a given channel will shut off the beam that reaches the corresponding ion; hence, control over the RF signal to each channel provides control over which ions perform operations. 2.3.2 32-Channel AOM Characteristics AOMs are a tool used frequently in atomic physics experiments. Laser light of frequency f and wavelength λ is passed through a transparent crystal, such as fused silica. A piezoelectric transducer, controlled by an applied RF signal of frequency F , transmits sound waves of wavelength Λ into the crystal, which varies the index of refraction in the crystal. In the Bragg regime, the incoming light is then diffracted into symmetric orders m (see Figure 2.8(a)), each of which is also modulated by its order number times the applied RF frequency, f → f +mF, (2.2) 25 and displaced by an angle mθ, where mλ sin θ = (2.3) 2Λ Most of the light can be directed into the first diffraction order by adjusting the incoming light’s angle of incidence in the plane of the acoustic waves. The 32-channel AOM varies from the typical single-channel model in that it consists of a single crystal with 32 evenly spaced transducers, as shown in Figure 2.8(b). The transducers are spaced about 450 µm apart, with a channel width of 100 µm. Each transducer is controlled by its own RF input channel; all of the channels are designed for a drive frequency at 200 MHz frequency and 700 mW in power. The channels are arranged horizontally, and activated channels deflect beams vertically. The crystal and transducers are watercooled to accommodate that much RF power while suppressing temperature-based signal fluctuations. We characterized the efficiency of the AOM, and found that all channels had > 65% efficiency into the first order, and a comfortable 20 MHz range over which the efficiency was constant (for a typical channel, see Figure 2.9.) Crosstalk was also found to be minimal, with < 1% optical and RF crosstalk. 2.3.3 32-Channel AOM Control Providing RF power to the AOM channels has some important considerations in order to work with the Raman beam setup, and so the circuit design and com- ponent selection must be performed carefully. The circuit design is shown in Figure 26 (a) (b) Figure 2.8: (a) Diagram showing how a single channel AOM works, with two positive and negative orders shown. (b) The 32-channel AOM differs from its more common single-channel counterparts in that it has 32 piezoelectric transducers on a single crystal, hence creating 32 channels that can independently modulate an input laser beam. 27 Figure 2.9: Efficiency for channel A15 of the 32-channel AOM. The 20 MHz range over which the efficiency is maximized is sufficient for experimental use, as the RF control frequency signal does not vary more than 10%. 2.10. First, to control 10 beams, we need 10 independent channels. A two-way power splitter (Minicircuit ZSCJ-2-1+) is used, and each output is then sent to a five-channel control box with a five-way splitter (Minicircuits ZFSC-5-1-SB+) for a total of 10 channels. Second, we must provide a means of control over when a beam is deflected. A TTL switch (Miniciruits ZASWA-2-50DR+) is used on the RF input for each channel, which is in turn operated by the experimental control computer. Since the switch time is at most 20ns, this allows a given ion’s Raman beam to be shut off or turned on quickly relative to qubit rotations that take tens of microseconds, or two-qubit gate times of hundreds of microseconds. We also provide a manual switch option. Third, the 32-channel AOM is rated by the manufacturer to take up to 700mW 28 Figure 2.10: Component layout for RF control of 32-channel AOM. (+28.45dBm) of RF power per channel. Hence, the signal control on each channel should be designed to produce no more than 700mW. To adhere to this limit, the beat note power output was measured, insertion loss from the TTL and splitters were calculated, and then amplifiers were chosen that would produce at least +29dBm (Minicircuits ZX60-P103LN+, gain +18dB; Minicircuits ZHL-2-12, gain +26dB). The channels were then individually calibrated by measuring the power output of each channel with a spectrum analyzer, and then adding passive attenuators before the pre-amplifier until the measured power output was less than +28.45dBm. To ensure steady Rabi frequencies on each ion, there should be no power fluctuations over time on any given channel. Such fluctuations can be caused by fluctuations in the DC voltage provided to power the channel amplifier, so we used low-ripple Acopian power supplies. For the large amplifier, we used a Gold Box A24H850, which has a 0.25 mV ripple and provides +24VDC; for the pre-amplifier, we used 5EB150, which has a 1 mV ripple and provides +5VDC. For additional protection, bypass capacitors were installed on each power supply, the DC bulkhead 29 connection on the control box, and the DC power connection to the amplifier. These bypass capacitors of 10 µF eliminate fluctuations by acting as a resistor to ground for any AC components that may be in the signal, while the DC component sees the capacitor as an open circuit and instead continues to the amplifier. It should be noted that small power differences (< 1dB) from one channel to the next are acceptable, as these differences can be measured and accounted for. Another important consideration is that there cannot be phase variations be- tween channels when individual phase control is not available. This requires that the path length of the RF signal be the same for all channels. Hence, we must be certain to use identical components for every channel, and same-length coaxial cables to connect corresponding components. Additionally, RF components were chosen that had very small phase imbalances. 2.3.4 Optics for Individual Addressing In order to the use the 32-channel AOM for individual optical addressing, an optical system must be designed and constructed to split 355nm light from the pulse laser into 10 beams, focus those beams into the AOM with 450 µm beam spacing between adjacent acoustic modes and < 50 µm waist on each beam, and then focus those beams onto the ions in the trap, which requies 5 µm beam spacing and < 2 µm beam waists. The problem is additionally complicated by the practical constraints of a lab: the limited real estate on an optical table limits the allowable beam path length, and due to commercial optics availability and quality, the total 30 Figure 2.11: An image of the 10 beams from the diffractive optic element, demon- strating even beam spacing, even power distribution, and very low-power higher- order beams. beam diameter should not exceed 2 inches at any point on the beam path. The 10-way beam splitting problem is solved with a diffractive optic element (DOE) from Holo-Or, part number MS-244-U-Y-A. It splits one incoming collimated Gaussian beam with waist at least 0.24 mm into 10 evenly spaced, collimated Gaus- sian beams of equal power and with the same waist as the input beam (see Figure 2.11). It has an observed diffraction efficency of > 56% and a diffraction angle of 4.3 mrad. Since the optical system to be designed has so many constraints, it was im- portant to perform careful analysis of the optical path for each solution attempted. Additionally, both the spacing between the beams and the waist of each beam are important. Consequently, ray transfer matrix analysis was used to simultaneously keep track of beam spacing and beam widths. In the ray picture, we assume that each beam acts like a ray, and then keep track of one beam’s displacement x from the optical axis to determine the spacing between beams. The optical system starts with the 10-way diffractive optic beamsplitter (10BS) splitting the incoming beam with a diffraction angle of θ = 4.3 mrad. For simplicity, we will examine the size and displacement of a beam that is deflected θ = 4.3 mrad from the optical axis, which is equivalent to examining the size and spacing between 2 beams that are each deflected θ = 2.15 mrad on either side of the optical axis, which will be the 31 closest 2 beams for a diffractive optic element producing an even number of beams. The initial ray vector is ( ) ( ) x0 0 r0 = = . (2.4) θ0 0.0043 In the Gaussian picture, we instead assume each beam is a Gaussian beam, and use the complex beam parameter q to keep track of one beam’s width and calculate relevant waists. For a Gaussian beam of wavelength λ = 0.000355 mm, radius of curvature R = ∞ for collimated beams, refractive index n ≈ 1 for air, and radius w, 1 1 iλ 0.000113i = + = . (2.5) q0 R πnw2 w2 Using the ABCD matrix method, we can find the final complex beam parameter for a given system using Aq0 +B q1 = (2.6) Cq0 +D and extract the beam radius from q1. For both the ray and Gaussian pictures, we use ( ) ( ) A B 1 d = (2.7) C D 0 1 to represent a distance d through free space, and ( ) ( ) A B 1 0 = (2.8) C D − 1 1f to represent a thin lens with focal length f . A naive solution would involve simply focusing the 10 beams into 10 channels of the AOM, and then using one lens to 32 (a) (b) Figure 2.12: The solution for the optics system. The system starts with the 10-way diffractive optic element beamsplitter (10BS) and is then focused into the 32-channel AOM (32AOM) with the correct spacing. After that, the system uses alternting pairs of lenses to bring the separate beams closer together, while expanding each beam’s radius to then focus very sharply with the final objective f5. The beam radius is expanded between f3 and f ′ 2 and between f ′ 3 and f4, while the individual beams are kept parallel. The individual beams are brought closer together between lenses f and f , f ′2 3 2 and f ′ 3, and f4 and f5, where each beam is also kept collimated. For appropriate lens selection, this will achieve our desired waist of 1 µm and spacing of 5 µm at the ions. (a) The Gaussian solution shows the width of a single beam passing through the lens system. (b) The ray solution shows the spacing of 2 beams (one red, one orange) passing through the lens system. image the AOM onto the ions. However, this solution requires that the imagining lens have a focal length of ∼ 8 meters, which is entirely impractical. Instead, using the ray transfer analysis, we found a solution that uses two telescope arrangements after the 32-channel AOM to achieve the desired waist and spacing on the ions, while reducing the total length of the system. The solution is shown in Figure 2.12. The first important step in the optics system is focusing the 10 beams into the AOM onto 10 separate channels. Since the diffractive angle θ = 4.3 mrad for the 33 10-way beamsplitter is small, we can write xAOM = f1θ (2.9) for the spacing of the beams at the 32-channel AOM. So f1 must be chosen to achieve the correct beam spacing at the AOM channels. After the AOM, the system’s magnification can be calculated, f4f ′ 2f m = 2′ = 28 (2.10)f1f3f3 where ray transfer matrix analysis shows that m = 28 is the ideal magnification to yield a 5 µm spacing at the ions given a choice of a 32.7 mm objective (see next paragraph.) Much like in Equation 2.9, the final spacing is f5θ xf = = 5µm. (2.11) m Taking into account contraints from commercially available lenses, the final lenses chosen are listed in Table 2.1. f1 was chosen experimentally; the 90mm Thorlabs lens allowed us to successfully line up 10 parallel beams into 10 channels. For an input waist of 0.2 mm to the 10-way beamsplitter, this yields a waist of 55 µm; while this yields a spot slightly larger than the AOM channel, we do not see any spillover since the channels are 450 µm apart, and the efficiency is still high enough that the solution is workable. For the telescopes after the 32-channel AOM, the magnification is m = 74.2, which is close to the ideal from equation 2.10. 