ABSTRACT Title of dissertation: QUANTUM SIMULATIONS OF THE ISING MODEL WITH TRAPPED IONS: DEVIL?S STAIRCASE AND ARBITRARY LATTICE PROPOSAL Simcha Korenblit, Doctor of Philosophy 2013 Dissertation directed by: Professor Christopher Monroe Joint Quantum Institute University of Maryland Department of Physics and National Institute of Standards and Technology A collection of trapped atomic ions represents one of the most attractive plat- forms for the quantum simulation of interacting spin networks and quantum mag- netism. Spin-dependent optical dipole forces applied to an ion crystal create long- range e ective spin-spin interactions and allow the simulation of spin Hamiltonians that possess nontrivial phases and dynamics. We trap linear chains of 171Yb+ ions [1] in a Paul trap [2], and constrain the occupation of energy levels to the ground hyper ne clock-states, creating a qubit or pseudo-spin 1/2 system. We proceed to implement spin-spin couplings between two ions using the far detuned M lmer-S renson scheme[3] and perform adiabatic quantum simulations of Ising Hamiltonians with long-range couplings. We then demonstrate our ability to control the sign and relative strength of the interaction between three ions. Using this control, we simulate a frustrated triangular lattice, and for the rst time establish an experimental connection between frustration and quantum entanglement. We then scale up our simulation to show phase transitions from paramagnetism to ferromagnetism for nine ions, and to anti-ferromagnetism for sixteen ions. The experimental work culminates with our most complicated Hamiltonian - a long range anti-ferromagnetic Ising interaction between 10 ions with a biasing axial eld, which I have led. Theoretical work presented in this thesis shows how the approach to quan- tum simulation utilized in this thesis can be further extended and improved. It is shown how appropriate design of laser elds can provide for arbitrary multidimen- sional spin-spin interaction graphs even for the case of a linear spatial array of ions. This scheme uses currently existing trap technology and is scalable to levels where classical methods of simulation are intractable. QUANTUM SIMULATIONS OF THE ISING MODEL WITH TRAPPED IONS: DEVIL?S STAIRCASE AND ARBITRARY LATTICE PROPOSAL by SIMCHA KORENBLIT Dissertation submitted to the Faculty of the Graduate School of the University of Maryland, College Park in partial ful llment of the requirements for the degree of Doctor of Philosophy 2013 Advisory Committee: Professor Christopher Monroe, Chair/Advisor Professor Edo Waks Professor Ted Jacobson Professor Eite Tiesinga Professor Ian Spielman Professor Gretchen Campbell c Copyright by Simcha Korenblit 2013 Dedication I dedicate this work to my family, who are my rock. ? ?????????????????????,???????????????,???????????????????? . ????????????,?????????????? How?manifold?are?Thy?works,?O?LORD!?In?wisdom?hast?Thou?made?them all?????the?earth?is?full?of?Thy?creatures. Psalms?104:24. ii Acknowledgments A doctoral program is rarely pursued in a vacuum. There are many people and organizations without whom this work would not be possible. Firstly, I would like to thank my advisor, Chris Monroe, for giving me the opportunity to work at the cutting edge of quantum information. None of this would have happened if he did not happen to be in his o ce the day I knocked on his door, uninvited, displaying way more con dence than warranted for someone who has not yet been accepted to the Michigan Physics Department. Chris is probably one of the most hands on advisors out there, the kind of advisor who helps his group put up optic table \clouds" and electronics room shelving, while not micro-managing group members, and allowing us room to grow as independent scientists. Throughout the years, I have bene ted greatly from the advice and discussions of scientists outside our group. Here I will only mention a few, and I ask forgiveness in advance from whoever I may have forgotten. Luming Duan has collaborated with us intensively over the years, and has guided us on many issues and directions worth pursuing. His students Guin-Dar Lin, Zeh-Xuan Gong and Chao Shen have worked closely with us and supported our work. We have also had many useful discussions and collaboration with Jim Freericks and his post-doc Joseph Wang, especially regarding the e ect of the transverse eld on ion chain phonons. We have also received numerical support from Howard Carmichael and his student Changsuk Noh on the extremely intensive quantum trajectory simulations of our experiment. Working in a world class lab means working with world class people. I have iii had the privilege to work with extremely talented and dedicated post doctoral re- searchers. When I arrived in Maryland I knew nothing about ion trapping and had no knowledge in atomic physics. I received a patient and dedicated introduction from Ming-Schien Chang and Kihwan Kim, who were friends and teachers to me. Later I had the privilege of working with the post-docs Emily Edwards and Wes Campbell. From them I have learned much as well. Soon after arriving I had a an amazing graduate student join me - Rajibul Islam, and he shared the trenches with me for most of my time at Maryland. Later other students have joined our project and contributed as well - Andrew Chew, an undergrad set up the iodine lock. Aaron Lee joined us as an undergrad and signi cantly improved my messy control program. He enjoyed this work so much he later joined us a grad student. Crystal Senko switched over from the fast gate project and was instrumental in trouble shooting our messy MBR problem. Jake Smith has joined as well and is quickly getting up to speed. Phil Richerme has joined the team as a post-doc and quickly began to contribute. I have also bene ted from working and playing with many grad students, undergrads and post-docs not directly on the project. I list them here in no particular order. Peter Maunz, Dzimitry Matsukevich, Susan Clark, Dan Stick, Yisa Ramula, David Moehring, Kathy-Ann Brickman, Steve Olmschenk, Qudsia Quraishi, Jon Sterk, Mark Acton, Le Luo, Charles Conover, Brian Fields, Kenny Lee, Ilka Geisel, Jonathan Mizrahi, Brian Neyenhuis, Andrew Manning, Geo rey Ji, Kale Johnson, Volkan Inlek, David Hucul, Dave Hayes, Ken Wright, Shantanu Debnath, Caroline Figgatt, Taeyoung Choi, Chenglin Cao, Daniel Brennan, and Volkan Inlek. It has been a pleasure and iv a terri c learning experience working with you. I would like to thank Crystal, Phil and Jake for reviewing my thesis. I thank my thesis committee for taking upon themselves this service to the department and to greater science community. I would also like to thank the Physics department sta , particularly Lorraine Desalvo and Jane Hessing, who have been very warm and welcoming to me when I made the transition from Michigan. A man cannot live on professional relationships alone. I would like to thank the close friends who were there for me during my time here, but they are too numerous to mention. I would like to send a special thanks to Rob Feldmier who has been like a brother to me, and Rabbi and Rebbetzin Teitelbaum, who have been amazing to me. I would like to thank MesorahDC and the entire Jewish community of DC, who have guided the spiritual journey that paralleled my intellectual journey in this program. This work is supported by the US Army Research O ce (ARO) with funds from the DARPA Optical Lattice Emulator (OLE) Program and the IARPA MQCO Program, the NSF Physics at the Information Frontier Program, the NSF Physics Frontier Center at JQI. Thank all of you and please forgive me if I omitted you. v Table of Contents List of Figures ix List of Abbreviations xi 1 Introduction 1 1.0.1 Classical Computation . . . . . . . . . . . . . . . . . . . . . . 2 1.0.2 Quantum Information Processing . . . . . . . . . . . . . . . . 3 1.1 The DiVincenzo Criteria . . . . . . . . . . . . . . . . . . . . . . . . . 5 1.2 Quantum Mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . . 6 1.3 Adiabatic quantum simulation . . . . . . . . . . . . . . . . . . . . . . 8 1.3.1 Hydrogen-Like Atom . . . . . . . . . . . . . . . . . . . . . . . 9 1.3.2 Transition Rules . . . . . . . . . . . . . . . . . . . . . . . . . 14 1.3.3 Interaction Hamiltonian . . . . . . . . . . . . . . . . . . . . . 14 1.3.4 E ective Hamiltonian . . . . . . . . . . . . . . . . . . . . . . . 15 1.4 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16 2 The 171Yb+ ion 17 2.1 Ionization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19 2.2 Dopppler Cooling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20 2.3 Optical Pumping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22 2.4 Detection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22 3 Quantum Control of The Ion 28 3.1 Light-Ion Interaction . . . . . . . . . . . . . . . . . . . . . . . . . . . 28 3.2 Raman Transitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29 3.3 Coupling Spin to Motion . . . . . . . . . . . . . . . . . . . . . . . . . 31 3.4 Multi-Ion Gates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35 3.5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38 4 Experimental Setup 40 4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40 4.2 Ion trap . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40 4.2.1 Normal Modes . . . . . . . . . . . . . . . . . . . . . . . . . . 43 4.3 Micromotion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49 4.3.1 The Linear Trap . . . . . . . . . . . . . . . . . . . . . . . . . 50 4.3.2 Trap Voltages and Control . . . . . . . . . . . . . . . . . . . . 52 4.3.3 Oscillating Trap Voltage . . . . . . . . . . . . . . . . . . . . . 54 4.4 Optical Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56 4.4.1 Ionization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56 4.4.2 935 nm Repump . . . . . . . . . . . . . . . . . . . . . . . . . . 57 4.4.3 638 nm Repump . . . . . . . . . . . . . . . . . . . . . . . . . . 58 4.4.4 Cooling, Pumping, and Detecting with 369 nm . . . . . . . . 58 4.4.5 Iodine Spectroscopy Lock . . . . . . . . . . . . . . . . . . . . 65 vi 4.4.6 Vanguard 355 nm Laser . . . . . . . . . . . . . . . . . . . . . . 68 4.4.7 Driving Raman and M lmer-S renson Transitions with the Vanguard . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72 4.4.8 Florescence Collection . . . . . . . . . . . . . . . . . . . . . . 78 4.5 Arbitrary Waveform Generator (AWG) . . . . . . . . . . . . . . . . . 87 5 Experimental Procedure 90 5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90 5.2 Temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90 5.3 Frequencies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91 5.4 Loading and Alignment . . . . . . . . . . . . . . . . . . . . . . . . . . 92 5.5 Raman Beams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94 5.6 Calibration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96 5.7 Imaging Calibration . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98 6 Summary of Experiments 99 6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99 6.2 Experimental Demonstration of the Ising interaction . . . . . . . . . 100 6.2.1 Two spin case . . . . . . . . . . . . . . . . . . . . . . . . . . . 102 6.2.2 Three spin case . . . . . . . . . . . . . . . . . . . . . . . . . . 103 6.3 Adiabatic quantum simulation . . . . . . . . . . . . . . . . . . . . . . 105 6.4 Phase diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109 6.5 Spin frustration and entanglement . . . . . . . . . . . . . . . . . . . 112 6.6 Scalability of the quantum simulation . . . . . . . . . . . . . . . . . 117 6.7 Scaling of imperfections . . . . . . . . . . . . . . . . . . . . . . . . . 121 6.7.1 Spin-motion coupling . . . . . . . . . . . . . . . . . . . . . . . 121 6.7.2 Diabaticity . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122 6.7.3 Spontaneous Emission . . . . . . . . . . . . . . . . . . . . . . 122 6.7.4 Intensity uctuations . . . . . . . . . . . . . . . . . . . . . . . 123 6.7.5 Detection Errors . . . . . . . . . . . . . . . . . . . . . . . . . 123 6.8 Conclusion and Outlook . . . . . . . . . . . . . . . . . . . . . . . . . 128 7 Quantum Simulation of the Devil?s Staircase 130 7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130 7.2 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131 7.3 Experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132 7.4 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148 8 Simulating the Ising Model with Arbitrary Control of the Couplings 153 8.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153 8.2 Control of an Arbitrary Lattice Hamiltonian . . . . . . . . . . . . . . 154 8.3 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157 8.4 Implementation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158 8.5 Scalability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163 8.6 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165 vii A Linear Trap External Pi Filter 167 B Optical Diagram Legend 169 C AWG Function Library 170 D Mathematica ROI and Discriminator Selection Script 181 Bibliography 186 viii List of Figures 2.1 171Yb+ States and transition wavelengths . . . . . . . . . . . . . . . . 18 2.2 Doppler cooling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21 2.3 Optical pumping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23 2.4 171Yb+ Detection Scheme . . . . . . . . . . . . . . . . . . . . . . . . . 25 2.5 Single ion histogram . . . . . . . . . . . . . . . . . . . . . . . . . . . 26 2.6 Idealized histograms of multiple ions . . . . . . . . . . . . . . . . . . 27 3.1 Raman . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30 3.2 RSB . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33 4.1 Paul Trap . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41 4.2 Transverse Modes For 10 Ions . . . . . . . . . . . . . . . . . . . . . . 47 4.3 Spin-Spin Couplings for 10 Ions . . . . . . . . . . . . . . . . . . . . . 48 4.4 Linear trap 3D diagram . . . . . . . . . . . . . . . . . . . . . . . . . 50 4.5 Trap Control Program . . . . . . . . . . . . . . . . . . . . . . . . . . 53 4.6 Trap Control Program - Electrodes Tab . . . . . . . . . . . . . . . . . 55 4.7 MBR-110 Laser . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59 4.8 369 nm Laser and Optics Setup . . . . . . . . . . . . . . . . . . . . . 63 4.9 Chamber and Resonant Beams . . . . . . . . . . . . . . . . . . . . . . 64 4.10 Iodine Spectroscopy Setup . . . . . . . . . . . . . . . . . . . . . . . . 67 4.11 MLUV Main Window . . . . . . . . . . . . . . . . . . . . . . . . . . . 69 4.12 MLUV Oscillator Service . . . . . . . . . . . . . . . . . . . . . . . . . 70 4.13 MLUV Ampli er Settings . . . . . . . . . . . . . . . . . . . . . . . . 71 4.14 MLUV Oven Settings . . . . . . . . . . . . . . . . . . . . . . . . . . . 71 4.15 MLUV Position Controls . . . . . . . . . . . . . . . . . . . . . . . . . 72 4.16 Frequency Combs for Raman . . . . . . . . . . . . . . . . . . . . . . . 74 4.17 Vanguard Raman Setup . . . . . . . . . . . . . . . . . . . . . . . . . 75 4.18 Raman Beams Close Up . . . . . . . . . . . . . . . . . . . . . . . . . 79 4.19 Raman Pro le . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80 4.20 Florescence Collection . . . . . . . . . . . . . . . . . . . . . . . . . . 81 4.21 10 Bright Ions, Binned . . . . . . . . . . . . . . . . . . . . . . . . . . 82 4.22 Single Ion on ICCD . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83 4.23 ICCD Histograms of 10 Ions . . . . . . . . . . . . . . . . . . . . . . . 86 5.1 Trap Corner on ICCD . . . . . . . . . . . . . . . . . . . . . . . . . . 94 5.2 10 Ion Raman Spectrum . . . . . . . . . . . . . . . . . . . . . . . . . 96 5.3 Automatic Side-Band Finder . . . . . . . . . . . . . . . . . . . . . . . 97 6.1 Two Spins Coupled . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102 6.2 Three Spins Coupled . . . . . . . . . . . . . . . . . . . . . . . . . . . 104 6.3 Energy Level Diagram for 3 Spins . . . . . . . . . . . . . . . . . . . . 107 6.4 Measured Ferromagnetic Order . . . . . . . . . . . . . . . . . . . . . 108 6.5 Phase Diagram for Three Ions . . . . . . . . . . . . . . . . . . . . . . 110 ix 6.6 Spin Frustration on Triangular Lattice . . . . . . . . . . . . . . . . . 113 6.7 Entanglement Witness . . . . . . . . . . . . . . . . . . . . . . . . . . 115 6.8 FM Simulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119 6.9 FM Data vs Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124 6.10 FM Magnetization . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127 7.1 6 Ion Spin-Spin Coupling Decay . . . . . . . . . . . . . . . . . . . . . 134 7.2 Controlling the Phases of Simulation Pulses . . . . . . . . . . . . . . 137 7.3 Magnetization Phase Diagram for 6 Ions . . . . . . . . . . . . . . . . 139 7.4 Axial Simulation for 6 spins - States . . . . . . . . . . . . . . . . . . . 141 7.5 6 Ion Staircase . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142 7.6 6 Ion Complete Energy Spectrum . . . . . . . . . . . . . . . . . . . . 143 7.7 6 Ion Low Energy Spectrum . . . . . . . . . . . . . . . . . . . . . . . 143 7.8 6 Ion Exponential Fit . . . . . . . . . . . . . . . . . . . . . . . . . . . 145 7.9 Filtered 6 Spin Staircase . . . . . . . . . . . . . . . . . . . . . . . . . 146 7.10 Power Filter VS Adiabaticity . . . . . . . . . . . . . . . . . . . . . . 147 7.11 10 Ion Coupling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148 7.12 Axial Simulation for 10 spins - States . . . . . . . . . . . . . . . . . . 149 7.13 10 Ion Exponential Fit . . . . . . . . . . . . . . . . . . . . . . . . . . 150 7.14 Power Filtered 10 Ion Data . . . . . . . . . . . . . . . . . . . . . . . 151 8.1 Spectrum of N Ions with N Detunings . . . . . . . . . . . . . . . . . . 155 8.2 Example Solutions for Arbitrary Lattice Generation . . . . . . . . . . 159 8.3 Micromirror Array Simulator . . . . . . . . . . . . . . . . . . . . . . . 160 8.4 AOM Array Simulator . . . . . . . . . . . . . . . . . . . . . . . . . . 161 8.5 Phase Array Simulator . . . . . . . . . . . . . . . . . . . . . . . . . . 162 8.6 1D Ising chain Scaling for Simulator . . . . . . . . . . . . . . . . . . . 164 A.1 Pi Filter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167 B.1 Optical Diagram Legend . . . . . . . . . . . . . . . . . . . . . . . . . 169 x List of Abbreviations and Symbols h Reduced Plank Constant KB Boltzman Constant AFM Antiferromagnetic AOM Acousto Optic Modulator BSB Blue Sideband COM Center of Mass CW Continuous Wave EOM Electro Optic Modulator FPGA Field Programmable Gate Array FM Ferromagnetic PCB Printed Circuit Board QE Quantum E ciency SE Schr odinger Equation SHG Second Harmonic Generator THG Third Harmonic Generator RF Radio Frequency ROI Region of Interest RSB Red Sideband UHV Ultra High Vacuum VCA Voltage Controlled Attenuator IREAP Institute for Research in Electronics and Applied Physics NA Numerical Aperture NSA National Security Agency xi Chapter 1 Introduction The work presented in this thesis is complex, and covers many subject areas, which cannot be given a full, bottom up presentation in this work. Furthermore, I do not posses this knowledge. Six years pursuing an experimental result does not allow for such a pursuit. Instead, I will present mostly information needed to understand the motivation for this work and the minimal amount of information needed to follow it as it is exposed. A deeper understanding can be pursued by the interested reader by reading the cited works. Many of the results presented here have already been presented in previously published works to which I have contributed in work and writing, but all the results presented are the fruits of a team e ort. Quantum systems are notoriously di cult to model e ciently using classical computers, owing to the exponential complexity in describing a general quantum state as the system grows in size. In the 1980s, Richard Feynman proposed to circumvent this problem by employing a control quantum system with tailored in- teractions and logic gates between quantum bits (qubits) in order to simulate the quantum system under investigation [4, 5]. Quantum spin models such as the Ising model have become a proving ground for Feynman?s proposal, with systems of qubits behaving as e ective spins and interactions engineered with external electromagnetic elds. The quantum Ising 1 model is the simplest spin Hamiltonian that exhibits nontrivial aspects of quantum magnetism such as spin frustration, phase transitions [6], and poorly understood spin glass and spin liquid phases [7, 8]. Indeed, solving for the ground state con guration of spins subject to a general fully-connected Ising interaction is known to be an NP-complete problem [9]. At the conclusion of this thesis I will show how a fully- connected Ising or more general Heisenberg spin model with arbitrary couplings across the spin network can be generated in a scalable system of trapped atomic ions, even for a one-dimensional chain in space. This may allow quantum simulations with hundreds of spins, where the physics cannot generally be predicted otherwise. 