34 Lens Focal length at 355 nm Lens used f1 87.0 mm 90 mm Thorlabs plano-convex f2 96.6 mm 100 mm Thorlabs plano-convex f3 -29.0 mm -30 mm Thorlabs plano-concave f ′2 120.8 mm 125 mm Thorlabs plano-convex f ′3 -57.3 mm Doublet: -150 mm +meniscus and -100 mm , Thorlabs plano-concave f4 483.4 mm Doublet: 1000 mm and 1000 mm +meniscus, Thorlabs plano-convex f5 32.7 mm Ronar-Smith objective, 0.25 NA, working distance 32.7 mm Table 2.1: Final focal lengths for lenses chosen for the focusing lens system. The final lens, f5, is a triplet objective made by Ronar-Smith with a 0.25 NA and working distance 32.7 mm that was chosen because it will minimize abberations on the ion. Since it will be focusing a large-diamater beam to a very small focal point, it is important that f5 be chosen carefully. The total system length, from 10-way beamsplitter to ions, is 1.59 m. To save more space, it is possible to move the objective f5 back toward f4, thus reducing the total length to slightly over 1 m. This is because f4 is a very slow lens; the individual beams have a very small deflection angle as they converge after the lens, and so placing the objective closer to f4 will still yield beams very close to parallel at the ions. The beam spacing will still be correct at the working distance of the objective. This yields the additional advantage that the individual beams can be focused on the ions without having to move the entire optics system; instead, small adjustments in the objective will ensure the individual beams are focused at the ions. While the system will no longer be telecentric, meaning the beam spacing will vary with objective position, the amount of variance is negligible. The lens system was assembled on the table (see Figure 2.13) and a camera 35 placed at the objective focus to image the results. The assembly yielded 10 beams separated by 5 µm and with waists of 1 µm, as shown in Figure 2.14(a-b). Once the optics setup was achieved, a retroreflector was added to the global beam arm of the Raman beam setup to adjust and match the optical path length of the global beam arm to the individual beam arm. It is important to note that while the individual beam spacing is fixed, the ion spacing can be adjusted using the five DC electrodes on the blade trap; this will allow the beams to be aligned as well as possible to the ions in the trap. Coarse beam alignment was achieved through a rough method of observing direct scatter off the trap blades to align the Raman beams to the middle of the trap. Daily fine-tuning consists of optimizing the alignment by maximizing the observed Rabi frequency. A higher-resolution 2D profile of 5 focused beams is shown in Figure 2.14(c), which was imaged using the ion response. With one ion in the trap, the individual beams were moved across the ion and the Rabi rate on the ion, which corresponds to the square root of the optical intensity seen by the ion, was measured. The beams were moved by moving the final imaging objective, which is mounted on a translation stage controlled by picomotors along all 3 axes. The picomotors allow the objective to be moved in steps of about 20 nm without opening the protective enclousure around the Raman optics, which reduces air currents that induce beam pointing instability and therefore noise on the experiment’s coherent operations. These picomotors are also used to fine-tune the individual beam alignment on the ions, which drifts on a daily timescale. The final ion positions relative to the center ion were [−10.74,−5.05, 0, 5.16, 10.85] 36 Figure 2.13: Photograph of the 32-channel AOM on the optics table. Beam path is shown in blue, along with the 10-way diffractive optic element (10BS), 32-channel AOM, and the ion chamber. (This is a slightly earlier version of the setup, with some different lenses.) (a) (b) (c) Figure 2.14: Images of focused beams to be applied to ions. 10 beams are available to shine on 10 ions; they can be turned off and on independently to achieve arbitrary individual addressing. (a) Camera image of 10 focused beams with 1 µm waist and 4.5 µm spacing; this image was taken with an earlier lens setup, and was later changed to a 5 µm beam spacing. (b) Camera image of beams for ions 1 and 2 turned on, and the rest off. (c) 2D profile across 5 beams focused to <1 µm waist and separated by 5 µm, imaged by the ions themselves by moving the beams across the ions and measuring the resulting Rabi frequencies experienced by the ions. 37 µm, yielding spacing differences of [5.69, 5.05, 5.16, 5.69] µm, measured with an error of about 0.16 µm. 2.4 Individual Addressing System Upgrades Some upgrades to the individual addressing system have been added since the results reported in [19, 20], which will apply to some but not all of the results reported in this thesis. The Grover search algorithm discussed in Chapter 6 was performed before the following upgrades, but several more results, including those in Chapter 7 were made possible by the following upgrades. For the earlier work on the experiment, a single AWG output provided RF control to the global beam AOM, supplying sidebands and pulse-shaping to con- trol two-qubit entangling gates. The multi-channel AOM controlling the individual beams was controlled by a single RF source from the beatnote lock output; only on/off TTL control was available for individual beams. A significant upgrade was enabled by the addition of a 4-channel AWG2, bring- ing to 5 the total number of AWG channels available. This allowed for total phase, frequency, and amplitude control for the signal on each ion, not just on/off TTL control. To use these new capabilities, the AWG controls were transferred to the RF signals to the individual beam channels on the multi-channel AOM, and the beat- note lock signal was applied to the AOM on the global beam. The controls for the AWG were then integrated into the Igor control program using the NI-VISA com- 2Model WX1284C-1 1.25 GS/s Four Channel Arbitrary Waveform Generator, PN: 126182, Tabor Electronics Ltd. 38 munication protocol; a suite of utility functions were added in a separate procedure file that managed communication with the device and ensured uploaded waveforms adhered to the device’s parameters. The additional, individually addressed controls provided by the AWG enabled several procedures previously impossible to implement, including simultaneous ro- tations, the simultaneous π pulses on different sidebands required to implement phonon hopping and blockades in [29], and the parallel two-qubit operations demon- strated in Chapter 7. Improved phase control for individual ions also enhanced ex- perimental performance for all subsequent work by better compensating for differ- ences in accumulated phases as operations are performed on different ions. Finally, the arbitrary phase control on each ion allowed for the classical implementation of Z rotations, further discussed in Section 3.1.2. For further scaling in ion traps, however, continuing to rely on AWG control for individual addressing is not ideal; at a cost of about $10,000 per channel, this is not cost effective. Efforts such as Sinara [54] have made important progress in this area, designing alternative control signal generators using an FPGA and a fast DAC to create an effective dynamic AWG; this design is estimated to lower costs by an order of magnitude to ∼$1000 per channel. Using only commercially available non- AWG signal generators to generate control signals is unfortunately not currently possible, due to the need to change frequencies while maintaining phase coherence. Direct control over phase coherence also allows for compensation of effects like Stark shifts, which can reduce fidelities if left unaddressed. 39 2.5 Expansion to 7-Ion System One experiment performed over the course of this thesis work required a 7-ion system (specifically [31]), and we anticipate performing more 7-qubit experiments in future. This presented several challenges. First and foremost was arranging for 7 ions positioned in the trap to align well with the individual addressing beams, which have fixed equal spacing due to the optical setup (see Section 2.3.4). As discussed in Section 2.2, perfectly equal spacing of more than 3 ions in this trap is not feasible. While 5 ions are still pretty close to equal spacing, more ions in the trap will have bigger and bigger spacing differentials between the outer and inner ions. Consequently, 7 ions will be less well-positioned relative to the individual ad- dressing beams than for 5 ions. However, this can be mitigated by instead trapping 9 ions, and ignoring the outer 2 ions. The inner 7 ions are then pushed together by the sacrificial outer ions, resulting in a subchain of 7 ions with smaller spacing differentials than would occur with just 7 ions in the trap. With this strategy, several candidate DC voltage sets based on simulations were tested on 9 ions in the trap to find the best experimental values. Once a voltage set was chosen, we next needed to align the individual beams on the not- quite-equally-spaced center 7 ions. Some experimentation demonstrated that the best alignment strategy was to try to maximize the alignment of all ions, but to particularly prioritize ions 2-6. This yielded reasonably comparable Rabi flops on all ions, further discussed in Section 3.1.1. The selected ion spacing yielded motional sidebands at [2.9424, 2.9604, 2.9774, 3.0044, 3.0149, 3.0264, 3.0384, 3.0534] 40 MHz detuned from the carrier. The ion positions relative to the center ion were [−23.24,−16.49,−10.69,−5.21, 0, 5.27, 10.59, 16.49, 23.40] µm, corresponding to dif- ferential spacings of [6.76, 5.80, 5.48, 5.21, 5.27, 5.32, 5.90, 6.91] µm, measured with an error of about 0.16 µm. As discussed in Section 2.4, we now have 5 AWG channels for full phase, frequency, and amplitude control on each individual addressing beam, but expanding to 7 qubits means we no longer have 1 AWG channel per qubit. We instead had some qubit controllers share AWG channels. For our initial setup, we had ions 1 and 7 share an AWG channel, and ions 2 and 6 share an AWG channel. This consisted of installing a TTL-controlled switch on the two shared AWG outputs that directed the AWG output to one qubit or the other, and adding software to the control program to seamlessly handle the shared controllers by queueing control sequences to the AWG and managing TTL timing accordingly. The tradeoff was that ions 1 and 7, and ions 2 and 6, could not be controlled at the same time, ruling out any 2-qubit gates between those pairs. For purposes of [31], we did not need those gates, so this was sufficient. Future experiments may require these gates, in which case we can set up a slightly more complicated TTL-controlled switching system to supply two different AWG channels to a given pair of ions to perform a 2-qubit entangling gate. Finally, we also wanted to maintain our existing 5-qubit infrastructure for ongoing 5-qubit experiments, so the control program was modified to allow easy switching between controls for 5 and 7 qubits; a few button pushes change the relevant parameters in the control program, 7 wires for the individual addressing RF signals need to be shuffled, and the alignment of the readout objective 41 is adjusted slightly to better align the ions to the PMT array. 42 Chapter 3: Quantum Control In this chapter I discuss our two mechanisms for quantum control: single qubit rotations, or R gates, and two-qubit entangling interactions, or XX gates. These two operations form the gate set native to our hardware, and together they form a universal gate set that can implement any desired operation. 3.1 Single-Qubit Rotations R In the matrix representation of quantum computation, we define the two qubit states as ( ) ( ) 1 0 |0〉 = |1〉 = . (3.1) 0 1 Superposition states can be represented as points on the surface of the Bloch sphere (Figure 3.1), where |0〉 and |1〉 define the poles of the Z axis of the sphere. The polar angle θ represents the angular distance from the Z axis, and the azimuthal angle φ gives the projection in the XY plane running orthogonal to the Z axis. We define the azimuthal angle as φ = 0 at the X axis; φ = π at the Y axis. 2 Consequently, coherent transformations of a qubit can be described as rotations on 43 z y x Figure 3.1: Quantum states can be represented as points on the surface of the Bloch sphere with the radial variables (θ, φ). Single-qubit operations can be represented as rotations performed about a selected axis on the Bloch sphere. the Bloch sphere. Rotations of an angle θ about an axis described by the azimuthal angle φ in the XY plane are given by ( ) cos θ −ie−iφ sin θ R(θ, φ) = 2 2 , (3.2) −ieiφ sin θ cos θ 2 2 with rotations about the X and Y axes given by ( ) cos θ −i sin θ R (θ) = R(θ, φ = 0) = 2 2x (3.3) −i sin θ cos θ 2 2 and ( π) ( )cos θ − sin θ Ry(θ) = R θ, φ = = 2 2 . (3.4) 2 sin θ cos θ 2 2 Such rotations can be implemented on the experiment by applying a Raman pulse and adjusting the duration and phase of the pulse accordingly. The duration of 44 1 Ion 1 0.9 Ion 2 Ion 3 0.8 Ion 4Ion 5 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 1 3 5 7 9 11 13 15 Duration ( s) Figure 3.2: Sideband-cooled Rabi flops on 5 ions. the pulse governs the magnitude of the rotation θ, to which it is directly proportional by way of the Rabi frequency. The duration required to transfer all population from the |0〉 to the |1〉 state, which corresponds to a R(θ = π) rotation, is known as the π time. Rotation calibration consists of measuring the π time for each ion, and using it to calculate the pulse duration necessary to implement each rotation: τ θθ = · τπ π. The rotation angle φ is set by the phase of the control signal, with φ = 0 defined to be the X axis. 3.1.1 Rabi Flopping on 5 or 7 ions With individual Raman addressing beams in place, we can coherently cycle between the |0〉 and |1〉 states of each ion using Rabi flopping. Sideband-cooled Rabi flopping on 5 ions is shown in Figure 3.1.1. Applying Raman pulses for a 45 Probability |1 1 Ion 1 0.9 Ion 2 Ion 3 0.8 Ion 4Ion 5 Ion 6 0.7 Ion 7 0.6 0.5 0.4 0.3 0.2 0.1 0 1 6 11 16 21 26 31 36 41 46 51 56 61 66 Duration ( s) Figure 3.3: Sideband-cooled Rabi flops on 7 ions in a 9 ion chain at low power. Ion 1 is directly behind ion 5 in the plot. controlled duration with a controlled phase allows us to implement coherent rotation gates. Typical rotation fidelities are 98-99%. Rotation fidelities are limited largely by intensity fluctuations on the Raman beams and imperfect cooling. Intensity fluctuations also inhibit our ability to accurately measure and calibrate the π times for gate implementation. Rabi flopping on 7 ions is shown in Figure 3.1.1. after sideband cooling, the third Rabi flop peak was still above 90% bright for all ions, which is comparable to sideband-cooled Rabi flopping with 5 ions. The Rabi flops decay faster for ions that are less well aligned to the individual beams, which is primarily due to beam pointing fluctuations on the individual addressing beams; while this is still manage- able for 7 ions, this problem gets worse as more ions are added to the chain and 46 Probability |1 the non-equally-spaced ions become less and less well matched to the equally-spaced individual beams. 3.1.2 Z Rotations Qubit rotations about the Z axis on the Bloch sphere, given by ( θ ) e−i 2 0 Rz(θ) = , (3.5)θ 0 ei 2 can be implemented on the ions in one of 2 ways. The first way is through a composite series of rotations about axes in the XY plane, as discussed in Section 5.1. The second way is by adjusting the phase of individual controllers relative to the ion. With this method, instead of performing a rotation on the qubit and moving it on the Bloch sphere, the controller is rotating the Bloch sphere around the qubit. This effectively rotates the XY plane around the Z axis and redefines the locations of the X and Y axes, and is equivalent to rotating the ion around the Z axis of the Bloch sphere. Subsequent operations are then implemented relative to the rotated axes. The upgrades discussed in Section 2.4 permitted individual phase offsets to be applied differentially to individual ions. Because they can be implemented purely classically by simply shifting the phase of the frequency source, the error introduced by Z rotations is the same as the error of the frequency source itself. Consequently, Z rotation contributions to the error in fidelity during an algorithm are negligible compared to the error introduced by an R(θ, φ) of the same θ and can be disregarded. 47 This has implications for composite sequence optimization further discussed in Sec- tion 5.2.1. 3.2 Two-Qubit Entangling XX Gates We implement two-qubit entangling gates using a Mølmer-Sørensen scheme [41, 52, 55, 56] that creates an XX spin-spin interaction between two ions using the normal motional modes of the chain as an information bus. In this experiment, we couple to the transverse phonon modes [57,58] to drive the interaction. The two-qubit XX entangling gate is   cos(χ) 0 0 −i sin(χ)  0 cos(χ) −i sin(χ) 0 XX(χ) =   , (3.6)0 −i sin(χ) cos(χ) 0  −i sin(χ) 0 0 cos(χ) where the parameter χ can be varied continuously, 0 < χ ≤ π , by adjusting the 4 overall power applied to the gate. The gate is maximally entangling for χ = ±π , 4 where ( π)       1 0 0 −i 1 0 0 i1 0 1 −i 0  ( π) 1 0 1 i 0 XX χ = = √   , XX χ = − = √   . 4 2  0 −i 1 0  4 2 0 i 1 0 −i 0 0 1 i 0 0 1 (3.7) 48 When applied to two ions initilized to |00〉, this yields a maximally entangled state: ( π) 1 XX (χ = √ ) |00〉 = (|00〉 − i|11〉)4 2π XX χ = |00〉 √1= (|00〉+ i|11〉) . (3.8) 4 2 The sign of χ for a given gate depends on the interactions between the two ions in question and the modes of motion primarily excited by the gate to produce the spin-spin entanglement. The sign of the gate parameter χ is measured empirically for each XX gate pairing. Accommodation for the difference in χ signs is handled by the compiler at the gate library level (see Chapter 5). For ease of discussion, I will only reference the χ = π case for the remainder of this chapter. 4 3.2.1 Pulse-Shaping Scheme We implement entangling XX gates using a pulse-shaping scheme designed to ensure that the motional modes are fully detangled from the qubits at the end of the gate, leaving only spin-spin entanglement [53, 57, 59]. Two ions in a chain of N ions are illuminated with red and blue sidebands that are detuned near the normal transverse modes of motion and couple the modes to the qubit spins. The pulse- shaping scheme is also robust against detuning errors, and the number of pulses needed to implement gate solutions grow linearly with the number of ions in the chain. The detuning µ and gate time τ are independent parameters that can be chosen before calculating a gate pulse shape, and a unit fidelity gate is possible at any detuning [53]. 49 Because we couple to the normal motional modes of the chain to induce XX gates, all ions participate in the motional modes, and so couplings are available between all possible ion pairs. This fully connected architecture has significant ad- vantages over less-connected architectures, which require additional gates to transfer quantum information between unconnected qubits [27]. On this experiment, we have also implemented entangling XX gates using a pulse-shaping scheme that modulates the frequency of the gate driver, rather than the amplitude [28]. In order to entangle ions i and j in a chain of N ions with N motional modes ωk using red and blue sidebands with detuning µ, we have the following 2-qubit unitary: ( ∑ ∑ ) U xXX(τ) = exp i φi(τ)σi + i χij(τ)σ xσxi j ( [ i i [70] and Rigetti’s Forest [71] have started to address these needs. While the control program described here certainly has limitations, and more 77 streamlined approaches will be needed as the size of quantum computers grows, this can certainly serve as a demonstration that, moving forward, quantum computer control software will require modularity, flexibile construction, and well-thought- out program design. 78 Chapter 5: Gate Library Now that we have a programmable quantum computer with high-quality qubits (Chapter 2) and coherent single- and two-qubit operations yielding a universal gate set (Chapter 3), the next step is to build a gate library. These gates are used in Chapters 6-7 and [30, 31] to construct complex, multi-qubit algorithms, and the modular nature of these gates will be crucial to flexible ion-qubit mappings and allow great ease of implementation. In this chapter, I present detailed circuit diagrams for all of the operations used in the course of the research for this thesis, shown in terms of the R(θ, φ) and XX(χ) gates directly implemented by the experiment. 5.1 Single-Qubit Gates From Chapter 3, the single-qubit rotation is defined as ( ) cos θ −ie−iφ sin θ R(θ, φ) = 2 2 . (5.1) −ieiφ sin θ cos θ 2 2 Rotations about the X-axis (Rx(θ)) are achieved by setting φ = 0, and rotations about the Y -axis (Ry(θ)) are achieved by setting φ = π . Rotations about the Z 2 axis (Rz(θ)) can be constructed from three rotations about axes in the XY plane, 79 (a) |q〉 Rz(θ) = Ry(π ) Rx(θ) Ry(−π )2 2 (b) |q〉 H = Ry(π )2 R π x(π) = Rz(π) Ry( )2 Figure 5.1: (a) Rz(θ) gate implementation using Rx(θ) and Ry(θ) gates. (b) Hadamard (H) gate implementations using Rx(θ), Ry(θ), and Rz(θ) gates. as demonstrated in Figure 5.1(a). A Hadamard (H) gate, ( ) √1 1 1H = , (5.2) 2 1 −1 can be performed with 2 rotations, as shown in Figure 5.1(b), with a few implemen- tation options. The fidelities of these composite rotations gates are comparable to the fidelities of individual rotations, typically 98-99%. 5.2 Multi-Qubit Circuit Construction From Chapter 3, the two-qubit XX entangling gate is    cos(χ) 0 0 −i sin(χ)  0 cos(χ) −i sin(χ) 0 XX(χ) =   . (5.3)0 −i sin(χ) cos(χ) 0 −i sin(χ) 0 0 cos(χ) The parameter χ can be varied continuously, 0 < χ ≤ π , by adjusting the overall 4 power applied to the gate, but the gates used here require only χ = ±π or χ = 4 80 ±π . The ability to vary χ continuously has proved useful in implementing circuits 8 for other applications not discussed in depth here, including the quantum Fourier transform [19] and a Bayesian quantum game [32]. The gate is maximally entangling for χ = ±π . 4 Two-qubit XX gates are combined with rotation R gates to construct com- posite gates. The parameter χ can be positive or negative, depending on what ion pair is chosen and the particulars of the pulse segmentation solution chosen for the ion pair in question; the sign of χ (sgn(χ)) is determined experimentally for each ion pair. Consequently, some composite gate circuits include rotations with parameters that depend on sgn(χ); all composite gates are constructed to account for both signs of χ for each XX gate used, so that composite gates are always fully modular. Composite gates are constructed by starting with known circuits, converting constituent parts into R and XX gates using lower-level constructions, and then optimizing the circuit. First, the number of XX gates is minimized (as in the Toffoli- 3 gate, described in Section 5.4). Second, the single-qubit gates are optimized by minimizing the sum of all rotation angles θ, as this minimizes the total time for the experiment. Several composite circuits used in algorithms performed in this thesis are pre- sented in this chapter. Other quantum algorithms may be implemented on this system in a similar fashion. First, decompose the algorithms subroutines into high- level circuits. Second, optimize those circuits to minimize the number of two-qubit interactions required. Third, decompose the high-level circuits into physical-level R and XX gates. Finally, perform further optimizations to first minimize the num- 81 ber of two-qubit XX gates required, and then to minimize the total rotation time (the sum of all rotation angles θ) across all R gates. However, the optimization of quantum circuits is in the QMA-Hard complexity class, which is the quantum ana- log of the classical NP-Hard complexity class. We therefore anticipate that future improvements in algorithm design, circuit synthesis, and circuit optimization tech- niques may result in more efficient circuit implementations, facilitating increased experimental performance. 5.2.1 Optimization Strategy Adjustments with Experimental Up- grades This chapter features composite gates developed for use before and after the improvements described in Section 2.4 were implemented, and consequently are opti- mized slightly differently. In particular, the introduction of individual ion phase con- trol allowing for the purely classical implementation of Z rotations (see Section 3.1.2) leads to different optimization strategies for minimizing rotations. Pre-upgrade, any Z rotations had to be implemented via a combination of several rotations in the XY plane. Consequently, composite gates were optimized by minimizing the total rota- tion angle θ for all rotations R(θ, φ) in the XY plane; Z rotations were to be avoided where possible, and compiled into X and Y rotations for further optimization when not. Post-upgrade, however, Z rotations became effectively error-free to perform relative to R(θ, φ) rotations. Therefore, when optimizing a given composite gate, the total rotation angle θ over all R(θ, φ) rotations can be further minimized by 82 (a) |q 〉 • R (απ ) R (−απc y y ) Rz(−π )2 = XX(απ 2 2 ) | 〉 4qt Rx(−π )2 (b) |qc〉 • R π π πy(− ) R2 x(− ) Ry( )2 2 = XX(απ ) |q 〉 R (απ ) R (απ 4t Z y x ) Ry(−απ )2 2 2 Figure 5.2: χct is the parameter for the XX gate between the two qubits. Let α = sgn(χct). (a) CNOT gate implementation using XX(χ), Rx(θ), Ry(θ), and Rz(θ) gates. (b) Controlled-Z gate implementation using XX(χ), Rx(θ), and Ry(θ) gates. employing Z rotations instead wherever possible. In particular, the following Pauli rotation identities were very useful for converting between X, Y , and Z rotations to minimize rotations other than Z: (π) ( ) ( ) ( ) Rz(θ) = Ry( ) · π π π Rx(θ) ·Ry(− ) = Rx(− ) ·Ry(θ) ·Rx (5.4)2 2 2 ( 2π π π π) Rx(θ) = Rz ( ) ·Ry(θ) ·Rz (− )= Ry2 2π π ( − 2 π) ·Rz(θ) ·Ry ( 2π) (5.5) Ry(θ) = Rx ·Rz(θ) ·Rx − = Rz − ·Rx(θ) ·Rz (5.6) 2 2 2 2 It is also useful to note that X rotations commute with XX gates, and can therefore be moved around more to minimize rotations. In this chapter, I will identify which composite gates were optimized under which set of circumstances. 