1.0.1 Classical Computation Over the last century computers and electronics have made enormous advances and transformed our world our and lives. The theoretical foundations of modern computers are rooted in the work of Charles Babbage in the mid 19th century. It was in the late 19th century, when the need to handle large amounts of data, and the emergence of the vacuum tube, spurred the creation of real world, albeit primitive computers. The conception of solid state transistors in the 1920?s and their technological application in the 1950?s enabled a breakneck acceleration in the development of computers. The constant rate of transistor miniaturization led to the observation that the number of transistors on a processor doubles every two years (\Moore?s Law")and processor performance doubles every 18 months. The average smartphone 2 that a consumer carries in their back pocket today is more powerful than a 1990?s Cray-2 supercomputer! Although the available computational power is increasing, this growth cannot continue forever, as technological hurdles such as thermal management and impu- rities become more severe as the density of transistors and processors continues to increase, and physical obstacles such as unwanted quantum e ects due to shrinking component size threaten reliability. Currently, the smallest transistors fabricated in mass are 22 nm in length [10]. Furthermore, there are more fundamental limits to computation speed. A computational algorithm can be analyzed abstractly, and a relationship between the number of steps or other resources needed to solve a problem and the size of the problem can be established. An important problem in computer science is factoring a large number into two prime numbers, as this can be used to decipher encrypted communications (using the RSA method). With known methods, the number of steps to nd the prime factors grows exponentially with a linear growth in the number of digits of the problem number. This makes this problem practically unsolvable and hence the encryption is secure. 1.0.2 Quantum Information Processing However, it is known that a computer with components that remain in a pure quantum state (undisturbed by uncontrolled external in uences) only requires a 3 polynomial increase in steps [11]. This surprising result shows that quantum com- puters likely posses computational powers far beyond those of classical computers [4] (this is only a hypothesis though as the fundamental theories of computation are still in their infancy). The other motivation for the development of quantum information processing machines is the problem of simulating quantum systems. As Feynman proposed in 1982 [4, 5], a well-controlled quantum system could e ciently simulate the behavior of a complex quantum model that is classically intractable. For example, a quantum simulator could be used to determine properties and dynamics derived of poorly un- derstood models in condensed matter such as quantum magnetism [12], spin glasses [7], spin liquids [13] and high temperature superconductors [14, 15]. As the resources required to simulate quantum systems on a classical computer grow exponentially with linear growth of the quantum system, a machine that is itself in a quantum state analogous to a system we wish to investigate will naturally bene t from this growth in its computational space as its number of components grows linearly. This is likely the more important application of quantum informa- tion processing, as the number of digital quantum computing algorithms known is limited, but the number of quantum systems whose understanding can bene t are unlimited [16]. Furthermore, it has been shown that a universal quantum simulator is equiv- alent in computational power to a universal quantum computer up to a polynomial factor in the time needed for computation [17], and as such cannot simply be dis- counted as a more limited form of quantum computing. 4 1.1 The DiVincenzo Criteria In order to implement the ideas discussed above, one needs a practical quantum system. DiVicenzo presented the grocery list of characteristics of such a system : Identi cation of well-de ned qubits Reliable state preparation Low decoherence Accurate quantum gate operations and Strong quantum measurements Although originally speci ed for a circuit quantum computer, the requirements for a quantum simulator are identical. The 171Yb+ system has all of these charac- teristics. The qubits are well de ned in the sense that each 171Yb+ ion is exactly the same. The only variation between these qubits is due to local gradients in electric and magnetic elds. The technical details of the 171Yb+ qubit will be discussed in detail in chapter 2. For now, it will be simply stated that in comparison to compet- ing implementations (mainly solid state Josephson junction based systems [18, 19]) of qubits, the ion systems are superior in all of these criteria except one that is not listed - scalability. The storage of ions for long periods of time is di cult as it requires strong electric elds and very low pressures. In this work, we were limited to working with 16 ions. Although we faced multiple sources of errors that would prevent us from controlling more than 16 ions, the main limiter is our inability to 5 store in a stable crystal the ions for more than 10 minutes. This is likely due to background atoms colliding with the chain, even at our working pressure of 10 12 Torr. 1.2 Quantum Mechanics The development of quantum mechanics was spurred by false predictions and contradictions encountered by physicists in the late 19th and early 20th century. These problems spanned several di erent aspects of physics. First, there was the famous ultraviolet catastrophe, which predicted the radi- ation of in nite energy from a blackbody at the limit of short wavelengths. Second there was the photoelectric e ect, where light shining on a metal would cause a current that would inexplicably cease when the light frequency was reduced. Third, the double-slit experiment demonstrated wave-like behavior for light, except when a detector was placed at one of the slits leading to the ba ing appearance of particle- like behavior. The fourth and most problematic discovery was that atoms were most likely composed of electrons in orbit around a small, positive nucleus. A classical description of this would seem to make matter itself impossible due to the electrons completely radiating away their energy, unless they were limited to discrete orbitals for an (at the time) inexplicable reason. These problems were resolved by the intro- duction of the photon, the de Broglie wave equation, and nally the development of quantum mechanics. In the quantum mechanical description of physical systems the state is de- 6 scribed by a vector in a Hilbert space. This vector and its associated conserved quantities (\quantum numbers") describe all that can be known about the system. The dynamics of such a system is described by the Schr odinger equation H j i = i h _j i (1.1) Where H is the Hamiltonian, an operator in Hilbert space that is the sum of the energy terms in the system, j i is the wavefunction, a vector in Hilbert space, and h is Planck?s constant, de ned as the ratio between photon energy and angular frequency. The most revolutionary aspect of this new theory is the wave-function descrip- tion of nature: although particles are discrete (they are always observed as particles), some of their properties can only be described probabilistically. This probability is given by the probability amplitude h j i = R dx integrated over the region of interest (this is the inner product of the vector in Hilbert space). The other ob- servables of the system are extracted from the wavefunction using an operator, such that the expectation value for quantum mechanical observable A is hAi = h j A j i (1.2) where A is a Hermitian matrix (so as to always have real expectation values). 7 If we assume the Hamiltonian is time independent, we can rewrite Eq. 1.1 E j i = H j i (1.3) by assuming that j i has an oscillatory form j (t)i = e iEt= h j (0)i (1.4) and E is the energy of the eigenvector j (t)i . In general, we de ne j (t)i = U(t) j (0)i where U(t) = Te i R Hdt= h (1.5) is the evolution operator, which simpli es to U(t) = e iHt= h for time independent Hamiltonians.1 1.3 Adiabatic quantum simulation One method of quantum information processing is adiabatic quantum simula- tion. This approach is motivated by the adiabatic quantum computation algorithm rst proposed as a method to solve NP-complete satis ability problems [20]. The process of quantum adiabatic computation works as follows: a quantum system is initialized to the ground state of a trivial Hamiltonian. Next, the Hamiltonian is 1T is Dyson?s time ordering operator. This operator handles the non-commutation of H(t) for di erent t. 8 adiabatically deformed into the Hamiltonian of interest, whose ground state encodes the solution of a problem that has been mapped to the nal Hamiltonian. If success- ful, the system will remain in the ground state and can be directly probed once the system arrives at the desired Hamiltonian. For quantum simulation of magnetically interacting spins, this approach allows the determination of ground states where the Hamiltonian can easily be written, yet the spin ground state cannot always be predicted, even with just a few dozen spins [21]. A precondition for the use of this state is satisfying the adiabaticity criterion[22] 2: T >> g2min (1.6) where T is total simulation time, gmin = min(E1(t) E0(t)) is the minimum energy gap between the ground state (or any other state we can trivially initialize) and clos- est coupled state, where = max(jh1; tj dHdt j0; ti j) is the change in the Hamiltonian at time t for the coupling between those energy levels. 1.3.1 Hydrogen-Like Atom In this work we be using 171Yb+ ions as qubits. To understand their behavior, we will model them using a quantum mechanical model, rst developed to describe the simplest atom - Hydrogen. In order to model the Hydrogen atom and other atoms that have only one valence electron, we assume a simple model of a negative electron orbiting a stationary positive nucleus in a spherically symmetric electric 2This is the criterion usually encountered in the literature. However, this criterion is not in true in general [23] 9 potential. Then we write our Hamiltonian H = V + T = 1 4 0 Ze2 r + p2 2m0 (1.7) where Z is the number of positive nuclear charges, e is the charge of the electron, r is the distance of the electron from the nucleus, 0 is the electric permittivity constant, and p 2 2m0 is the kinetic energy of the electron (m0 is its mass): h 2 2m0 r2. We nd the eigenfunctions that satisfy the Schr odinger equation: (r; ; ) = R(r)Y ( ; ) (1.8) Y ml ( ; ) s (2l + 1)(l jmj)! 4 (l + jmj)! eim Pml (cos ) (1.9) R(r)nl s 2Z na 3 (n l 1)! 