5.3 Two-Qubit Composite Gates The controlled-NOT , or CNOT , gate, is a two-qubit interaction used as the building block for many quantum algorithms. A control qubit |qc〉 acts on a target 83 qubit |qt〉 by flipping the target qubit’s state if the control qubit is in the |1〉 state; otherwise, it does nothing. This has the unitary evolution operator  1 0 0 00 1 0 0UCNOT =  . (5.7)0 0 0 1 0 0 1 0 Another useful two-qubit interaction is the controlled-Z, or C(Z), gate, which flips the phase of the two-qubit state if the two qubits are in the |11〉 state:  1 0 0 00 1 0 0 UC(Z) =  . (5.8)0 0 1 0 0 0 0 −1 Circuits for the CNOT and (C()Z) gates are shown in Figures 5.2(a-b) respectively. They each require one XX π gate and several rotations. Both gates were first 4 demonstrated on this system in [19], with typical fidelities of 96-98% [19, 20]. The upgrades described in Section 2.4 have resulted in typical CNOT gate fidelities increasing to 98-99%. In this work, the CNOT gate will be used in Chapters 6 and 7, while the controlled-Z gate will only be needed for Chapter 6. 5.4 Toffoli Gates Next, we scale up to a three-qubit composite gate by implementing a controlled- controlled-NOT (C2(NOT )), or Toffoli-3, gate. The Toffoli gate [72] requires two control qubits (q1 and q2) and one target qubit (qt). Like the CNOT gate, it flips 84 • • • • • = • • U V V † V Figure 5.3: Any doubly-controlled unitary U can be constructed out of 5 two-qubit interactions using CNOT gates and controlled-V operations, where V 2 = U [63]. the target qubit state if and only if both control qubits are in the |1〉 state:   1 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 1 0 0 0 0UT3 =   0 0 0 0 1 0 0 0 . (5.9)0 0 0 0 0 1 0 00 0 0 0 0 0 0 1 0 0 0 0 0 0 1 0 ( ) Our(T)offoli-3 gate is constructed from 5 two-qubit gates (three XX π and 8 two XX π gates) in a manner similar to the Toffoli gate demonstrated in [73]. 4 As shown in Figure 5.3, any doubly-controlled unitary C2(U) operation can be performed with 5 two-qubit interactions (two CNOT s, two C(V )s, and one C(V †)) if a controlled-V operation is available such that V 2 = U [63]. Since [ (π)]2 (π) XX = XX , (5.10) 8 4 we can add single-qubit rotations to construct a Toffoli-3 gate with minimal use of two-qubit gates, as shown in Figure 5.4(a). This compares favorably to t(he)6 two-qubit gates that would be necessary if only CNOT (or equivalently, XX π ) 4 gates were available. These constructions also provide for the implementation of 85 (a) |q π1〉 • Ry(−β ) Rx(β 3π ) XX(β π )2 4 8 |q2〉 • = Ry(−γ π ) R πx(γ )2 2 XX(γ π ) 8 |qt〉 R (πx ) XX(β π )4 8 |q1〉 · · · XX(απ ) |q2〉 4 · · · R(−2π , (γ+1)π − P ) 3 2 R(−αβγ 2π , (αβ+1)π − αβγP ) 3 2 |qt〉 · · · |q1〉 · · · R πy(β ) π 2XX(α ) |q2〉 · · · 4 R(π,−αβγ π ) XX(γ π 4 ) |qt〉 · · · 8 (b) |q1〉 • Ry(−β π ) R 3πx(β ) XX(β π )2 4 8 |q2〉 • = Ry(−γ π ) R (γ πx )2 2 XX(γ π ) 8 |qt〉 Z Ry(π ) Rx(π ) XX(β π )2 4 8 |q1〉 · · · XX(απ ) |q 〉 4· · · R(−2π2 , (γ+1)π − P ) 3 2 R(−αβγ 2π , (αβ+1)π − αβγP ) 3 2 |qt〉 · · · |q1〉 · · · Ry(β π ) XX(απ 2 ) |q2〉 · · · 4 R(π,−αβγ π ) 4 XX(γ π ) |qt〉 · · · 8 R πy(− )2 Figure 5.4: Three-qubit composite gates using XX(χ), Rx(θ), Ry(θ), a(n√d R)(θ, φ) gates. Let α = sgn(χ12), β = sgn(χ1t), γ = sgn(χ2t), and P = arcsin 2 . (a) 3 Toffoli-3 gate implementation. (b) Controlled-controlled-Z (CCZ) gate implemen- tation. 86 1 1 0.5 0 0.5 000 111 111 000 detected input 0 Figure 5.5: Measured truth table for a Toffoli-3 gate. The average process fidelity is 89.6(2)%, corrected for a 1.5% average state preparation and measurement (SPAM) error. C2(Z) gates, which can be constructed by adding a few single-qubit rotations to a Toffoli-3 gate (see Figure 5.4(b)) and has the unitary evolution operator  1 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 1 0 0 0 0 0  0 0 0 1 0 0 0 0 UC2(Z) =    0 0 0 0 1 0 0 0   . (5.11)   0 0 0 0 0 1 0 0 0 0 0 0 0 0 1 0  0 0 0 0 0 0 0 −1 For all circuits, the single-qubit rotations are further optimized to minimize total rotation time [74], optimized under pre-upgrade constraints. Here, we show results for a Toffoli-3 gate, with a process fidelity of 89.6(2)% (see Figure 5.5). Toffoli-3 gates have been previously performed in NMR systems [75] and ion traps [76], including this system [24,27]. We employed a limited tomography procedure to verify that the Toffoli-3 gate performed had no spurious phases on the 87 probability outputs (see Section 5.4.1). This gate will be used for the Grover search algorithm in Chapter 6. 5.4.1 Toffoli-3 Characterization We employed a limited tomography procedure to characterize the outputs of the Toffoli-3 gate performed. A global rotation into the X basis was applied to all 3 ions before and after the Toffoli-3 gate for each input (see Figure 5.6(a)): R (πy ) for the even inputs (000, 010, 100, 110) and R π y(− ) for the odd inputs (001,2 2 011, 101, 111). An ideal Toffoli-3 gate will result in an anti-diagonal input-output matrix in the Z basis when this procedure is applied. The experimental results of this verification procedure are shown in Figure 5.6(b) with an average success probability of 82.1(2)%, indicating the Toffoli-3 is faithful for arbitrary input states. 88 (a) ( ) ( ) |q 〉 R ±π1 y • R ±π( 2 ) y ( 2 ) |q2〉 Ry ±π • R ±π( 2 ) y ( 2 ) |q 〉 R ±πt y Ry ±π2 2 (b) 1 1 0.8 0.8 0.6 0.6 0.4 0.2 0.4 0 000 0.2 011 111 011 111 0 detected 000 input Figure 5.6: (a) Circuit for implementing the Toffoli-3 limited tomography procedure. The global rotations are positive for even input states and negative for odd input states. (b) Limited tomography check performed on the Toffoli-3 gate to verify phases. The average success probability is 82.1(2)%, corrected for a 2.4% average SPAM error. 89 probability |q1〉 • Ry(α1α π2 ) XX(α π1 )2 4 |q2〉 • Ry(π ) XX(α π2 )2 4 |q3〉 • = |qt〉 |qa〉 Ry(α πt ) XX(α π2 ) R(−αtπ,Qπ) XX(α π1 ) Ry(α π4 4 4 t )4 |q1〉 · · · |q π2〉 · · · XX(α2 )4 |q 〉 · · · R (−β π ) R (β π3 y 2 x )2 XX(β π ) · · · π 8|qt〉 Rx( )4 XX(α πt ) |qa〉 · · · Rx(− 8 α π2 ) XX(α π 2 ) R π 3π 2 4 y (−αt )4 Rx(αt )4 |q1〉 · · · |q2〉 · · · |q3〉 · · · R(−2π , (β+1)π − P ) XX(α π ) R(−α 2π α3αt+13 2 3 4 3αtβ , ( )π − α3αtβP )3 2 |qt〉 · · · |qa〉 · · · XX(α π3 )4 |q1〉 · · · Rx(−α1π) |q2〉 · · · XX(α π2 )4 |q3〉 · · · R(π,−α π π3αtβ ) XX(α )4 3 4 XX(β π ) |qt〉 · · · 8 |qa〉 · · · XX(α π ) R (α π ) XX(α π3 4 y t 4 2 )4 |q 〉 · · · XX(α π ) R π1 1 4 y(−α1α2 )2 |q2〉 · · · XX(α π2 ) Ry(−π )4 2 |q3〉 · · · |qt〉 · · · |qa〉 · · · R(−αtπ,Qπ) XX(α π1 ) R (α π π π π4 y t ) Rx(−α4 2 ) XX(α ) R2 2 4 y(αt )4 Figure 5.7: Toffoli-4 gate implementation using XX(χ) and R(θ, φ) gates. 90 1 1 0.8 0.8 0.6 0.6 0.4 0.2 0.4 0 0000 1111 0.2 1000 1000 1111 0 detected 0000 input Figure 5.8: Measured truth table for a Toffoli-4 gate performed with 3 controls, 1 target, and 1 ancilla qubit. The average process fidelity is 70.5(3)%, corrected for a 1.9% average SPAM error. 91 probability 5.4.2 Toffoli-4 Gate We use a related strategy to construct a Toffoli-4 gate as was done for the Toffoli-3 gate, where the Toffoli-4 gate is a triply-controlled NOT gate,  1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 00 1 0 0 0 0 0 0 0 0 0 0 0 0 0 00 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0  0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0    0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 U =   0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 T4  . (5.12)0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 00 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0   0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 00 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 Using the methods described in [77], we construct a circuit between 3 control qubits (q1, q2, and q3), one target qubit (qt), and an additional ancilla qubit (qa), shown in Figure 5.7. In the circuit diagram, α1 =(√sgn)(χ1a), α2 = sgn(χ2a), α3 = sgn(χ3a), αt = sgn(χta), β(=)sgn(χ3t), P( =) arcsin 2 , and Q = 1(4 − 3α2αt). By again3 8 using both XX π and XX π gates, we are able to save one two-qubit gate 4 8 relative to a construction limited to CNOT gates [77]. When executing the gate on the system, we report an average process fidelity of 70.5(3)% (see Figure 5.8) [24]. 92 5.5 The SWAP Gate A SWAP gate exchanges the quantum states of two qubits, like so:  1 0 0 00 0 1 0USWAP =  . (5.13)0 1 0 0 0 0 0 1 If one assumes that qubits are fixed wires, one could naively implement a SWAP gate as follows [65]: • • × . = (5.14) • × However, if one has a fully-connected system - for example, an ion-trap quantum computer where the normal modes of motion yield multi-qubit entangling gates with all possible pairwise interactions available - then one can simply swap the qubit assignments, a purely classical operation that saves us 3 two-qubit interactions. This will come in handy in Section 5.9. A comparison to another architecture, the IBM Quantum Experience, shows that our system benefits from the absence of SWAP -gate overhead due to its complete connectivity graph [27]. 5.6 Controlled-SWAP Gate 93 (a) |qc〉 • • |q2〉 × = • |q3〉 × • • (b) |qc〉 • Rx(γ π ) R (−π )2 z 2 |q2〉 × = Rz(−β π ) R (πx + (1− β)π )2 4 4 XX(β π ) XX(β π ) 4 8 |q3〉 × Ry(β π ) R π π π2 z(− ) Rx(−β + )2 2 4 |q 〉 · · · XX(γ π ) R(−2π , (γ+1c 8 )π − P )3 2 XX(απ ) |q2〉 4 · · · |q3〉 · · · XX(γ π )8 |qc〉 · · · R(−αβγ 2π , (αβ+1)π − αβγP ) XX(γ π ) R(π,−αβγ π )3 2 8 4 |q2〉 · · · |q3〉 · · · XX(γ π )8 |qc〉 · · · XX(απ ) | 〉 · · · 4q π2 Rz(−β )2 XX(β π ) | 〉 · · · 4q3 R (β πy ) R πy(−β ) Rz(−π )2 2 2 Figure 5.9: Controlled-SWAP (or Fredkin) gate implementation. (a) Controlled- SWAP gate in terms of CNOT and Toffoli gates, used for optimized circuit con- struction. (b) Controlled-SWAP gate circuit optimized for use on the experi- ment using XX(χ), Rx(θ), Ry(θ), Rz(θ), an√d R(θ, φ) gates, where α = sgn(χ12), β = sgn(χ23), γ = sgn(χ13), and P = arcsin 2 . 3 94 1 1 0.9 0.8 0.8 0.7 0.6 0.5 0.6 0.4 0.3 0.2 0.4 0.1 0 000 0.2 010 110 100 100 110 010 detected state 000 0input state Figure 5.10: Measured truth table for a Controlled-SWAP gate. The average pro- cess fidelity is 86.8(3)%, corrected for a 1% average state preparation and measure- ment (SPAM) error. The controlled-SWAP (CSWAP , or Fredkin) gate [78] is a three-qubit in- teraction that operates by swapping the last two qubits |q2〉 and |q3〉 if the control qubit |qc〉 is in the state |1〉, and has the unitary evolution operator   1 0 0 0 0 0 0 0  0 1 0 0 0 0 0 0  0 0 1 0 0 0 0 00 0 0 1 0 0 0 0 UCSWAP =   . (5.15)0 0 0 0 1 0 0 00 0 0 0 0 0 1 0  0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 1 It has been previously demonstrated on NMR [79] and photonic systems [80, 81]; this represents its first demonstration on a trapped ion system. 95 probability The CSWAP can be implemented using two CNOT gates and a Toffoli gate, as shown in Figure 5.10(a). The final composite circuit, shown in Figure 5.10(b), was constructed by assembling the pre-optimized CNOT and Toffoli circuits in Figures 5.2(a) and 5.4(a) respectively, and then further optimizing the combined rotations. The Toffoli and CNOT gates used here were optimized for use on the machine pre- upgrade, but the CSWAP was optimized for use post-upgrade; consequently, it is possible further rotation savings could be achieved were a more optimal Toffoli gate determined first. Future uses of the CSWAP may benefit from a more thorough optimization procedure taking advantage of the upgraded experimental capabilities. The CSWAP gate was used as a readout module to measure the Renyi-2 entropy in a two-site Fermi-Hubbard model system [30]. Data for all bitwise inputs and outputs is shown in Figure 5.10, with an average process fidelity of 86.8(3)%. For use in the Reyni-2 entropy measurement scheme, only the outcome of the control qubit is needed; its process fidelity is 94.0(2)%. 5.7 Using Z-Rotations to Obviate χ Sign in XX Gates With Z rotations implemented classically (see Section 5.2.1), an opportunity now arises to abstract away the considerations made( fo)r the eff(ects )of the sign of χ on the constuction of a composite gate. Since XX π 6= XX −π , a mechanism 4 4 is needed to ensure correct gate sequences. Previously, a parameter for the sign of the χ for each XX gate had to be introduced, and affected the properties of other rotations in the circuit in order to ensure the gate was fully modular and could be 96 |q1〉 Rz(π) Rz(−π) XX(+χ) = XX(−χ) |q2〉 |q1〉 XX(+χ) = XX(−χ) |q2〉 Rz(π) Rz(−π) Figure 5.11: Adding an appropriate Z rotation on one qubit on either side of an XX(χ) gate flips the sign of its χ parameter, obviating the need to consider all possible combinations of χ signs when optimizing a modular, programmable gate. When implementing a gate on a given set of ions, simply add Z rotations on either side of any XX gates on ion pairs with experimentally-verified negative χ signs. used on any qubit selection, regardless of the χ-signs of the XX gates in use. This made circuit optimization considerably more complicated, and led to some circuits (such as the Toffoli-3 in Figure 5.4(a)) that were not necessarily optimal for all possible χ sign combinations. Conveniently, it turns out that adding a classical Z rotation on the same qubit before and after an XX(χ) gate flips the sign of its χ parameter, as shown in Figure 5.11. By adding these classically-implemented rotations whenver an XX gate with a negative χ parameter is used, all XX gates are rendered identical without an error penalty, making composite gate optimzation much simpler (I optimized all of the following gate sequences by hand) and ensuring that all composite gates are always optimal. 5.8 The Square Root of CNOT In Section 7.5, to construct the full adder circuit from [82] we will need to construct a Controlled-V (C(V )) gate, where C(V ) is the square root of a CNOT gate, or C(V )2 = CNOT , and consists of a controlled-X(θ) operation. As discussed 97 (a) |qc〉 • R π π πy( ) Rx(− ) R (− )2 2 y 2 = XX(π ) |qt〉 4 R (−πx )2 (b) |qc〉 • R (−π πy ) R ( )2 y 2 Rz(π) = XX(π ) | 4q 〉 R (πt x )2 (c) |qc〉 • Ry(π ) R πx(− ) R (−π )2 y = XX(π 4 2 ) |qt〉 8 V Rx(−π )4 (d) |qc〉 • Ry(π ) Rx(π ) R πy(− )2 4 2 = XX(−π ) |qt〉 8 † R πV x( )4 (e) |qc〉 • • • • • = = |q 〉 V V V †t V † Figure 5.12: Component circuits for CNOT and its square root gates using XX(χ), Rx(θ), Ry(θ), and Rz(θ) gates.. (a) and (b) Alternate CNOT gate implementa- tions. (c) Controlled-V (C(V )) gate implementation. (d) Controlled-V † (C(V †)) gate implementation. (e) C(V ) and C(V †) gates are equivalent to the square root of a CNOT gate. 98 in Section 5.4, we can adjust the χ parameter of the XX(χ) gate to achieve such a relationship, shown in Equation 5.10. Figure 5.12(a) shows a variation on the CNOT √ gate from Figure 5.2(a) that allows for the creation of a CNOT by adjusting the χ of the XX gate, and concurrently adjusting the length of the following X rotations on each ion; the resulting circuit for a C(V ) gate is shown in Figure 5.12(c). Additionally, C(V †) is needed, shown in Figure 5.12(d); from Section 5.7, we can easily implement the XX(−π ) gate by adding Z rotations on either side of 8 the XX gate. By squaring the C(V ) or C(V †) gates, a CNOT gate is performed (C(V †) (Figure 5.12(e)). √ The unitary for the C(V ) = CNOT gate is  1 0 0 00 1 0 0 UV =  . (5.16)0 0 1(1− i) 1(1 + i) 2 2 0 0 1(1 + i) 1(1− i) 2 2 5.9 Quantum Scrambling Library Quantum scrambling is a phenomenon of interest that describes how infor- mation spreads through a system. It has direct implications for out of time order correlators (OTOCs) and has implications for the study of quantum information in black holes [83]. We have implemented scrambling unitaries on our experiment and probed their properties in several ways [31], as proposed in [84]. Here, I discuss the library of quantum circuits used for implementing and probing scrambling unitaries. The general circuit diagram to test whether a unitary U is a quantum scram- 99 bling unitary is shown in Figure 5.13. In the case of a scrambling unitary, the input state |ψ〉 will teleport to the output state |ϕ〉 with fidelity F = |〈ψ|ϕ〉|2 = 1, conditioned on the outcome of the Bell measurement. The Bell measurement it- self will yield the correct Bell state |00〉 + |11〉 with post-selection probability P . Three possible pairs of qubits can be used for Bell measurements: (3,4), (2,5), and (1,6). If successful teleportation with the correct post-selection probability P for all possible input states and all possible choices of Bell pair, quantum scrambling is verified. Using the teleportation fidelity F and the post-selection probability P , we can probe the relationship between scrambling, OTOCs, and decoherence [84]. |ψ〉 |0〉 EPR U |0〉 EPR Bell Bell Bell |0〉 |0〉 EPR U∗P |0〉 EPR |0〉 |ϕ〉 Figure 5.13: General figure for testing the scrambling properties of a unitary U . The unitary U∗P is a permutation of the unitary U∗ that permutes the first 100 and third input, such that U∗P = PU∗P , where   1 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0  0 0 1 0 0 0 0 0 0 0 0 0 0 0 1 0 P =    . (5.17) 0 1 0 0 0 0 0 00 0 0 0 0 1 0 00 0 0 1 0 0 0 0 0 0 0 0 0 0 0 1 This can also be modeled by simply swapping the inputs of qubits 4 and 6 into the unitary U∗, and swapping them back afterwards. 5.9.1 Deterministic Teleportation Protocol An alternative protocol that teleports |ψ〉 deterministically, eliminating the need to perform post-selection, can be performed using a variant of the Grover search protocol [85] and implemented as shown in Figure 5.14(a). The Grover operations, GD and GA, have the form    0 0 0 −1 G = I − 2|EPR〉〈EPR| =  0 1 0 0  , (5.18)0 0 1 0  −1 0 0 0 which is a SWAP operation and a few rotations. The circuit is shown in Figure 5.14(b). We can implement the SWAP by simplly re-assigning the qubits to the other ion; no quantum operation needed. 101 (a) |ψ〉 |0〉 EPR U |0〉 EPR GD |0〉 |0〉 EPR U∗ UT |0〉 EPR GA |0〉 |ϕ〉 (b) |q1〉 Rz (π) Rx (π) × Rz (π) |q2〉 Rx (π) × Figure 5.14: (a) General figure for testing the scrambling properties of a unitary U , using a Grover protocol requiring G. (b) Circuit for Grover protocol G. 5.9.2 Bell States and Bell Measurements For the quantum scrambling experiments presented in [31], we will need to create and measure Bell pairs. Below are the schemes used for dealing with Bell pairs, optimized under the post-upgrades capabilities. The four Bell states can be created using the following recipe (or related vari- ations) as a method of constructing the EPR pairs required by the scramble-tester 102 circuit in Figure 5.13: | +〉 √1 ( ) ( | 〉 | 〉 π π ) Φ = ( 00 + 11 ) = RZ,1 (α X) X α( |)00〉 (5.19)2 2 4 | 1 π πΦ−〉 = √ (|00〉 − |11〉) = RZ,1 (−α) XX α |(00〉 ) (5.20)2 2 4 |Ψ+〉 1= √ (| 〉 | 〉 π π01 + 10 ) = RZ,1 (α R)X,2 (π)XX α( |)00〉 (5.21)2 2 4 | −〉 √1 π πΨ = (|01〉 − |10〉) = RZ,1 −α RX,2 (π)XX α |00〉 (5.22) 2 2 4 Our native XX(χ) gates create the entangling state √1 (|00〉 − i|11〉). As such, 2 XX(π ) gates can be used by themselves to create EPR pairs. Since the resulting 4 output in Figure 5.13 simply adds a global phase to the teleported state |ϕ〉, the bare XX(π ) gates can be used without modification in the circuit. Note that proper 4 Bell measurement as described next is still necessary - we can’t use bare XX gates there. Bell measurements can be performed using the circuit shown in Figure 5.15(a) [65]. This takes as input a state |q1q2〉 in the Bell basis, so |q q 〉 = a|Φ+〉+ b|Ψ+1 2 〉+ c|Φ−〉+ d|Ψ−〉, (5.23) where a2 + b2 + c2 + d2 = 1, and outputs the result |q1q2〉 = a|00〉+ b|01〉+ c|10〉+ d|11〉. (5.24) This circuit can be optimized using our experimentally available gates R(θ, φ) and 103 (a) |q1〉 • H |q2〉 (b) ( ) ( ) |q π1〉 (Rz ) R π 2 y 2 ( ) XX απ 4 ( ) ( ) |q2〉 Rz (1− α) π Rz (α− 1) π R −π2 2 x 2 Figure 5.15: (a) Simple Bell measurement circuit implemented with a CNOT gate and a Hadamard gate. (b) Optimized Bell measurement gate sequence using R and XX gates. XX(χ) to the Bell measurement circuit shown in Figure 5.15(b). 5.9.3 Scrambling Unitary US The scrambling unitary US used for much of the data discussed in [31],  −1 0 0 −1 0 −1 −1 0   0 1 −1 0 −1 0 0 1  0 −1 1 0 −1 0 0 1  1  1 0 0 1 0 −1 −1 0 US =   , (5.25)2  0 −1 −1 0 1 0 0 1 1 0 0 −1 0 1 −1 0 1 0 0 − − 1 0 1 1 0  0 −1 −1 0 −1 0 0 −1 104 is verified to be a scrambling unitary by showing U †(X ⊗ I ⊗ I)U = X ⊗ Z ⊗ Z U †(I ⊗X ⊗ I)U = Z ⊗X ⊗ Z U †(I ⊗ I ⊗X)U = Z ⊗ Z ⊗X U †(Y ⊗ I ⊗ I)U = Y ⊗X ⊗X U †(I ⊗ Y ⊗ I)U = X ⊗ Y ⊗X U †(I ⊗ I ⊗ Y )U = X ⊗X ⊗ Y U †(Z ⊗ I ⊗ I)U = Z ⊗ Y ⊗ Y U †(I ⊗ Z ⊗ I)U = Y ⊗ Z ⊗ Y U †(I ⊗ I ⊗ Z)U = Y ⊗ Y ⊗ Z (5.26) where {X, Y, Z, I} are the Pauli operators. The circuit to implement this unitary consists of 3 XX gates followed by 3 Y Y gates; the latter can be implemented by adding rotations to an XX gate to move the interaction to the Y Y basis. Because this unitary is fully real, US = U ∗ S. The gate sequence to produce this unitary is shown in Figure 5.16(a). A circuit with a continuously-adjustable parameter θ that scans how much scrambling the unitary performs is shown in Figure 5.16(b); for θ = 0, the unitary is the identity I, and for θ = π , the unitary is the maximally- 2 scrambling unitary US. We note that the 3 XX gates in each group of consecutive XX gates can be done in any order. 105 (a) ( ) ( ) |q 〉 XX R (π1 z ) XX Rz −π2 2XX XX ( ) |q2〉 R πz Rz −π2 2 XX ( ) XX ( ) |q 〉 XX R π π3 z XX Rz −2 2 (b) |q1〉 XX Rz (θ) XX Rz (−θ) XX XX |q2〉 Rz (θ) Rz (−θ) XX XX |q3〉 XX Rz (θ) XX Rz (−θ) Figure 5.16: (a) Circuit to produce the unitary US. (b) Circuit for U(θ), with an adjustable parameter θ that goes from the id(ent)ity U(θ = 0) = I to the maximally-scrambling U(θ = π ) = US, allowing for a scan of the amount of scrambling per-2 formed by the unitary. XX indicates a XX π gate. 4 5.9.4 Scrambling Unitary UCZ An additional scrambling unitary, UCZ , is also probed in [31] for use with the Grover protocol, where    1 1 1 −1 1 −1 −1 −1  1 −1 1 1 1 1 −1 1  1 1 −1 1 1 −1 1 1 √1 −1 1 1 1 −1 −1 −1 1 UCZ = 2 2    . (5.27) 1 1 1 −1 −1 1 1 1 −1 1 −1 −1 1 1 −1 1−1 −1 1 −1 1 −1 1 1  −1 1 1 1 1 1 1 −1 The gate sequence yielding this unitary is shown in Figure 5.17(a), and is based on an arrangement of 6 Control-Z gates. An optimized gate sequence is shown in Figure 5.17(b). Additionally, a gate sequence with the same number of 106 operations as the unitary but instead performing the identity is shown in Figure 5.17(c), allowing for direct comparisons and error characterization. As with US, we note that the 3 XX gates in each group of consecutive XX gates can be done in any order. Additionally, the rotations before the first XX gate and after the last XX gate can all have their signs flipped without changing the unitary performed. This comes in handy when optimizing the overall circuit with the Grover protocol, and allows for some XX gate cancellations (see Section 5.9.6.) Finally, we note that U = U∗ = UTCZ CZ CZ . (a) UCZ |q1〉 • • H • • |q2〉 • Z H Z • |q3〉 Z Z H Z Z (b) UCZ, opti(mi)zed ( ) ( ) |q1〉 Rz (π) Ry (π) XX R (πy ) XX R π2 2 y 2XX XX ( ) |q2〉 Rz (π) Ry (π2 ) R πy ( 2 ) Ry (π2XX XX ) |q3〉 Rz (π) R πy XX R π R π2 y 2 XX y 2 (c) Identity I( ) ( ) |q π1〉 Rz (π) Ry ( ) XX Rz (π) XX R π2 y 2XX XX ( ) |q2〉 R πz (π) Ry ( ) Rz (π) R π2 y 2XX XX ( ) |q3〉 R (π) R πz y 2 XX Rz (π) XX R π y 2 Figure 5.17: (a) Basic circuit to produce the scrambling unitary UCZ . (b) Opti- mized circuit to produce the scrambling unitary UCZ . (c) Identity circuit requiring the s(am)e number of gates as UCZ , allowing for error characterization and direct com-parisons between maximally- and minimally-scrambling unitaries. XX indicates a XX π gate. 4 107 5.9.5 Classical Scrambler Some unitaries will scramble quantum information non-maximally. This means they can transform 1-body information into 3-body information in some bases, but not others. Maximally-scrambling unitaries transform 1-body information into 3- body information in all bases. One such unitary is the “classical” scrambling unitary UCS, which scrambles information in the X and Y bases, but not information in Z. Here,   1 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 1 0 0 0 0 0  0 0 0 −1 0 0 0 0  UCS = 0 0 0 0 1 0 0 0   (5.28) 0 0 0 0 0 −1 0 0 0 0 0 0 0 0 −1 0  0 0 0 0 0 0 0 −1 and we have U †CS(X ⊗ I ⊗ I)UCS = X ⊗ Z ⊗ Z U †CS(I ⊗X ⊗ I)UCS = Z ⊗X ⊗ Z U †CS(I ⊗ I ⊗X)UCS = Z ⊗ Z ⊗X U †CS(Y ⊗ I ⊗ I)UCS = Y ⊗ Z ⊗ Z U †CS(I ⊗ Y ⊗ I)UCS = Z ⊗ Y ⊗ Z U †CS(I ⊗ I ⊗ Y )UCS = Z ⊗ Z ⊗ Y. (5.29) 108 However, U †CS(Z ⊗ I ⊗ I)UCS, U † CS(I ⊗ Z ⊗ I)UCS, and U † CS(I ⊗ I ⊗ Z)UCS are not equal to any combination of three Pauli matrices, and therefore no scrambling occurs in that basis. The gate sequence yielding this unitary (controlled-Z gates connecting each pair of qunits) is shown in Figure 5.18(a), with its optimized version in Figure 5.18(b). Because this unitary only scrambles classical information, the circuit in Figure 5.13 will only teleport classical information; in other words, information about pop- ulations will be teleported from the input to the output state, but not information about the phase on each state. For an input state |ψ〉 = a|0〉 + b|1〉 on qubit 1, a measurement on the output qubit ϕ will yield a2 population in |0〉 and b2 population in |1〉 conditioned on the Bell pair |00〉 + |11〉. Additionally, the probability P of getting the Bell pair |00〉+ |11〉 will be 50% rather than 25%. (a) |q1〉 • • |q2〉 Z • |q3〉 Z Z (b) ( ) ( ) ( ) |q 〉 R (π) R −π ( ) XX π R π1 z y ( 2 ) 4 y 2XX π4 ( ) |q2〉 Rz (π) R −π ( ) R πy ( 2 ) ( ) y ( 2XX π4 ) |q3〉 Rz (π) R πy − XX π R π2 4 y 2 Figure 5.18: (a) Basic circuit to produce the unitary UCS, which scrambles classical but not phase information. (b) Optimized circuit to produce the unitary UCS. 109 5.9.6 Circuit Optimization We can eliminate several XX gates by cancelling XX gates repeated between repeated scrambling unitaries on the same qubits. Additionally, we can use prop- erties of EPR pairs that allow us to “move” gates from one half of an EPR pair to the other half, subject to a transformation. Specifically, any unitary U applied on one half of an EPR pair is the same as that unitary’s transpose UT applied to the other half (see Figure 5.19). As an example of how this improves circuit perfor- mance, Figures 5.20 and 5.21 showsthe gates eliminated with these techniques for UCZ implemented in the circuit from Figure 5.13. |q1〉 U EPR = EPR |q2〉 UT Figure 5.19: A unitary U performed on one half of an EPR pair is the same as its transpose UT performed on the other half of the EPR pair. ( ) ( ) ( )|q1〉 R πy Rz (−π) R π2 y − 2 ( ) XX π XX π 4 ( ) ( ) |q 〉 R π π 4 2 y R2 z (−π) Ry − 2 ( ) Rx (π) ( ) = XX π XX π4 4 = Rx (π) Figure 5.20: Equivalence used to eliminate XX gates in neighboring, repeated UCZ unitaries. 