2n[(n+ l)!]3 e rZ=na 2Zr na l [2L2l+1n l 1(2Zr=na)] (1.10) a 4 0 h 2 mRe2 (1.11) Where Pml are the associated Legendre functions, 2L 2l+1 n l 1 are the associated La- guerre polynomials, and mR is the reduced mass of the nucleus and electron. n; l and m are quantized numbers that label orthogonal eigenfunctions. These numbers are described as \good" quantum numbers. Their goodness is a result of the simplifying assumptions of the model. In the 171Yb+ ion we will see that these assumptions are only roughly true, leading to metastable states de ned by other quantum numbers. For Hydrogen-like atoms, n determines the energy of the sta- 10 tionary state, l (the azimuthal quantum number) orbital angular momentum., and m (the magnetic quantum number) the direction of angular momentum vector. The de nition of the Legendre polynomials restricts the values of l and m: l = 0; 1; 2:::;m = l; l + 1; :::; 0; 1; ::l (1.12) The Laguerre polynomial solution is found when one restricts the radial solution R(r) to a non-diverging series. The same restriction truncates l at n 1. Intuitively, we can understand that a nite kinetic energy implies nite angular momentum. Using the standard de nition of angular momentum, L = r p and our pre- vious implicit de nition of the momentum operator, we can show the commutation relations for the Cartesian components of L to be [Li; Lj] = i h ijkLk (1.13) and if we de ne the raising and lowering operators as L+ = Lx + iLy L = Lx iLy (1.14) 11 then we nd that these are ladder operators for the atomic eigenstates L+ jl;mi = h p (l m)(l +m+ 1) jl;m+ 1i (1.15) L jl;mi = h p (l +m)(l m+ 1) jl;m 1i (1.16) There are additional complications to this simple model we must address before we can discuss the 171Yb+ ion and its energy structure. First, the electron posses an innate angular momentum, \Spin", which is described by exactly the same math as angular momentum, however l for the electron is xed at 1=2 and so m only spans 1=2; 1=2. We label the operators for spin ~S h2~ where ~ is the Pauli spin vector - a vector of operators with components x 0 B B @ 0 1 1 0 1 C C A ; y 0 B B @ 0 i i 0 1 C C A ; z 0 B B @ 1 0 0 1 1 C C A (1.17) Complicating thins further, the proton(s) in the nucleus of the atom also have spin. These spins give the proton and electron a magnetic moment, and these experience a magnetic dipole force due to the magnetic eld created by the orbiting electron. These add additional terms to the the total atom Hamiltonian, of the form Hd = B. We de ne two new quantities, the total orbital angular momentum J L + S and F I + J , where S is the total electronic spin and I is the total nuclear spin. For most stable atomic states, where we ignore coupling between the outer electron and the inner electrons (LS coupling), we will label the atomic states using 12 the ?Russel-Saunders? convention 2S+1LJ (1.18) The interaction of the electron with the magnetic eld created by the rotating nucleus (in the electron?s rest frame) gives rise to the ne structure splitting of the energy levels, and the interaction of the nucleus? dipole moment with the magnetic eld created by the electron leads to a smaller perturbation of the energy levels, called the hyper ne splitting. In atoms such as 171Yb+ , where the nuclear spin magnetic moment is non-zero (due to the odd number of nucleons), the hyper ne coupling will be strong, and so it will be necessary to add an energy term to the Hamiltonian for the dipole coupling between the nucleus? magnetic moment and the magnetic eld from the electron, HHF = I (Bl +Bs). However, the dominant coupling term will be that of the total angular momen- tum, so we will write the eigenstates of the atom as superpositions of our previously uncoupled basis: jJ;Mi = X ml+ms=mj CL;S;L+Sml;ms;mj jl;mli js;msi (1.19) where the coe cients CL;S;L+Sml;ms;mj are the Clebsch-Gordan coe cients. These will be necessary to calculate transition strengths. 13 1.3.2 Transition Rules What is the signi cance of all this? Well, from here we arrive at the rules which tell us what atomic transitions are allowed and which are forbidden. The restriction of atomic transitions will allow us to construct a useful qubit. When the atom absorbs a photon, angular momentum is conserved. We con- sider a perturbation H1 to the atom from an external electromagnetic eld. The probability for a transition from a stationary state 1 to 2 will depend on the matrix element h 1jH1 j 2i . Considering only the dipole component of the multi- pole expansion of the eld, the perturbation to the potential of the electron will be / e~r ~E which requires l = 1 and m = 1; 0 for a non-zero transition prob- ability. We will see how these rules come in to play when we examine the allowed transitions of 171Yb+ in chapter 2. 1.3.3 Interaction Hamiltonian For many systems, the Hamiltonian is a sum of a static term HS and an interaction term HI . If this is the case we may simplify our analysis by rotating into the static frame, so that the time evolution is only a function of HI . So our total Hamiltonian in the Schr odinger picture (where all operators are static in time) is HT = HS +HI (1.20) 14 And in the interaction ?picture? I(t) = eiHSt= h T (t) (1.21) HIS = e iHSt= hHSe iHSt= h = HS (1.22) HII = e iHSt= hHIe iHSt= h (1.23) where the superscript I (T ) indicates interaction (Schr odinger ) picture. We see that the static term remains una ected, but now the interaction term in the interaction picture is time dependent. Rewriting the Schr odinger equation 1.1 in terms of the interaction picture, the static component of the Hamiltonian drops out and we are left with HII I(t) = i h _j I(t)i (1.24) and now the dynamics only depend on the interacting component of the Hamil- tonian. 1.3.4 E ective Hamiltonian For highly complex systems the interaction picture is still too complex to allow for analytical analysis. For many of these systems, including the spins in the ion trap, simplifying assumptions can be made to rewrite an even simpler Hamiltonian. The following treatment, succinctly presenting the formalism developed by D.F.V. 15 James[24], can be applied to interaction Hamiltonians of the form HI(t) = NX n=1 Hne i!nt +Hyne i!nt (1.25) where we are summing over N oscillatory terms, and 0 < !1 < !2::!N Then we can derive the e ective Hamiltonian given by Heff (t) = NX n;m=1 1 h !mn Hm; H y n ei(!m !n)t (1.26) and 1 !mn = 1 2( 1 !m + 1!n ). This Hamiltonian is the result of several approximations. First the interaction evolution operator 1.5 is time averaged, so that all fast changes are neglected (general case of rotating wave approximation). The second approximation is neglecting the higher order terms of the averaged evolution operator?s time ordered expansion. Applying a second rotating wave approximation leads to 1.26. We will utilize this powerful tool when we treat light-ion interactions in 3. 1.4 Conclusion This brief introduction to quantum information processing with ions and quan- tum mechanics has laid down the ground work for understanding the quantum as- pects of manipulating a 171Yb+ ion, that will be introduced in the following chapter. 16 Chapter 2 The 171Yb+ ion Now that we have discussed the basic physics necessary to understand atomic transitions and energy states, we will introduce our atomic system of choice: the 171Yb+ ion[1]. We will discuss in 4.2 how we actually trap and create these ions from a Yb metal. The discussion of how we generate and direct the light on the ion will have to wait until chapter 4. Many ions are used throughout the world in quantum information experi- ments, mostly from the alkaline earth metals. These are chosen so that once singly ionized the ion will have a hydrogen-like electronic structure, which is more easily understood and controlled. The second important feature is that the energy structure is such that the ion can be forced through optical pumping and cycling transitions to be constrained to two energy levels, so that it behaves as an e ective spin (a qubit). Our ion has a nuclear spin of 1/2, giving it the simplest hyper ne structure possible in its ground state. For our qubit, we will be using the ground state 2S1=2 energy manifold, and the clock states jF = 0;mf = 0i as our \down" (j#i ) state and jF = 1;mf = 0i as our \up" (j"i ) state. These states are the eigenstates of the z operator with 17 369.5262nm (99.5% ) 638.6151n m 638.6102n m F=1 F=0 F=0 F=1 F=1 F=2 F=1 F=0 F=3 F=4 329n m 12.642812118466 GHz+310.8B2Hz 2.105 GHz 2P1/2 935.1879n m 2S1/2 2P3/2 3D[3/2]1/2 3D3/2 2D5/2 2F7/2 1D[5/2]5/2 F=3 F=2 100 THz F=3 F=2 Figure 2.1: The 171Yb+ ion states with transition wavelengths. eigenvalues 1 and -1 respectively. The clock states are so named as they are used in atomic clocks due to only having a second order Zeeman energy shift, making them robust to magnetic noise. Correspondingly, this ensures the qubits a long T2 coherence time - 2.5 s [1] without magnetic shielding, orders of magnitude longer than non-clock state qubits with magnetic shielding or magnetic- eld feedback. We use the resonant 2S1=2 to 2P1=2 transition for cooling, state initialization and detection. As we scatter on this transition, there is a 0:5% chance of decaying out of this cycle to the 2D3=2 state. In order to correct for this we irradiate the ion with 935:1879 nm light modulated at 3:07 GHz in order to remove the population from 2D3=2 jF = 1i and 2D3=2 jF = 2i to the 3D[3=2]1=2 state. This is a metastable state (?Bracket State?) where the usual ?good? quantum numebrs are no longer valid 18 1. The ion has a 98:5% chance of decaying back to 2S1=2 jF = 1i from this state, so it is quickly restored to the correct qubit state (this will however create an error in the phase of one qubit relative to other qubits in the chain). A much less likely event is the ion transitioning to the 2F7=2 state. This happens once in about an hour per ion. It is likely caused by a collision with a background atom in the Ultra High Vacuum Chamber (UHV), as this is a forbidden octopole transition [26]. Previously, we have used 638:61 nm light that is scanned between the two resonance frequencies. We have found that our intense 355 nm Raman beams are actually better at retrieving the ion from this long lived ( 6 years) dark state, but it is possible that this is due to YbH+ formation rather than transition to the 2F7=2 state [27]. The 355 nm radiation would then disassociate this ionic molecule and return the ion to a coolable state. 2.1 Ionization The source of our ions is a nearly pure source of neutral 171Yb+ . This source is a solid metal, sliced into small shards that are then forced into a narrow stainless steel tube, which is clamped o at one end. A wire is spot welded to the tube, and the tube is fed into a titanium holder, designed to have a high electric resistance and low thermal conductivity so as to isolate the Ohmic heating to the 171Yb+ oven. A current of roughly 2.3 Amps is run through this circuit (the holder is grounded 1In this state the outer electron couples to the core electrons, and the state is written as 2s0+1[K]J , Where s0 is the spin of the outer electron, K is the total angular momnetum of the core and the angular momentum of the outer electron, and J is K and the spin of the outer electron. The letter preceding the bracket is not part of the notation but is L for the core in the usual LS notation. [25] 19 to the chamber). The oven is oriented towards the trapping region, where a 399 nm beam will excite the neutral Yb from 1S0 to 1P1. A powerful 355 nm beam or 369 nm beam will then excite an electron to the continuum. Some of these 171Yb+ ions will be su ciently cooled by Doppler cooling to be trapped. 