110 ( ) |q1〉 Rz (π) Ry (π · · ·2 ) |q2〉 EPR Rz (π) Ry (π2 ) ( ) · · ·XX π4 |q 〉 R (π) R (π3 z y ) · · ·2EPR |q4〉 Rz (π) Ry (π2 ) ( ) · · ·XX π4 |q 〉 EPR R (π) R (π5 z y ) · · ·2 |q6〉 R (π) R πz y · · ·2 EPR |q7〉 · · ·( ) R (π) R πz y · · ·2 EPR · · · · · · EPR = ( ) ( ) ( )RTy XX π T 4 ( π 2 ) RT πz (π) Rz (π) Ry ( 2 ) ( ) · · ·XX π4 EPR RT πy ( 2) RTz (π) Rz (π) R πy · · ·2 Rz (π) R π y · · ·2 EPR · · · ( ) R (π) R πz y · · ·2 EPR · · · · · · EPR = Rx (π) · · · EPR Rx (π) ( ) · · · Rz (π) R π y · · ·2 EPR · · · Figure 5.21: Equivalence used to eliminate XX gates from stacked UCZ unitaries using the EPR property in Figure 5.19. 111 Chapter 6: Complete 3-Qubit Grover Search The Grover quantum search algorithm is a hallmark application of a quan- tum computer with a well-known speedup over classical searches of an unsorted database. Here, we report results for a complete three-qubit Grover search algo- rithm using the scalable quantum computing technology of trapped atomic ions, with better-than-classical performance. Two methods of state marking are used for the oracles: a phase-flip method employed by other experimental demonstrations, and a previously-undemonstrated Boolean method requiring an ancilla qubit that is directly equivalent to the state-marking scheme required to perform a classical search. The data presented here is also presented in [24]. 6.1 The Grover Search Algorithm Searching large databases is an important problem with broad applications. The Grover search algorithm [86,87] provides a powerful method for quantum com- puters to perform searches with a quadratic speedup in the number of required database queries over classical computers. It is an optimal search algorithm for a quantum computer [88], and has further applications as a subroutine for other 112 Initialize Oracle Amplification Figure 6.1: Evolution of relative amplitudes for each state during a Grover search algorithm. The initialization stage creates an equal superposition of all possible input states, so the amplitude αx = 1 for all basis states |x〉. The oracle stage marks the desired state, so the amplitude αm of the marked state |m〉 becomes negative while the amplitudes αb∑of the unmarked states |b〉, b 6= m remain∑unchanged. Theamplification stage performs a reflection about the mean vector N−1x=0 |x〉, which has amplitude A = 1 N−1 1x=0 αx = (−αm + (N − 1)αN N b), to amplify the marked state. An appropriate number of repetitions of the oracle and amplification stages will maximize the amplitude of the correct answer. All qubit states are normalized by the factor √1 . The algorithm can also be generalized to mark and amplify the N amplitude of t desired states. quantum algorithms [89, 90]. Searches with two qubits have been demonstrated on a variety of platforms [91–97] and proposed for others [98], but larger search spaces have only been demonstrated on a non-scalable NMR system [73]. The Grover search algorithm has 4 stages: initialization, oracle, amplifica- tion, and measurement, as shown in Fig∑ure 6.1. The initialization stage creates an equal superposition of all states, √1 N−1x=0 |x〉 for all basis states |x〉. The or-N acle stage ma(rks th∑e solution(∑s) |m〉 by fl)ipping the sign of that state’s amplitude, yielding √1 −αm m |m〉+ b=6 m αb|b〉 for marked state amplitude(s) αm andN non-marked state amplitudes αb. Inverting an amplitude α about an average am- plitude A can be written as −α + 2A or A + (A −∑α). Here, the amplification stage performs a reflection about the mean vector N−1x=0 |x〉, thus increasing the 113 amplitude of the marked state: [ ∑ ∑ ] |Ψamp〉 √ 1 = (2A+ αm) |m〉+ (2A− αb) |b〉 , (6.1) N m b 6=m where A is the amplitude of the mean vector. Finally, the algorithm output is measured. For a search database of size N with t possible solutions, the single- shot probability of measuring the correct answ√er is maximized to near-unity by repeating the oracle and amplification stages O( N/t) times before measurement. The probability of failing to measure a correct answer is guaranteed to be less than t N if the optimal number of iterations, j = b π c, is used, where sin2(θ) = t , 0 < θ ≤ π ; 4θ N 2 for t N , the probability of failure is negligible [86,87]. By comparison, a classical search algorithm will get the correct answer after an average of N/2 queries of the oracle, and in the worst case may require up to N queries to find the correct answer. For large databases, this quadratic speedup represents a significant advantage for quantum computers. Here, we implement the Grover search algorithm on n = 3 qubits, which corresponds to a search database of size N = 2n = 8. All searches are performed with a single iteration (j = 1). After a single iteration, the initial∑amplitudes αx = 1 on all basis states |x〉, so the mean vector has amplitude A = 1 N−1x=0 αx =N 1 (−αm · t+ (N − t)α N−2tb) = . Plugging in to Equation 6.1 above and rewritingN N 114 the inversion about the mean vector as A+ (A− α), we have [ ∑ ∑ ] | 1Ψamp(j = 1)〉 = √ [(A+ (A+ αm)) |m〉+ (A+ (A− αb)) |b〉N ( ( m )) b 6=m √1 N − 2t N − 2t ∑ = + + 1 |m]〉N ( N ( N )) m N − 2t N − 2t ∑ + [ + − 1 |b〉( N N ) b6=m ( ) ] √1 N − 2t 2(N − t ∑ ∑ = + | N − 2t 2tm〉+ − ( |b〉 . N N N N Nm b6=m (6.2) Since there are t possible correct solutions |m〉, the probability of measuring one correct solution after one iteration is ([ ] )2 · N − 2t 2(N − t) √1Pm = t + . (6.3) N N N In contrast, the optimal classical search strategy consists of a single query (equiva- lent to a single Grover iteration) followed by a random guess in the event the query failed. Therefore, the total probability of finding a correct solution is P (success) = P (query correct) + P (query incorrect) ∗ P (guess correct|query incorrect), where P (query correct) = t , P (query incorrect) = N−t , and P (guess correct|query incorrect) N N = t− . In general, for t solutions in the classical case,N 1 t N − t t P (success) = + · . (6.4) N N N − 1 115 Consequently, for a single-solution algorithm (t = 1),(t[he algorithmic]prob)abil-2 (ity of)measuring the correct state after one iteration is t · N−2t + 2(N−t) √1 =N N N2 √5 = 78.125% [87], compared to t + N−t · t 1 7 1− = + · = 25% for the optimal4 2 N N N 1 8 8 7 classical search strategy. In the two-solution case (t = 2), where two states are marked as correct answers during the oracle stage and both states’ amplitudes are amplified in the algorithm’s am(p[lification stage], the)probabi(lity o)f measur(ing)one of the two correct answers is2 2 2 t · N−2t + 2(N−t) √1 = 2 · 1√6 = 2 · √1 = 100% for the quantum case, N N N 8 8 2 as compared to 2 + 6 · 2 = 13 ≈ 46.4% for the classical case. 8 8 7 28 The algorithm is performed with both a phase oracle, which has been previ- ously demonstrated on other experimental systems, and a Boolean oracle, first re- ported here, which requires more resources but is directly comparable to a classical search. All quantum solutions are shown to outperform their classical counterparts. 6.2 Oracles We examine two alternative methods of encoding the marked state within the oracle. While both methods are mathematically equivalent [65], only one is directly comparable to a classical search. The previously-undemonstrated Boolean method requires the use of an ancilla qubit initialized to |1〉, as shown in Figure 6.2(a). The oracle is determined by constructing a circuit out of NOT and Ck(NOT ) (k ≤ n) gates such that, were the oracle circuit to be implemented classically, the ancilla bit would flip if and only if the input to the circuit is one of the marked states. By using 116 (a) Init Amplification |q1〉 : |0〉 H H X • X H |q2〉 : |0〉 H H X • X H | 〉 | 〉 Oracleq3 : 0 H H X Z X H |qa〉 : |1〉 H H (b) Init Oracle Amplification |0〉 H X • X H X • X H  ∑ |0〉 • •  √5 1H H X X H |011〉+ √4 2 4 2 x 6=011 |x〉 |0〉 H • H X Z X H |1〉 H H |1〉 (c) Init Amplification |q1〉 : |0〉 H H X • X H |q2〉 : |0〉 H Oracle H X • X H |q3〉 : |0〉 H H X Z X H (d) Init Oracle Amplification |0〉 H • H X • X H  |0〉 H • H X • 1X H √ (|011〉+ |101〉)2 |0〉 H Z Z H X Z X H Figure 6.2: (a) General circuit diagram for a Grover search algorithm using a Boolean oracle, depicted using standard quantum circuit diagram notation [65]. The last qubit qa is the ancilla qubit. (b) Example single-solution Boolean oracle marking the |011〉 state. (c) General circuit diagram for a Grover search algorithm using a phase oracle. (d) Example two-solution phase oracle marking the |011〉 and |101〉 states. classically available gates, this oracle formulation is directly equivalent to the classi- cal search algorithm, and therefore can most convincingly demonstrate the quantum algorithm’s superiority. On a quantum computer, because the initialization sets up an equal superposition of all possible input states, the Cn(NOT ) gate targeted on the ancilla provides a phase kickback that flips the phase of the marked state(s) in 117 the data qubits. An example oracle is shown in Figure 6.2(b) to illustrate this. The phase method of oracle implementation does not require the ancilla qubit. Instead, the oracle is implemented with a circuit consisting of Z and Ck(Z) (k ≤ n−1) gates that directly flip the phase(s) of the state(s) to be marked (see Figures 6.2(c-d)). 6.3 Circuit Implementation The Grover search algorithm is implemented using circuits that are equivalent to those shown in Figures 6.1(b,d), but with the initialization and amplification stages optimized to minimize gate times, as shown in Figures 6.3(a-b). The circuits shown are for use with Boolean oracles; in the phase oracle case, the ancilla qubit qa is simply omitted. To preserve the modularity of the algorithm, the initialization stage and amplification stage were each optimized without regard to the contents of the oracle, so each possible oracle can simply be inserted into the algorithm without making any changes to the other stages. Oracles for the Grover search algorithm were constructed using a combination of reversible and classical logic synthesis techniques. For Boolean oracles, reversible logic synthesis was employed to find a set of X,CN(NOT ) gates that marked the desired state(s) for each oracle. For phase oracles, EXOR polynomial synthesis was used to find a set of Z,CN(Z) gates that marked the desired state(s) for each oracle. For example, for Boolean oracles, the selection was limited to the classically available X (or NOT ) and CN(NOT ) gates, and a reversible circuit was constructed such that the output bit (corresponding to the ancilla qubit in the quantum oracle) would 118 (a) |q1〉 : |0〉 Rx(π) Ry(−π )2 |q2〉 : |0〉 Rx(π) Ry(−π )2 |q3〉 : |0〉 Rx(π) R πy(− )2 |qa〉 : |0〉 Ry(−π )2 (b) ( ) |q1〉 Ry ((1− β)π) Rx(β 3π ) XX(β π )2 4 8|q2〉 Ry (1− γ)π R (γ π ) 2π γ+12 x 2 R(− , ( )π − P )3 2 XX(γ π ) |q 83〉 Ry(π) Rx(π ) XX(β π )4 8 |qa〉 R (πy )2 Rx(π) · · · XX(απ ) 4 · · · R(−αβγ 2π , (αβ+1)π − αβγP ) R(π,−αβγ π ) 3 2 4 XX(γ π ) 8 · · · · · · ( ) · · · Ry (β − 1)π2 XX(απ ) 4 · · · Ry(−π )2 · · · Ry(π) · · · Figure 6.3: Grover search algorithm implementation by substage using XX(χ), Rx(θ), Ry(θ), and R(θ, φ) gates. The circuits shown are for use with Boolean ora- cles; removing the ancilla qubit |qa〉 produces the necessary circuits for use(w√ith)a phase oracle. Let α = sgn(χ12), β = sgn(χ1t), γ = sgn(χ2t), and P = arcsin 2 . 3 (a) Grover initialization stage implementation. (b) Grover amplification stage im- plementation. 119 be flipped if and only if a marked state was used as the input to the circuit. While there are many possible circuit constructions for each oracle, the oracle chosen for implementation was one that first minimized the number of two-qubit interactions required, and then minimized the number of single-qubit interactions needed. The synthesis techniques used are scalable and can be applied to oracles of any size. The oracles used here were implemented as per the circuit diagrams shown in Table 6.1 for single-solution oracles and Table 6.2 for two-solution oracles. The algorithm is executed for all 8 possible single-result oracles and all 28 possible two-result oracles. Mark Boolean Oracle Phase Oracle Mark Boolean Oracle Phase Oracle 000 |q 1001〉 X • X |q1〉 X • X |q1〉 • |q1〉 • |q2〉 X • X |q2〉 X • X |q2〉 X • X |q2〉 X • X |q3〉 X • X |q3〉 X Z X |q3〉 X • X |q3〉 X Z X |qa〉 |qa〉 001 |q1〉 X • X |q1〉 X • X 101 |q1〉 • |q1〉 • |q2〉 X • X |q2〉 X • X |q2〉 X • X |q2〉 X • X |q3〉 • |q3〉 Z |q3〉 • |q3〉 Z |qa〉 |qa〉 010 |q1〉 X • X |q1〉 X • X 110 |q1〉 • |q1〉 • |q2〉 • |q2〉 • |q2〉 • |q2〉 • |q3〉 X • X |q3〉 X Z X |q3〉 X • X |q3〉 X Z X |qa〉 |qa〉 011 |q1〉 X • X |q1〉 X • X 111 |q1〉 • |q1〉 • |q2〉 • |q2〉 • |q2〉 • |q2〉 • |q3〉 • |q3〉 Z |q3〉 • |q3〉 Z |qa〉 |qa〉 Table 6.1: Table of all oracles used for the single-solution Grover search algorithm. 120 Marked Boolean Oracle Phase Oracle 000, 001 |q1〉 X • X |q1〉 Z • |q2〉 X • X |q2〉 Z Z |q3〉 |q3〉 |qa〉 000, 010 |q1〉 X • X |q1〉 Z • |q2〉 |q2〉 |q3〉 X • X |q3〉 Z Z |qa〉 000, 011 |q1〉 X • X |q1〉 Z • • |q2〉 X • • X |q2〉 Z Z |q3〉 • |q3〉 Z Z |qa〉 000, 100 |q1〉 |q1〉 |q2〉 X • X |q2〉 Z • |q3〉 X • X |q3〉 Z Z |qa〉 000, 101 |q1〉 X • • X |q1〉 Z • |q2〉 X • X |q2〉 Z Z • |q3〉 • |q3〉 Z Z |qa〉 000, 110 |q1〉 X • • X |q1〉 Z • |q2〉 • |q2〉 Z • |q3〉 X • X |q3〉 Z Z Z |qa〉 000, 111 |q1〉 X • • • • X |q1〉 Z • • |q2〉 • |q2〉 Z Z • |q3〉 • |q3〉 Z Z Z |qa〉 001, 010 |q1〉 X • X |q1〉 • • |q2〉 • • |q2〉 Z Z |q3〉 • |q3〉 Z Z |qa〉 Table 6.2 – Continued on next page 121 Table 6.2 – Continued from previous page Marked Boolean Oracle Phase Oracle 001, 011 |q1〉 X • X |q1〉 • |q2〉 |q2〉 |q3〉 • |q3〉 Z Z |qa〉 001, 100 |q1〉 • • |q1〉 Z • |q2〉 X • X |q2〉 Z • |q3〉 • |q3〉 Z Z |qa〉 001, 101 |q1〉 |q1〉 |q2〉 X • X |q2〉 • |q3〉 • |q3〉 Z Z |qa〉 001, 110 |q1〉 • • • • |q1〉 • • |q2〉 X • X |q2〉 Z • |q3〉 • |q3〉 Z Z Z |qa〉 001, 111 |q1〉 X • • X |q1〉 • |q2〉 • |q2〉 • |q3〉 • |q3〉 Z Z Z |qa〉 010, 011 |q1〉 X • X |q1〉 • |q2〉 • |q2〉 Z Z |q3〉 |q3〉 |qa〉 010, 100 |q1〉 • • |q1〉 Z • |q2〉 • |q2〉 Z • |q3〉 X • X |q3〉 Z Z |qa〉 010, 101 |q1〉 • • • • |q1〉 • • |q2〉 • |q2〉 Z Z • |q3〉 X • X |q3〉 Z Z |qa〉 Table 6.2 – Continued on next page 122 Table 6.2 – Continued from previous page Marked Boolean Oracle Phase Oracle 010, 110 |q1〉 |q1〉 |q2〉 • |q2〉 Z • |q3〉 X • X |q3〉 Z |qa〉 010, 111 |q1〉 X • • X |q1〉 • |q2〉 • |q2〉 Z Z • |q3〉 • |q3〉 Z |qa〉 011, 100 |q1〉 • • • • |q1〉 Z • • |q2〉 • |q2〉 Z • |q3〉 • |q3〉 Z Z |qa〉 011, 101 |q1〉 • • |q1〉 • |q2〉 • |q2〉 • |q3〉 • |q3〉 Z Z |qa〉 011, 110 |q1〉 • • |q1〉 • |q2〉 • |q2〉 Z • |q3〉 • |q3〉 Z |qa〉 011, 111 |q1〉 |q1〉 |q2〉 • |q2〉 • |q3〉 • |q3〉 Z |qa〉 100, 101 |q1〉 • |q1〉 Z • |q2〉 X • X |q2〉 Z |q3〉 |q3〉 |qa〉 100, 110 |q1〉 • |q1〉 Z • |q2〉 |q2〉 |q3〉 X • X |q3〉 Z |qa〉 Table 6.2 – Continued on next page 123 Table 6.2 – Continued from previous page Marked Boolean Oracle Phase Oracle 100, 111 |q1〉 • |q1〉 Z • • |q2〉 X • • X |q2〉 Z |q3〉 • |q3〉 Z |qa〉 101, 110 |q1〉 • |q1〉 • • |q2〉 • • |q2〉 Z |q3〉 • |q3〉 Z |qa〉 101, 111 |q1〉 • |q1〉 • |q2〉 |q2〉 |q3〉 • |q3〉 Z |qa〉 110, 111 |q1〉 • |q1〉 • |q2〉 • |q2〉 Z |q3〉 |q3〉 |qa〉 Table 6.2: Table of all oracles used for the two-solution Grover search algorithm. Detailed information and individual fidelities for constituent composite gates can be found in Chapter 5. 6.4 Data Figures 6.4 and 6.5 show the results, respectively, of single- and two-solution Grover search algorithms, each using both the Boolean and phase marking methods. All possible oracles are tested to demonstrate a complete Grover search (see Tables 6.1 and 6.2). Two figures of merit are provided with the data for each oracle. The 124 Boolean Oracles ASP(%) SSO(%) Phase Oracles ASP(%) SSO(%) 50 000 34(1) 80(2) 34.1(7) 79(1) 001 38(1) 83(2) 50.2(7) 89(1) 40 010 40(1) 85(2) 33.3(7) 78(1) 011 41(1) 86(2) 46.4(7) 87(1) 30 100 31(1) 76(2) 42.7(7) 86(1) 20 101 35(1) 81(2) 55.1(7) 91(1) 110 41(1) 85(2) 41.4(7) 85(1) 10 111 50(1) 90(2) 46.5(7) 84(1) 0 000 010 100 110 000 010 100 110 detected state detected state Figure 6.4: Results from a single iteration of a single-solution Grover search algo- rithm performed on a 3-qubit database. Data for the Boolean oracle formulation is shown on the left, and data for the phase oracle formulation is shown on the right. The plots show the probability of detecting each output state. All values shown are percents, with a theoretical ASP of 78.1% and theoretical SSO of 100%. Data is corrected for average SPAM errors of 1%. algorithm success probability (ASP) is the probability of measuring the marked state as the experimental outcome. For the two-solution algorithm, the ASP is calculated by summing the probabilities of measuring each of the two marked states. The squared statistical overlap (SSO) measures the statis(t∑ical overlap b)etween the√ 2 measured and expected populations for all states: SSO = Nj=0 ejmj , where ej is the expected population and mj is the measured population for each state j [99]. Additionally, all of the data shown in this paper is corrected to account for state preparation and measurement (SPAM) errors (see figure captions for values), similar to the method proposed in [39] while also accounting for multi-ion crosstalk [19]. All uncertainties given are statistical uncertainties based on the number of experiments performed. The single-iteration, single-solution Grover search algorithm shown in Figure 6.4 has a theoretical ASP of 78.1%, as discussed above. The SSO takes into account 125 marked state Boolean Oracles ASP(%) SSO(%) Phase Oracles ASP(%) SSO(%) 50 000,001 67(1) 67(1) 77(1) 77(1) 000,010 66(1) 66(1) 79(1) 79(1) 000,011 66(1) 66(1) 77(1) 75(1) 000,100 68(1) 68(1) 74.7(9) 72(1) 000,101 65(1) 65(1) 79(1) 79(1) 000,110 62(1) 62(1) 69(1) 69(1) 40 000,111 66(1) 66(1) 74(1) 73(1) 001,010 64(1) 64(1) 79(1) 79(1) 001,011 71(1) 71(1) 72(1) 71(1) 001,100 68(1) 67(1) 79(1) 78(1) 001,101 69(1) 69(1) 80(1) 79(1) 001,110 60(1) 60(1) 82(1) 81(1) 30 001,111 65(1) 65(1) 73(1) 72(1) 010,011 75(1) 75(1) 73(1) 73(1) 010,100 66(1) 65(1) 70(1) 70(1) 010,101 65(1) 64(1) 73.8(9) 69.7(9) 010,110 72(1) 72(1) 74(1) 73(1) 010,111 73(1) 72(1) 75(1) 74(1) 20 011,100 65(1) 65(1) 74(1) 74(1) 011,101 66(1) 66(1) 74(1) 74(1) 011,110 73(1) 73(1) 75(1) 75(1) 011,111 74(1) 74(1) 80(1) 78(1) 100,101 74(1) 74(1) 72(1) 71(1) 10 100,110 67(1) 66(1) 66.7(9) 67(1) 100,111 64(1) 64(1) 75.5(9) 71(1) 101,110 65(1) 65(1) 72(1) 71(1) 101,111 70(1) 69(1) 77(1) 77(1) 110,111 74(1) 74(1) 80(1) 80(1) 0 000 010 100 110 000 010 100 110 detected state detected state Figure 6.5: Results from the execution of a two-solution Grover search algorithm performed on a 3-qubit database. Data for the Boolean oracle formulation is shown on the left, and data for the phase oracle formulation is shown on the right. The plots show the probability of detecting each output state. All values shown are percents. The ASP is the sum of the probabilities of detecting each of the two marked states. Data is corrected for average SPAM errors of 1%. 126 marked state that the 7 unmarked states then have equal expected probabilities totaling 21.9% of being measured. For all Boolean oracles, the average ASP is 38.9(4)% and the average SSO is 83.2(7)%, while phase oracles have an average ASP of 43.7(2)% and an average SSO of 84.9(4)%; the reduced use of resources in the phase oracles (10 XX(χ) gates and 3 qubits for phase oracles compared to 16 XX(χ) gates and 5 qubits for Boolean oracles) results in better performance, as expected. These results compare favorably with the classical ASP of 25%. The two-solution Grover search algorithm shown in Figure 6.5 has a theoretical ASP of 100%, as discussed above. For all Boolean oracles, the average ASP is 67.9(2)% and the average SSO is 67.6(2)%, while phase oracles have an average ASP of 75.3(2)% and an average SSO of 74.4(2)%; the reduced use of resources in the phase oracles (6-8 XX(χ) gates and 3 qubits for phase oracles compared to 10- 14 XX(χ) gates and 4 qubits for Boolean oracles) results in better performance, as expected. For all oracles in both cases, the two states with the highest measurement probability are also the two marked states. These results compare favorably with the classical ASP of 46.4%. 6.5 Additional Iterations Performing an additional iteration on the single-solution Grover search algo- rithms was inhibited by circuit-depth limitations in the experimental control pro- gram, which will be fixed for future work. Here, we estimate the impact of an additional iteration on algorithm performance. While a single iteration of the single- 127 solution Grover search algorithm has a maximum ASP of 78.125%, performing two iterations raises the maximum ASP to 94.5312%. Applying the error estimation models used in [27], the likely performances of two Grover search algorithm cases were examined: the Boolean 000 oracle and the phase 111 oracle, which correspond to the worst- and best-case oracles by gate count. The random error estimation √ model assumes random error propagation for each operation of the form (1−  ) Ng , and the systematic error estimation model assumes coherent over- or under-rotations for each operation and has the form (1− g)N , where N is the number of operations and g is the error per operation. Based on the analysis in [27] on this same system, we expect the actual results to fall somewhere between these two models. For a single iteration of the phase 111 oracle, we estimate an SSO of 86% and an ASP of 41% using the random error model, or an SSO of 61% and an ASP of 16% using the systematic error model; as in the analysis in [27], we compare this to the measured SSO of 84(1)% and ASP of 46.5(7)% and see that the experiment performs slightly worse than the random error model, and better than the systematic error model. Extending the analysis to two iterations of the phase 111 oracle, we estimate an SSO of 81% and ASP of 60% using the random error model, or an SSO of 40% and an ASP of 19% using the systematic error model. Similarly, for a single iteration of the Boolean 000 oracle, we estimate an SSO of 83% and an ASP of 37% using the random error model, or an SSO of 45% and an ASP of 6% using the systematic error model; as before, we compare this to the measured SSO of 80(2)% and ASP of 34(1)% and see that the experiment performs slightly worse than the random error model. Extending the analysis to two 128 iterations of the phase 111 oracle, we estimate an SSO of 77% and ASP of 55% using the random error model, or an SSO of 22% and an ASP of 6% using the systematic error model. We expect the experiment would perform somewhere between these two models, although we do not know how well these error models hold for very deep circuits; it is not clear whether the experiment would have outperformed the best classical strategy with two iterations, which has a success probability of 37.5%. 6.6 Outlook We note that this implementation of the Grover search algorithm scales linearly in the two-qubit gate count and ancilla count for increasing search database size as a function of the number of qubits n, and for a constant number of solutions t. For a database of size N = 2n stored on n qubits, the amplification stage requires one Toffoli-n gate, and the t-solution oracle stage requires at worst t Toffoli-n (for a phase oracle) or Toffoli-(n+ 1) (for a Boolean oracle) gates; optimal oracles for particular sets of marked states may require even fewer two-qubit gates. The method used in Section 5.4.2 to construct the Toffoli-4 circuit scales to Toffoli-n gates as 6n− 13 in the two-qubit gate count and as dn−3e in the ancilla count [77]. This paves the way 2 for more extensive use of the Grover search algorithm in solving larger problems on quantum computers, including using the circuit as a subroutine for other quantum algorithms. 