2.2 Dopppler Cooling Doppler cooling is achieved by scattering red detuned light from the resonant transitions [28] for the 2S1=2 to 2P1=2 transitions, as illustrated in Fig. 2.2. The light is tuned to a wavelength of 369.521525 nm , and detuned 25 MHz from resonance, while the transition has a 20 MHz linewidth. To ensure that the ion is cooled for both F=0 and F=1 states of the 2S1=2 manifold, the light is modulated by an Electro Magnetic Modulator (EOM) at half the combined hyper ne splitting of the 2S1=2 and 2P1=2 states. The second order sideband allows the o resonant cooling of the j#i state. Doppler cooling utilizes the Doppler shift of the resonant frequency in the frame of the moving ion, such that when the ion is moving towards the oncoming light it is blue shifted towards resonance. The ion will then spontaneously emit a photon into a random direction, so on average losing momentum in the direction of the oncoming beam. The temperature limit of this random process is the ?Doppler cooling limit?, h =2KB, where is the linewidth of the transition and KB Boltzmann?s constant. For the S-P transition linewidth of 19.7MHz, this corresponds to 5 10 3 K. The relation KBT = n h! [29] implies for our transverse direction of motion an average of 12 phonons after Doppler cooling. 20 F=1 F=0 F=1 2P1/2 2S1/2 F=0 mF=- 1 mF=1 mF=0 Figure 2.2: Doppler cooling the 171Yb+ ion. Note that the 935 nm repump transition is not shown here. The wide lines represent the lasers driving transitions. The narrow lines represent the allowed decays. 21 2.3 Optical Pumping Once the ion chain is su ciently cooled (after Doppler cooling and sideband cooling), we must prepare the ions in a pure and known quantum state. This can be done with high delity. We apply light resonant with a transition from the j"i state to the 2P1=2 F=1 state (as shown in 2.3), leading to a cycle that with our beam power of 10 W pumps the ion to the j#i state within 1.5 s with over 99% delity. 2.4 Detection The detection scheme of the 171Yb+ ion relies on the large separation of j"i and j#i states versus the linewidth of the detection transition. We use light reso- nant with the j"i to the 2P1=2 jF = 0i transition, and polarize the light so as to be resonant with all the mf states (i.e. both linear and circularly polarized light). The linewidth of this transition is 19:7 MHz , while the laser light linewidth is 100 KHz wide, and the separation between the j"i and j#i states is 12.6 GHz , so o - resonant scattering is small (but not negligible). Also, note that transitions from 2P1=2 jF = 0i to j#i are forbidden by transition laws. This allows us to create a closed cycle of absorption and emission - a cycling tran- sition. There is a small complication here - without a magnetic eld to break the degeneracy, a coherent dark state may occur [30]. In this situation, the bright states interfere destructively, and detection fails. However, we apply a magnetic eld via an electromagnet under the UHV chamber. This magnet serves multiple purposes - 22 F=1 F=0 F=1 2P1/2 2S1/2 F=0 mF=- 1 mF=1 mF=0 Figure 2.3: Optical pumping of the 171Yb+ ion. 23 it de nes the quantization axis (the Z-spin axis corresponding to the z operator), and it breaks the degeneracy of the 2S1=2 manifold, thus preventing this coherent dark state. We estimate the eld to be roughly 5 Gauss, providing a Zeeman shift of 7 MHz (310 52 Hz ). This can be measured by taking a Raman spectrum of the excitable sidebands. As the ion uoresces, the distribution of detected scattered light should ide- ally follow a Poissonian. However, the possibility of o resonant excitation of the dark state forces us to convolute this distribution with an exponential function[31]. A similar analysis is appropriate for the dark state. The atomic physics derived distributions are deformed by complicating factors. First, there is unwanted scatter o the electrodes of the trap. Second, there is the device physics of the imaging device, which will lead to spurious counts from false photon detections and noise during electrical readout. I will discuss this more in depth in 4.4.8. When detecting the state of a single ion, we can expect a histogram of counts as in gure 2.5. As is seen in the gure, there are two distributions, one centered around the average dark count, and one around the average bright state count of roughly 10 counts. In this case we can use a discriminator value that leads to an equal probability of mistaking j"i for j#i and vice versa. i.e the overlap of the distributions determines our single spin detection error. However, when we must detect multiple ions, we encounter two problems when simply collecting the light into a Photo Multiplying Tube (PMT). First, we do not have the ability to detect the true eigenstate of the chain, only how many ions are bright. This is acceptable in some experiments, where total number of bright ions 24 F=1 F=0 F=1 2P1/2 2S1/2 ?+ ?? ? F=0 F=1 F=2 F=1 F=0 9 3 5.1 8 7 9 n m 3D[3/2]1/2 3D3/2 0.86 GHz 2.2095GHz mF=- 1 mF=0 mF=1 mF=0 Figure 2.4: The 171Yb+ ion detection scheme. The additional 3D F=2 to F=1 resonant light is needed for the optical pumping, since the ion may occupy either 3D3=2 hyper ne states. Only pertinent states and transitions are shown. is the quantity of interest. Second, as the number of spins increases, the average number of bright counts per ions bright should naively increase as Nions x average count for one ion, assuming a uniform detection beam intensity across the chain (we will see this is not exactly the case). But this leads to increasing overlap between the di erent distributions, so even for the task of just detecting the total number of ions bright the PMT will under-perform compared to a device with spatial information, such as a Charge Coupled Device (CCD) camera. 25 Figure 2.5: Measured count histogram after Doppler cooling with typical experimental conditions. The ion is Doppler cooled for 3ms, during which the ion scatters 10 photons on average into the light collecting Photo Multiplying Tube (PMT). Here, we are not detect- ing the scattered Doppler cooling light. Rather, we are following the cooling by a 800 s resonant detection pulse and light collection. This is repeated 100 times and then a histogram is constructed from the number of counts collected by the PMT via the imaging optics for each detection. As can be seen from the graph, the ion is not com- pletely bright following cooling. This screen shot from the control program also shows the Poissonian ts for the dark and bright states. The mean value of the ts are entered as parameters. The tted am- plitudes are interpreted as probabilities for the number of ions bright. 26 0 20 40 60 80 0 20 40 60 80 Photon Counts o fEvent s Figure 2.6: Poissonians for the dark state and up to ve ions bright. These are idealized bright states as there is no leakage to the dark state. The leftmost distribution is the dark state, and the distribution peaking at 10 counts is a single bright ion. The other distributions are centered at multiples of 10 (N 10), corresponding to N ions bright. Realistic distributions would have worse overlap due to this e ect, including a broadening of the dark histogram as well. 27 Chapter 3 Quantum Control of The Ion 3.1 Light-Ion Interaction The last ingredient necessary to make the ion a useful qubit or simulated spin is the ability to control its internal state. For the purpose of this exposition I will utilize the e ective Hamiltonian theory introduced in 1.3.41. The simplest application of this theory is to the analysis of the AC Stark shift [24]. Applying an o -resonant eld, detuned by from the hyper ne splitting of the qubit states !hf with coupling coe cient , we derive the interaction Hamiltonian HI = h 2 (j"ih#j e i h t + j#ih"j ei h t) (3.1) Applying equation 1.26 to the single harmonic term H1 to get HStark = h2 2 4 1 2 h ( j"ih#j j#ih"j 2 j#ih"j j"ih#j 2 ) (3.2) resulting in HStark = h 2 4 (j"ih"j j#ih#j ) (3.3) 1I am still following here the compact approach presented by James, which simpli es the treat- ment but somewhat obscures the physical justi cations. For a di erent and lengthier presentation, see [32]. 28 We see that the laser is dressing the qubit states and shifting their e ective energy in opposite directions and by equal amounts. This is equivalent to a phase gate on a single spin - as the spin is precessing while shifted by the eld, it will become out of phase with other spins not under the in uence of the eld. Alternately, it may become out of phase with an oscillator that was supposed to be locked to the spin?s natural precession frequency - a persistent concern for the experiment. We will see in 4 how we can mitigate this. It is important to note that this is only the simplest A.C. Stark shift possible, as a Stark shift can be a multi-photon process, rather than the two photon process presented here. 3.2 Raman Transitions We now add a third, higher energy level to the system, jei , as depicted in Fig 3.1. Applying two light frequencies !1 and !2 detuned by from jei , the beat frequency !hf will drive this transition, while only negligibly populating jei . We write the interaction Hamiltonian HI = h 1 2 jeih"j e i t + h 2 2 jeih#j e i t + h:c: (3.4) And apply 1.26: Heff = h 21 4 (jeihej j"ih"j ) h 22 4 (jeihej j#ih#j )+ h 1 2 4 (j#ih"j j"ih#j ) (3.5) where i = q Ii ISat is the coupling coe cient for beam i, Ii is the intensity of beam i, 29 ?? ?? ? ?? ?? ?? ??? Figure 3.1: O resonant coupling of the spins. jei is negligibly pop- ulated for >> i 30 ISat is the saturation intensity of the transition for that beam, and is the radiative linewidth for that transition. We see the e ect is an AC Stark shift on all levels (a symmetric e ect for both qubit levels for equal elds) and a coupling between the qubit levels. This is equivalent to a spin under the in uence of a transverse eld, as (j#ih"j j"ih#j ) = i y. We will refer to the beatnote that drives this transition from hereon as \carrier". 3.3 Coupling Spin to Motion So far we have neglected the fact that the ion is con ned in a harmonic well. This will be treated more in depth in 4.2. Using the standard quantum mechanical treatment of the harmonic oscillator, we use the ladder operators to lower and raise the number of motional quanta (\phonons") of the ion with mass m and harmonic vibration frequency ! with the Hamiltonian H = p^ 2 2m + 1 2m! 2x^2 a = r m! 2 h (x^+ i m! p^) (3.6) ay = r m! 2 h (x^ i m! p^) (3.7) 31 where the state representing the number of phonons is designated jni , and the ladder operators give a jni = p n jn 1i (3.8) ay jni = p n+ 1 jn+ 1i (3.9) aya jni = n jni (3.10) and the commutator [a; ay] = 1. Applying two elds with a beat frequency detuned to the red (blue) by the energy of the phonon, we can couple the spin to the ions motion and remove (add) a phonon while ipping the spin, as in gure 3.2. This scheme transfers momentum from the elds to the ion. The mode excited is selected spatially by the beat frequency wavevector k, and spectrally by the beat frequency = !2 !1 = !hf + !mode This treatment assumes that we are in the resolved sideband limit, i.e. the exposure of the ion is long enough and low power enough that the driven transition is not broadened to overlap with the carrier or other modes. We can illuminate this e ect by writing the interaction Hamiltonian for this situation HI = h 2 (e i( k ~r ( !)t ) + + e +i( k ~r ( !)t ) ) (3.11) where is the coupling strength of the qubit states, k is the wavevector di erence of the two beams, is the phase di erence of the beams, ~r is the location 32 ???? ?? ? ?? ?????? ?? ???? ???? ?????? ?????? ?????? ???? Figure 3.2: Driving a Red Side Band (RSB). The beat frequency ! = !2 !1 = !hf !COM imparts a spin-dependent momentum kick that takes the ion from j"; ni to j#; n 1i by absorbing a photon from !1 and emitting a photon into !2. 