129 Chapter 7: Parallel 2-Qubit Operations In any computer, quantum or classical, parallel operations are highly desir- able. Parallel operations save considerable time over performing operations in series. Current trends in the (classical) computer industry include a focus on developing multi-core processors and developing software with multiple parallel threads. Quan- tum computers have a long way to go before such a scale will be possible, but the ability to perform parallel operations is nevertheless crucial to our ultimate ability to scale up. In ion-trap quantum computers, two-qubit interactions are mediated by the normal modes of motion in the ion chain. However, as the chain grows in size, so do the number of modes of motion, and spectral crowding makes sideband resolution more difficult. Two-qubit interactions can be implemented in less time by using more optical power in the Raman beams, but this has the consequence of reducing sideband resolution, which degrades gate fidelity. Consequently, gate speed is lim- ited by sideband resolution, a limitation that gets worse as the processor size grows. Parallel two-qubit operations are a tool to speed up computation that avoids this problem. Parallel gate operations also reduce the overall gate depth of a given pro- cess, permitting more operations before decoherence and error accumulation obviate 130 any meaningful outputs. Several quantum computing subroutines and composite gates have been shown to benefit directly from parallel entangling gates, including the quantum Fourier transform [100,101], multiply controlled Toffoli gates, and sta- bilizer circuits [101]. Quantum algorithms that will similarly benefit from parallel entangling gates include Shor’s integer factoring [102], solving the discrete logarithm problem over the elliptic curve group [103], simulating Hamiltonian dynamics using the Suzuki-Trotter formula [104], and quantum chemistry [105]. Here, we present experimental results for a pair of two-qubit gates performed simultaneously in a single chain of trapped ions. We employ a pulse shaping scheme that modulates the phase and amplitude of the Raman laser to drive programmable high-fidelity 2-qubit XX gates in parallel by coupling to the collective modes of motion of the ion chain. Ensuring the interaction produced yields only spin-spin in- teractions between the desired pairs with neither residual spin-motion entanglement nor crosstalk spin-spin entanglement between non-desired ion pairs is a nonlinear constraint problem, and optimal pulse shapes are found using optimization tech- niques. As an application of these parallel operations, we demonstrate the quantum full adder using a depth-4 circuit requiring the use of parallel 2-qubit operations as well as single- and two-qubit operations previously demonstrated on this system. 7.1 Theory We perform parallel gates by modulating the Rabi frequency of the individ- ual Raman beams. This is accomplished in a similar manner to the 2-qubit gates 131 already implemented on this experiment [19,53] (see Section 3.2), which implement entangling XX gates using a Mølmer-Sørensen scheme [41,52,55,56], where red and blue sidebands are applied to entangle the ion spins of illuminated ions with the radial normal motional modes of the ions in the trap. Using the motional modes as an information bus, the spin states of separate ions become entangled, and the pulse shape is engineered so that at the end of the gate, the motional modes are entirely disentangled from the spins, leaving only spin-spin entanglement [53,57,59]. The pulse shape is controlled by dividing the gate time into several equal-length segments, and adjusting the amplitude of the gate modulation in each segment. Alternative possible schemes include varying the detuning µ [106] or the beatnote phase [107] to engineer high-fidelity gates; the former scheme has been demonstrated to generate 2-qubit entangling XX gates on this system [28]. In order to perform parallel entangling operations involving 2 pairs of qubits (i, j) and (m,n) in a chain of N ions with N motional modes ωk using red and blue sidebands with detuning µ, we have a 4-qubit unitary, as follows: ( ∑ ∑ ) U x x x||(τ) = exp i φi(τ)σi + i χij(τ)σi σj ( [ i i: The language integrated quantum operations simulator. http://stationq.github.io/Liquid/. [71] Rigetti Computing. Forest. https://www.rigetti.com/forest. [72] Tommaso Toffoli. Reversible computing. In Automata, Languages and Pro- gramming, Lecture Notes in Computer Science, pages 632–644. Springer, Berlin, Heidelberg, 1980. [73] Lieven M. K. Vandersypen, Matthias Steffen, Mark H. Sherwood, Costantino S. Yannoni, Gregory Breyta, and Isaac L. Chuang. Implementation of a three-quantum-bit search algorithm. Applied Physics Letters, 76(5):646– 648, 2000. [74] Dmitri Maslov. Basic circuit compilation techniques for an ion-trap quantum machine. New Journal of Physics, 19(2):023035, 2017. 190 [75] David G. Cory, Mark D. Price, and Timothy F. Havel. Nuclear magnetic resonance spectroscopy: An experimentally accessible paradigm for quantum computing. Physica D, 120(1):82–101, 1998. [76] T. Monz, K. Kim, W. Hänsel, M. Riebe, A. S. Villar, P. Schindler, M. Chwalla, M. Hennrich, and R. Blatt. Realization of the quantum Toffoli gate with trapped ions. Physical Review Letters, 102(4):040501, 2009. [77] Dmitri Maslov. Advantages of using relative-phase Toffoli gates with an application to multiple control Toffoli optimization. Physical Review A, 93(2):022311, 2016. [78] Edward Fredkin and Tommaso Toffoli. Conservative logic. International Jour- nal of Theoretical Physics, 21(3):219–253, 1982. [79] Jiangfeng Du, Ping Zou, Xinhua Peng, Daniel K. L. Oi, L. C. Kwek, C. H. Oh, and Artur Ekert. Experimental quantum multimeter and one-qubit fin- gerprinting. Physical Review A, 74(4):042319, 2006. [80] Raj B. Patel, Joseph Ho, Franck Ferreyrol, Timothy C. Ralph, and Geoff J. Pryde. A quantum Fredkin gate. Science Advances, 2(3):e1501531, 2016. [81] Takafumi Ono, Ryo Okamoto, Masato Tanida, Holger F. Hofmann, and Shigeki Takeuchi. Implementation of a quantum controlled-SWAP gate with photonic circuits. Scientific Reports, 7:45353, 2017. [82] D. Maslov, G. W. Dueck, D. M. Miller, and C. Negrevergne. Quantum circuit simplification and level compaction. IEEE Transactions on Computer-Aided Design of Integrated Circuits and Systems, 27(3):436–444, 2008. [83] Patrick Hayden and John Preskill. Black holes as mirrors: quantum informa- tion in random subsystems. Journal of High Energy Physics, 2007(9):120–120, 2007. [84] Beni Yoshida and Norman Y. Yao. Disentangling scrambling and decoherence via quantum teleportation. forthcoming, 2018. [85] Beni Yoshida and Alexei Kitaev. Efficient decoding for the Hayden-Preskill protocol. arXiv:1710.03363, 2017. [86] Lov K. Grover. Quantum Mechanics Helps in Searching for a Needle in a Haystack. Physical Review Letters, 79(2):325–328, 1997. [87] Michel Boyer, Gilles Brassard, Peter Høyer, and Alain Tapp. Tight bounds on quantum searching. Fortschritte der Physik, 46(4):493–505, 1998. [88] C. Bennett, E. Bernstein, G. Brassard, and U. Vazirani. Strengths and weak- nesses of quantum computing. SIAM Journal on Computing, 26(5):1510–1523, 1997. 191 [89] F. Magniez, M. Santha, and M. Szegedy. Quantum algorithms for the triangle problem. SIAM Journal on Computing, 37(2):413–424, 2007. [90] C. Dürr, M. Heiligman, P. Høyer, and M. Mhalla. Quantum query complexity of some graph problems. SIAM Journal on Computing, 35(6):1310–1328, 2006. [91] Isaac L. Chuang, Neil Gershenfeld, and Mark Kubinec. Experimental Imple- mentation of Fast Quantum Searching. Physical Review Letters, 80(15):3408– 3411, 1998. [92] N. Bhattacharya, H. B. van Linden van den Heuvell, and R. J. C. Spreeuw. Implementation of Quantum Search Algorithm using Classical Fourier Optics. Physical Review Letters, 88(13):137901, 2002. [93] K.-A. Brickman, P. C. Haljan, P. J. Lee, M. Acton, L. Deslauriers, and C. Mon- roe. Implementation of Grovers quantum search algorithm in a scalable sys- tem. Physical Review A, 72(5):050306(R), 2005. [94] P. Walther, K. J. Resch, T. Rudolph, E. Schenck, H. Weinfurter, V. Vedral, M. Aspelmeyer, and A. Zeilinger. Experimental one-way quantum computing. Nature, 434(7030):169–176, 2005. [95] L. DiCarlo, J. M. Chow, J. M. Gambetta, Lev S. Bishop, B. R. Johnson, D. I. Schuster, J. Majer, A. Blais, L. Frunzio, S. M. Girvin, and R. J. Schoelkopf. Demonstration of two-qubit algorithms with a superconducting quantum pro- cessor. Nature, 460(7252):240–244, July 2009. [96] Stefanie Barz, Elham Kashefi, Anne Broadbent, Joseph F. Fitzsimons, Anton Zeilinger, and Philip Walther. Experimental Demonstration of Blind Quantum Computing. Science, 335(6066):303–308, 2012. [97] T.A. Manning. Quantum information processing with trapped ion chains. PhD thesis, University of Maryland, 2014. [98] Klaus Mølmer, Larry Isenhower, and Mark Saffman. Efficient Grover search with Rydberg blockade. Journal of Physics B: Atomic, Molecular and Optical Physics, 44(18):184016, 2011. [99] J. Chiaverini, J. Britton, D. Leibfried, E. Knill, M. D. Barrett, R. B. Blakestad, W. M. Itano, J. D. Jost, C. Langer, R. Ozeri, T. Schaetz, and D. J. Wineland. Implementation of the Semiclassical Quantum Fourier Transform in a Scalable System. Science, 308(5724):997–1000, 2005. [100] R. Cleve and J. Watrous. Fast parallel circuits for the quantum fourier trans- form. In Proceedings 41st Annual Symposium on Foundations of Computer Science, pages 526–536, 2000. 192 [101] D. Maslov. Linear depth stabilizer and quantum fourier transformation circuits with no auxiliary qubits in finite neighbor quantum architectures. Physical Review A, 76(5):052310, 2007. [102] Austin G. Fowler, Simon J. Devitt, and Lloyd C. L. Hollenberg. Implementa- tion of Shor’s algorithm on a linear nearest neighbour qubit array. Quantum Information and Computation, 4:237–251, 2004. [103] Donny Cheung, Dmitri Maslov, Jimson Mathew, and Dhiraj K. Pradhan. On the design and optimization of a quantum polynomial-time attack on ellip- tic curve cryptography. In Theory of Quantum Computation, Communica- tion, and Cryptography, Lecture Notes in Computer Science, pages 96–104. Springer, Berlin, Heidelberg, 2008. [104] Tiago Januario and Sebastian Urrutia. An edge coloring heuristic based on vizing’s theorem. In Proceedings of the Brazilian Symposium on Operations Research, pages 3994–4002, 2012. [105] Ian D. Kivlichan, Jarrod McClean, Nathan Wiebe, Craig Gidney, Aln Aspuru- Guzik, Garnet Kin-Lic Chan, and Ryan Babbush. Quantum simulation of elec- tronic structure with linear depth and connectivity. Physical Review Letters, 120(11):110501, 2018. [106] S. Korenblit, D. Kafri, W. C. Campbell, R. Islam, E. E. Edwards, Z.-X. Gong, G.-D. Lin, L.-M. Duan, J. Kim, K. Kim, and C. Monroe. Quantum simulation of spin models on an arbitrary lattice with trapped ions. New Journal of Physics, 14(9):095024, 2012. [107] Todd J. Green and Michael J. Biercuk. Phase-modulated decoupling and error suppression in qubit-oscillator systems. Physical Review Letters, 114(12):120502, 2015. [108] Shengtao Wang. Applications of Atomic Systems in Quantum Simulation, Quantum Computation and Topological Phases of Matter. PhD thesis, Uni- versity of Michigan, 2017. [109] Chingiz Kabytayev, Todd J. Green, Kaveh Khodjasteh, Michael J. Biercuk, Lorenza Viola, and Kenneth R. Brown. Robustness of composite pulses to time-dependent control noise. Physical Review A, 90(1):012316, 2014. [110] Kenneth R. Brown, Aram W. Harrow, and Isaac L. Chuang. Arbitrarily accurate composite pulse sequences. Physical Review A, 70(5):052318, 2004. [111] C. A. Sackett, D. Kielpinski, B. E. King, C. Langer, V. Meyer, C. J. Myatt, M. Rowe, Q. A. Turchette, W. M. Itano, D. J. Wineland, and C. Monroe. Experimental entanglement of four particles. Nature, 404(6775):256–259, 2000. [112] Richard P. Feynman. Quantum mechanical computers. Optics News, 11(2):11– 20, 1985. 193 [113] Daniel M. Greenberger, Michael A. Horne, and Anton Zeilinger. Going beyond bells theorem. In Bells Theorem, Quantum Theory and Conceptions of the Universe, Fundamental Theories of Physics, pages 69–72. Springer, Dordrecht, 1989. [114] Ketan N. Patel, Igor L. Markov, and John P. Hayes. Optimal synthesis of linear reversible circuits. Quantum Information and Computation, 8(3):0282– 0294, 2008. [115] Dmitri Maslov and Yunseong Nam. Use of global interactions in efficient quantum circuit constructions. New Journal of Physics, 20(3):033018, 2018. [116] Mark Hillery, Vladimr Buek, and Andr Berthiaume. Quantum secret sharing. Physical Review A, 59(3):1829–1834, 1999. [117] G.-D. Lin, S.-L. Zhu, R. Islam, K. Kim, M.-S. Chang, S. Korenblit, C. Monroe, and L.-M. Duan. Large-scale quantum computation in an anharmonic linear ion trap. Europhysics Letters, 86(6):60004, 2009. [118] C.J. Ballance, T.P. Harty, N.M. Linke, M.A. Sepiol, and D.M. Lucas. High- Fidelity Quantum Logic Gates Using Trapped-Ion Hyperfine Qubits. Physical Review Letters, 117(6):060504, 2016. [119] J.P. Gaebler, T.R. Tan, Y. Lin, Y. Wan, R. Bowler, A.C. Keith, S. Glancy, K. Coakley, E. Knill, D. Leibfried, and D.J. Wineland. High-Fidelity Universal Gate Set for 9Be+ Ion Qubits. Physical Review Letters, 117(6):060505, 2016. [120] Ye Wang, Mark Um, Junhua Zhang, Shuoming An, Ming Lyu, Jing-Ning Zhang, L.-M. Duan, Dahyun Yum, and Kihwan Kim. Single-qubit quantum memory exceeding ten-minute coherence time. Nature Photonics, 11(10):646– 650, 2017. [121] D. Kielpinski, C. Monroe, and D. J. Wineland. Architecture for a large-scale ion-trap quantum computer. Nature, 417(6890):709–711, 2002. [122] I.V. Inlek, C. Crocker, M. Lichtman, K. Sosnova, and C. Monroe. Multi- species trapped-ion node for quantum networking. Physical Review Letters, 118(25):250502, 2017. 194