33 of the ion, and ! = !2 !1 , i.e. the beat frequency. Simplifying this general description to account for only motion along the transverse-x direction, and writing the coordinate as x^ = x0 + qp 2 (a+ ay), and q = q h M! we rewrite 3.11 as HI = h 2 (e i( k(x0+ qp 2 (a+ay)) ( !)t ) + + (e i( k(x0+ qp 2 (a+ay)) ( !)t ) ) (3.12) where k now only refers to the wavevector di erence in the x direction. Assuming the spatial spread of the ion relative to the beat wavelength is small, we make the Lamb-Dicke approximation = kq= p 2 << 1, and keep leading order terms in the phonon operators HI = h 2 (e i( kx0 ( !)t )(1 i (ae i!t + ayei!t)) + +(ei( kx0 ( !)t )(1 + i (ae i!t + ayei!t)) ) (3.13) where we have rotated into the frame of the vibrational modes ! [32]. If we set ! = !, HBSB = h 2 e i( kx0 )ay + + h:c: (3.14) HRSB = h 2 e i( kx0 )a + + h:c: (3.15) The RSBs will allow us to cool our ion chain beyond the Doppler cooling limit by coherently removing phonons. 34 3.4 Multi-Ion Gates Now that we have introduced a way to couple the ion spin to its motion, we can show a way to create gates between ions, using their motion as a bus to transport information between disparate spins [3, 33, 34, 35]. By applying beatnotes symmet- rically detuned from carrier we will virtually excite normal modes, and induce sign dependent Stark shifts on ions, where the phase accrued on each spin will depend on the mode coupled and the ion location in the chain. Generalizing equation 3.13 to multiple ions, while still only coupling to x direction phonons, we rewrite equation 3.13 HI = NX i h i 2 [(e i( kx i 0 ( !)t i)(1 NX m=1 i i;m(ame i!mt + ayme i!mt)) i+ +(ei( kx i 0 ( !)t i)(1 + NX m=1 i i;m(ame i!mt + ayme i!mt)) i )] (3.16) where i is an index summed over N ions, and m is an index summed over N modes (for a linear chain of ions con ned in an harmonic trap, there are N modes per axis [36], as we shall see in 4.2.1. !m is the frequency of the mth x direction mode, and now i;m = kqi;m= p 2, where qi;m = bi;m q h M!m and bi;m is the normalized matrix of ion displacements for ion i due to mode m, as described in 4.2.1. We will be applying two global, symmetrically detuned from carrier beats, so 35 we now sum equation 3.16 over two beatnotes !: HI = h 2 NX i (1 NX m i i;m(ame i!mt + ayme i!mt)) i [e i( k Rxi0 !t R i ) + e i( k Bxi0 !t B i )] i+ + h:c: (3.17) Where kR(B); R(B) refer to the wavevector and phase di erence for the two beams tuned to the red (blue) of carrier. Distributing the evolving coordinate and tem- porarily dropping the terms that go as ei(!m+ !)t to simplify the equations we are left with HI = h 2 NX i X m i;m(ame i(!m !)te i( k Rxi0 R i )+ayme i(!m !)te i( k Bxi0 B i )) i i ++h:c: (3.18) which can be more concisely written as 2 HI = h 2 NX i X m i i;m(ame i(!m !)t i im + ayme i(!m !)t+i im) i is + h:c: (3.19) where im = ( k Rxi0 R i k Bx0;i + Bi )=2 is the phase of the force on the i th spin, is = ( k Rxi0 R i + k Bx0;i Bi )=2 is the spin phase of the i th spin and iS = i +e i iS + i e i iS In our experimental setup the wavevectors of the BSB and RSB are counter oriented, i.e. ~ k R ~ k B , causing the spin phase to be insensitive to beam path length uctuations. As both follow the same path, any uctuation will cause equal and opposite phase change in both, i.e. iS = ( k R + kB)xi0=2 = 0 2The interested reader is welcome to compare this to the special case of 2 ions presented in [32] 36 Although the Hamiltonian above can create displacements in phase space while coupling to the ion spins, we are only interested in driving transitions where motion is only weakly excited and can be ignored. This happens when the beatnotes are su ciently far from the side-bands so to not drive them o -resonantly, i.e. << !m !. This is known as the \slow" M lmer-S renson gate [3, 33]. Applying 1.26 to 3.19, we will get many terms, however, we can observe that there are three types of commutators that could possibly contribute to the e ective Hamiltonian: 1. Same mode, same ion 2. Same mode, di erent ions 3. Di erent modes, di erent ions Assuming we set the spin phase so that we only have spin operators in the x-basis, only the second option contributes to the e ective Hamiltonian. 1. [an i; ayn i] = an iayn i ayn ian i = ( i)2(anayn a y nan) = I So this term does not e ect the e ective Hamiltonian. 2. [an i; ayn j] = an iayn j ayn jan i = i j(anayn a y nan) = i j 3. [an i; aym j] = an iaym j aym jan i = 0 Thus, the resulting e ective Hamiltonian is (here I am including the terms evolving at ! + !m previously omitted) HMS = h 4 X i;j;m ( i j i;m i;m ! + !m + i j i;m i;m ! !m )e i m;ie i m;j i j = (3.20) 37 HMS = h 2 X i;j;m i j i;m i;m!m !2 !2m e i m;ie i m;j i j (3.21) which we can write succinctly as HMS = P i;j Ji;j i j where Ji;j = h 2 X m i j i;m i;m!m !2 !2m e i m;ie i m;j (3.22) is the coupling strength between spin i and spin j. For the duration of this work, the relative modulation phase will be assumed to be equal for all ions, and will only be meaningful when compared to other force generating lasers. This is due to the ions being addressed with global beams, and the modulation frequencies in the MHz regime have phase fronts that do not signi cantly change over the length of the ion chain, that is no longer than 30 m . This assumption will change in chapter 8, where individual phase control for each ion will allow us to control the relative sign of spin-spin couplings. 3.5 Conclusion In this chapter all the necessary components for creating the simulated or e ective components of the Hamiltonians we will investigate numerically and exper- imentally in 6 have been introduced and succinctly explained. When we simulate these Hamiltonians we will be adding in several frequencies to create both spin-spin interactions as in 3.21 and single spin simulated magnetic elds as in 3.5. The hid- den assumption there is that combining the needed modulations of the light will 38 create an e ective Hamiltonian that is a sum of the desired e ective Hamiltonians. This is a valid assumption only if the beatnotes do not create cross terms with a low or stationary frequency in the e ective Hamiltonian. For a more in depth treatment of the possible unwanted e ects of this approach, see [37]. 39 Chapter 4 Experimental Setup 4.1 Introduction The experimental setup for the quantum simulation experiments reported in this thesis is highly complex. In fact, it is so complex that a single grad student starting from an empty lab would likely need more time just to set up part of the experiment, let alone perform any experiment, than an entire PhD program should reasonably require. The setup described here was luckily not started from scratch. The linear ion trap [2] was refurbished from its previous role as a Cadmium ion trap. We also also greatly bene ted from previous experience in the group [1, 38, 39] regarding how to use 171Yb+ . Control circuits for lasers and optical cavities, and the FPGA pulser were designed by authors mentioned in the citations above. 4.2 Ion trap The type of ion trap we use for our experiment is the Paul ion trap [40]. As ions are electrically charged particles, we wish to create an electric potential with a stable minimum to which the ion will relax. Unfortunately, Earnshaw?s theorem [41] informs us that a static electric potential cannot have a stable minimum. Paul?s trap circumvents this problem by combining a static electric eld and an oscillating 40 ???/2 ???/2 ???/2 ???/2 ??? x y z Figure 4.1: Paul trap with ions. the distance between hyperbolic electrodes is 2r0. eld to create a time-averaged pseudo-potential that has a stable minimum. The simplest example of such a setup is the four rod Paul trap, where the rods have a hyperbolic cross-section, as shown in gure 4.1. In order to create a con ning potential, we apply the voltage to the electrodes 0 = U + V cos(!t) (4.1) where U is a DC voltage and V is the amplitude of an RF voltage oscillating at frequency !. r0 is the distance from the electrode to the center of the trap, as shown 41 in gure 4.1. The resultant potential is = 0 2r20 (x2 y2) (4.2) After taking the gradient to nd the electric eld, we can write the equations of motion (known as the Matheiu equations): dx2 d 2 + (a+ 2qcos(2 ))x = 0 (4.3) dy2 d 2 (a+ 2qcos(2 ))x = 0 (4.4) where a = 4eU Mr20!2 ; q = 2eV Mr20!2 ; = !t 2 (4.5) When the a and q parameters are chosen so that neither coordinate grows exponen- tially rather than oscillate, the ions are stably con ned. The missing element of the con nement is in the z-direction, the trap axis (I will use this convention from here on out). That can be corrected for by simply adding a repulsive DC electrode to each end of the trap axis. 42 4.2.1 Normal Modes A linear chain of N ions, trapped in a harmonic trap, experience a pseudo- potential due to an AC eld, a static eld and their electric charge[36] V = M 2 NX n=1 3X i=1 !2i x 2 ni + e2 8 0 NX n;m=1 m6=n " 3X i=1 (xni xmi) 2 # 1=2 (4.6) from which we can nd their equilibrium positions in terms of a rescaled co- ordinate um = x0m=l, where x 0 m is the small displacement from equilibrium de ned by xni (t) = xni + x0mi, and l = e2 4 0M!23 1=3 (4.7) where e is the charge of the ion, M is the mass of a single ion, and !3 is the longitudinal collective center of mass frequency - the lowest normal mode frequency. We can now consider the Lagrangian of the system, L = T V = M 2 NX n=1 3X i=1 _x0ni 2 M 2 NX n=1 3X i=1 !2i xni + x 0 mi 2 e2 8 0 NX n;m=1 m6=n ( 3X i=1 xni + x 0 ni xmi x 0 mi 2 ) 1=2 (4.8) And expanding to lowest order, neglecting couplings between di erent spatial 43 directions1 L M 2 ( NX n=1 _x0ni 2 !23 NX m;n=1 Amnxm3xn3 + 2X i=1 " NX n=1 _x0ni 2 !23 NX m;n=1 Bmnxm3xn3 #) (4.9) and Amn = 8 >>>>>>>>< >>>>>>>>: 1 + 2 P p=1 p 6=m 1 jum upj 3 if m = n 2 jum unj 3 if m 6= n (4.10) Bmn = 1 2 + 1 2 mn 1 2 Amn (4.11) where A and B are the matrices for the longitudinal and transverse eigenvalue equations. We de ne the normal modes bm that describe the oscillatory displace- ment of ion i from its equilibrium with longitudinal (transverse) frequency !z p m (!z p m) NX n=1 Amnbi;n = mbi;m (i = 1; : : : ; N) (4.12) NX n=1 Bmnbi;n = mbi;m (i = 1; : : : ; N) (4.13) and assumed the collective transverse mode frequencies are equal. We de ne , the trap anisotropy !x = !z (4.14) 1This is a valid assumption as long as the trap anisotropy parameter is small enough, i.e. the chain is not close to buckling from a linear chain into a zig-zag con guration. 44 where !x is the COM frequency for the x-direction transverse mode and !z is the COM frequency along the ion chain axis. The relationship between the axial and transverse normal-mode eigenvalues is p = 1= 2 + 1=2 p=2 (4.15) Transcribing this analysis into a Wolfram Mathematica script generates the ion equilibrium positions: k = p 2 2 369:5 10 9 ; NumofIons = 10; !cm = 2 4:863232 106; !tilt = 2 4:81294 106; = q 1 !tilt 2 !cm2 ; !z = !cm; l = qe2 4 0MYb171!z2 1=3; Table[uH[m]; fm; NumofIonsg]; (* create position variables *) Table h fH[m] = uH[m] Pm 1 n=1 1 (uH[m] uH[n])^2 + PNumofIons n=m+1 1 (uH[m] uH[n])^2 ;uH fm; NumofIonsg]; EqH = Table[fH[m] == 0; fm; NumofIonsg]; (* generate coupled equations *) IniH = Table[fuH[m];m=10g; fm; NumofIonsg]; (* generate initial conditions *)condition solH = FindRoot[EqH; IniH];FindRFindR uHa = Table[uH[m]/.solH[[m]]; fm; NumofIonsg]; Print[uHa l 10^6; \ position in microns of ions"]ositioositio And we get the ion distance on the z-axis from the center of the trap in m for the trap parameters used in chapter 7: f-10.0021,-7.31662,-5.05321,-2.97461,-0.982864,0.982864,2.97461,5.05321,7.31662,10.0021g 45 With these distances we can calculate the transverse frequencies for the x- direction, which we will be using to create our spin-spin interactions: TMatrix = Table h If h i 6= j;TransMat[i; j] = 1Abs[uHa[[i]] uHa[[j]]]3 ;6 6 TransMat[i; j] = (1= )2 Pi 1 j=1 1 Abs[uHa[[i]] uHa[[j]]]3 PNumofIons j=i+1 1 Abs[uHa[[i]] uHa[[j]]]3 i ; fi; 1; NumofIonsg; fj; 1; NumofIonsg]; TransEigVal = Eigenvalues[TMatrix];alues[TMatrix] TransNormalFreq = Sqrt[TransEigVal]; Print [TransNormalFreq !z/ 2/ =106; \ Trap frequencies in MHz"] f4.86323,4.81294,4.74058,4.64815,4.53638,4.40509,4.25343,4.07989,3.88217,3.65694g Trap frequencies in MHz The transverse frequencies displayed here run from the highest, the COM mode, to the lowest - the \zig-zag" mode, as shown in gure 4.2. The second highest mode is the \tilt" mode, and the spacing between COM and tilt, unlike any other two modes, is xed at !2 = !1 p 1 2 (from here on the normal mode frequencies are referred to as !n, running from high to low). When generating the spin-spin coupling matrix J we will have exibility in control of the form of the J by setting our detuning close to the mode that has the form similar to our desired couplings, as can be seen in gure 4.3. Using a single detuning will not give us complete control, however the modes form a complete basis, and we will use this in 8 to gain complete control over J. 46 Figure 4.2: Transverse mode components for 10 ions. The arrows represent the maximum displacement of the ions from equilibrium while oscillating at mode frequency. The modes are ordered from highest frequency (COM) to lowest (zig-zag). 47 -1500 -1000 -500 0 -3 -2 -1 0 1 2 3 DetuningHKHzL J i, jHKH zL J89,10< J88,10< J88,9< J87,10< J87,9< J87,8< J86,10< J86,9< J86,8< J86,7< J85,10< J85,9< J85,8< J85,7< J85,6< J84,10< J84,9< J84,8< J84,7< J84,6< J84,5< J83,10< J83,9< J83,8< J83,7< J83,6< J83,5< J83,4< J82,10< J82,9< J82,8< J82,7< J82,6< J82,5< J82,4< J82,3< J81,10< J81,9< J81,8< J81,7< J81,6< J81,5< J81,4< J81,3< J81,2< Figure 4.3: The mess of couplings for 10 ions. The x-axis is detuning in KHz from the COM. As can be seen in the plot, near the COM all the couplings are nearly equal, as can be expected from the form of b1. Near each mode the couplings diverge to in nity, however this behavior is non-physical as we must maintain the weak excitation of phonons condition. 48 4.3 Micromotion In an actual trap, the alignment of the DC saddle point and the RF node will be imperfect. This will alter the form of the equations of motion adding an extra term of excess micromotion - oscillatory motion around the psuedopotential minimum, in addition to the inherent micromotion seen in 4.4. This unwanted motion creates an oscillating Doppler shift on the ion?s reso- nance frequencies, reducing scatter rates for cooling and detection and broadening the transitions. We have found that ions are less stable in the trap when this condi- tion exists. To correct this condition, the DC voltages must be altered as changing the RF node is di cult (or impossible, as in our setup). There are several indica- tors useful for reducing the excess micromotion[42]. One is measuring the Doppler cooling scattering linewidth. The second is measuring a correlation between the trap frequency phase and the scatter rate using a Time to Digital Converter. For micromotion along an axis parallel to an imaging device, one can also reduce V to see a displacement of the ion from DC saddle-point. In our trap we have found that it was very di cult to reduce micromotion, especially in the transverse direc- tion not coupled to our gates (y), as reducing micromotion would rotate the trap axis, creating unwanted components of the Raman beams along the y axis. This would require us to change the incident angle of our Raman beams on the ion chain, which is quite di cult. Instead, we would reduce micromotion by maximizing the Rabi rate. It is likely that reducing the micromotion along the x-axis leads to more e cient Doppler cooling on that axis, leading to a higher Rabi frequency. 49 4.3.1 The Linear Trap The trap used in the work in this thesis is of a di erent design than the one discussed above. It is constructed of three layers of thin alumina, with electrodes segmented by laser machining of the alumina and formed by gold coating, as shown in 4.4. (a) (b) 250 ?m dc dc dc dc rf rf 375 ?m 200 ?m Figure 4.4: (a)Linear trap schematic with ions in trapping region. The RF electrode is a single electrode with a slit of equal dimensions to the DC electrodes central slit. The top and bottom DC layers are identical. (b) Linear trap photograph. Here the capacitors and resistors in a pi- lter con guration on the DC electrodes can be seen. These prevent capacitive RF signal coupling to the DC electrodes. Although this con guration is quite di erent from that of the four rod con g- uration, the analysis is quite similar. As there is no easily found analytical solution for the potential generated by these boundary conditions, the approach we take is modeling the potential numerically with a commercial electromagnetic modeler - 50 Charged Particle Optics. For each electrode, we set all other electrodes to ground, and set said electrode to some voltage. Thus the linear combination of scaled po- tential maps gives us the total potential for the trap. Using this numerical solution, a ponderomotive psuedo-potential can be calculated = e2 4m!2 rV (x; y; z)2 (4.16) and treating the linear trap as we would treat the hyperbolic Paul trap - a eld exerting a harmonic restoring force on the ion, we nd the secular frequency (the axial COM frequency) [43, 44, 45] !2z = e2 4m2!2 @2 @2z (jrV (x; y; z)j)2 (4.17) Then the secular frequency of the trap is !z = e2V 2 4 p 2m!2r4 @2 @2z r(jV (x; y; z)j)2 (4.18) which is the identical to the secular frequency of the hyperbolic trap, but with a rede ned e ective distance from the electrode to the ion r and a geometric factor . This factor is the ratio between the quadrapole portion of the linear trap (or any other trap) to the oscillating term of the hyperbolic trap. 51 4.3.2 Trap Voltages and Control As shown in 4.4(a), the trap has six electrodes on both of the DC layers. Wiring all electrodes would give more control than necessary to control the trap axis and minimize micromotion, and would increase signi cantly the di culty of assembly and wiring. Therefore, only 3 DC electrodes in each layer are wired to external leads. The remaining three are wired to trap ground. These electrodes are wired to spot welded on board pi- lters. The leads running from these lters then connect to an external breakout box, which itself has additional pi- lters for each electrode, as discussed in 8.6. The trap voltages are supplied by an Iseg EHS-80-05XK3 high voltage module. This module is powered and controlled by an Iseg MPOD mini- crate. The crate provides a network based control of the voltages. The module has eight channels, all with SHV connectors. These are not to be confused with BNC connectors, which are very similar in appearance. We constructed custom cables to connect between the SHV connectors on the ISEG box to the BNC connectors on the breakout box. It is important to note that the channels have a single polarity (dual polarity requires a oating ground that reduces stability) Channels 0-5: up to +500V /15mA Channels 6-7: down to -500V The control program will automatically switch the driven channels from 4 and/or 5 to 6 and/or 7 if the set voltage on electrodes DC 3 and/or DC 4 is negative. 52 Figure 4.5: Trap control program. The upper right corner provided settings for a timed relaxation and tightening of the trap RF, via the LabJack controlled VCA. Upper right side controls the DC trap voltages. Lower half indicates sensed DC voltages and currents for tab displayed, as well as controls power to individual channels and the Iseg chasis (Main Power). 53 It is assumed the user has switched the cables on the breakout box if this is the case. 4.3.3 Oscillating Trap Voltage The RF voltage V is provided via a helical RF resonator [46]. This resonator has a quality factor Q of 300 and is resonant at 38.86 MHz . As the power deliv- ered through the resonator is changed, the ohmic losses change and the equilibrium temperature of the resonator and perhaps the trap itself changes. This leads to geometric changes that change the resonant frequency of the resonator and the RF voltage of the trap. As this leads to a change in the secular frequency of the ion, we have implemented a stabilization scheme. The rst stage of this stabilization scheme is thermal stabilization of the RF resonator. This is accomplished by a Thorlabs TC200 Heater controller. This heater senses the temperature of the resonator using a thermistor, and stabilizes with a resistive coil. The stabilization is improved by in- sulating the resonator and the thermal feedback system with an adhesive foam with a re ective aluminum layer. We nd that this temperature is stable to less than a degree when the set point temperature is 250C, slightly above room temperature. The second stage of this stabilization scheme is active frequency stabilization of the HP-8640B driving the trap RF to resonance. This helps stabilize the delivered power. It is helpful, but not su cient, to stabilize the trap transverse modes. We pass the driving RF through a Minicircuits BDC ZFBDC-62HPS bidirec- tional coupler, and mix the re ected signal with a phase shifted reference signal 54 Figure 4.6: Trap control program electrodes tab. Here the sensed voltages on each electrode are displayed. The electrodes are dis- played without the RF layer, facing towards the imaging optics. The electrodes labeled ?GND? are grounded. from the HP-8640B (this signal is impedance matched by passing through a 50 MHz low-pass lter) frequency generator providing the driving signal (which is ampli ed to 26dBm). The phase is adjusted using a potentiometer providing a set-point voltage to a Minicircuits JSPHS-51+ narrow band phase shifter (this component must be selected carefully for the resonant driving frequency of the trap). This mixed signal is low pass ltered so only the DC component remains, and is fed back as a DC frequency shift to the driving 8640B (typical modulation peak of 320KHz). 55 4.4 Optical Setup 4.4.1 Ionization There are three lasers used for excitation of the neutral Yb, ionization, cooling, optical pumping, and detecting. These are all DL-100 Toptica lasers - temperature and current controlled laser diodes. Aligning the diode with a re ection from an external optical grating allows for mostly single mode selectivity for the modes selected by the diode cavity. In general, once the grating is aligned to generate a frequency in the neighborhood of the desired frequency (for the minimal lasing threshold current) frequency adjustment should be done by adjusting the grating angle and diode current. If mode hopping occurs, then one may try a sensitive adjustment of the diode temperature. As discussed in 2.1, it is necessary to excite the neutral 171Yb before it may be ionized by a 369 nm or 355 nm beam. As the beam exiting the laser has an elliptical pro le, we rst circularize the beam with a prism pair. Next we protect the laser from back re ections from any of the downstream optical elements, as these create unwanted optical cavities with the laser and unnecessary mode hopping (this is true for all the lasers in this experiment except the Vanguard). A mechanical shutter controlled by the trap control program exposes or conceals the trap to the beam, which after passing through the nal lens is roughly 50 m vertically and 100 m horizontally, at roughly 2mW . When the 171Yb+ oven is driven at 2.3A after warming up for about a minute, the light intensity is powerful enough to load an ion roughly every 10s when using the Raman beams for ionization of the excited 56 Yb. This is achieved with good isotopic selectivity: the 171Yb oven is roughly 90% pure, but loading unwanted isotopes occurs less frequently than 10%. As the oven is perpendicular to the excitation beam, the atoms do not experience a doppler shift with a thermal distribution. The desired frequency of the excitation beam is 751.527640 MHz , however we do not lock this laser, rather we tune it to the desired frequency and load before it drifts too far o resonance. The rate of frequency drift depends on the temperature of the room and how long the laser has been operating at given temperature and current settings. The laser takes roughly an hour to equilibrate. 4.4.2 935 nm Repump The 935 nm laser providing the repump light discussed in 2 is provided by a DL-100 as well. As this light is needed to repump both the F=1 and F=2 manifolds of the 3D3=2 state, the 320.56922THz carrier is modulated by an EOSpace ber EOM at 3.07 GHz . This laser is only stabilized for long timescales, by a feedback signal to its grating angle provided by a software PID. This PID generates a feedback signal based on the frequency di erence between the laser wavelength, measured by a HighFinesse wavemeter via optical ber and a software controlled set point. The wavemeter accuracy may drift by up to 20 MHz , depending on room temperature, and is calibrated to our iodine spectroscopy setup, discussed below. Its precision is roughly a MHz . 57 4.4.3 638 nm Repump As discussed in 2, repumping the ion from the long lived but unlikely 2F7=2 state requires two close frequencies of 638.61 nm light. As this light is not often required, and when it is required the ion chain has likely decrystalized (from a collision) and data taking cannot immediately continue, the laser is only stabilized by the HighFinesse wavemeter ( 5 MHz short time scale stability). To provide both wavelengths of 638.6103 nm and 638.6151 nm , the PID feedback to the laser grating setpoint is alternated between the two desired wavelengths, once every minute. DL- 100 laser is rst passed through a prism pair to circularize the beam pro le, and then coupled into a ber optic. The output of the ber is directed at the ion chain with several mW and a large waist at the ion chain, so careful alignment is not required. 4.4.4 Cooling, Pumping, and Detecting with 369 nm To generate the 369 nm light we require, we use a MBR-110 from Coherent. This is a tunable Continuous Wave (CW) Ti:Sapphire laser, pumped by a Coher- ent 18W 532 nm Verdi-V18. It outputs roughly 1.5W at our desired wavelength of 739.052526 nm . The MBR-110 is a made of a single block of aluminum, and is thus passively stable. The bow tie con guration passes through an optical diode that prevents backward re ections. The wavelength of the laser is selected by controlling the gain pro le for multiple laser mode selecting elements. These include the laser cavity length, an etalon (that oscillates around a rotation angle at 80 KHz to gener- 58 Ti:Sapphire crystal Reference Cavity 532nm Pump Light Etalon Verdi Figure 4.7: The MBR-110 laser. 59 ate an error signal used to lock its angle to the laser cavity length), and a birefringent lter (with a wavelength dependent response). The Free Spectral Range (FSR) of a bow-tie cavity is FSR = CL where L is the length of the cavity. As the largest cavity is the laser itself, the mode-hop free frequency range tuning and locking is controlled by laser cavity length ( C1M = 300 MHz ). The laser has a linewidth of 75 KHz . The 739 nm output power of the MBR-110 is ber coupled into a high power ber, except 100mW that is ber coupled to the wavemeter, and 40mW to an EOSpace ber EOM, used for the iodine lock, described in section 4.4.5. The high power ber delivers 800mW to a WaveTrain frequency doubler, as shown in gure 4.8. After exiting the doubler the beam is 430MHz red detuned from resonance. It is divided in three, by a piece of glass. The front surface re ection is modulated by an New Focus EOM driven at 2.105GHz to provide the optical pumping, and is shuttered by an AOM that up shifts the beam by 424 MHz , into resonance with the transitions discussed in 2.2. This beam is then combined with the detection beam before being ber coupled. The back surface re ection is modulated by an EOM driven at 7.374GHz. It is then passed through a Brimrose AOM. The rst order di racted beam is up shifted by 400MHz and used for cooling (30 MHz detuned from resonance). The zeroth order is used as a \protection beam". This powerful beam is used to recrystallize the ion chain when a decrystallization is detected by the cooling light scatter PMT. As it is modulated for cooling, it is resonant with both the 2S1=2 jF = 1i and 2S1=2 jF = 0i to 2P1=2 transition, and we have found this to improve the speed and probability of recovering all the ions. The transmission 60 through the glass is used as the \protection" beam, which is shuttered by an Intra- Action AOM and further red detuned by 200 MHz . This beam is more powerful than the three other beams, but does not cool both transitions as the previous beam. All three beams are coupled into optical bers to reduce unwanted scatter light in the direction of the rst order beam. This scatter was found to limit coherence time in long experiments. The added bene t is clean laser modes and ease of aligning and overlapping the beams. A small portion of the transmission is siphoned o by a beam splitter, to be shuttered and shifted into resonance by a Brimrose AOM. The \protection+" beam was combined with the 399 nm beam in order to simplify alignment. Unfortunately, we have found that overlap is di cult and the 399 nm beam alignment drifts wildly from day to day due the need to change the laser grating angle and the long beam path to the chamber. Therefore, alignment with the cooling or protection beam at the ion position does not allow perfect overlap along the entire beam path. Be- fore entering the chamber window, the detection, cooling and protection beams are horizontally stretched by cylindrical lenses, and then focused at the ion chain to a horizontal beam waist of 100 m . However, as shown in gure 4.9, the beams enter the chamber at 450 to the ion chain, giving an e ective beam width at the chain of 150 m . This gives a roughly equal detection beam pro le across our ion chain, that for 16 ions is 26 m for our current operating parameters for the ion trap. As our beam pro les are roughly Gaussian, the beam dimensions and power are limiting factors on the detection homogeneity. As the power in the beam is limited, 61 handling longer ion chains with optimal exposure and detection beam powers for all ions requires a larger horizontal beam pro les. In order to achieve this with a xed laser power, we are required to squeeze the vertical beam waist. This leads to more stringent requirements on beam alignment, and more noise on the e ective laser power at ions, due to beam steering by air turbulence. Currently, with our detection beam vertical waist of 10 m this is not a limiting factor on our detection error, rather we su er from power uctuations of 5%. In order to remedy this issue we are currently investigating the use of a feedback stabilization system (\noise eater"). In fact, it is questionable if we can widen the beam much more without increasing background scatter from the trap o the imaging window into the imaging optics. A possible remedy is to create a beam with a \top hat" pro le using di ractive optics, and thus avoid wasted beam power in the wings of the beam and ensure equal detection conditions for all the ions. Note: for some of the earlier results reported in 6, we used the MBR-110 to drive the Raman transition, and a Toptica TA-100 ampli ed diode laser to drive all resonant transitions. In that setup, we only generated 200 mW of 739 nm light and su ered as a result for larger numbers of ions. The TA-100 was locked using the iodine spectroscopy setup described in 4.4.5. The major di erence is the use of a home built confocal cavity, that was locked by length to the transition, and the use of a Pound-Drever-Hall feedback signal to lock the laser frequency to the cavity length. This feedback signal was separated into a low (< 100KHz) component that controls the laser grating angle, and a high frequency component (> 10MHz) for the laser diode current, controlled via a Bias-T [47]. 62 MBR Wave Meter Monitor cavity protection 7.374GHz Protection+ Chamber Verdi - V18 532nm Iodine lock 2.105GHz X2 LBO 39 9 n m To Chamber Figure 4.8: A schematic diagram of our 369 nm and (399 nm excita- tion for neutral Yb) optical setup. The nal cylindrical lens, mounted on a micrometer positioning stage, is a vertical lens with focus 150 mm. The micrometer is used for aligning the vertical direction of the detection beam, which is the most sensitive. Chamber is shown in g 4.9 63 Figure 4.9: A schematic diagram of chamber and nal stage of res- onant beams. The cylindrical lens the beam pass through focuses the vertical direction of the beams, with an 80 mm focal length. It demagni es the intermediate focus by 5. The dashed line represents the rough trajectory of the high purity 90%171Yb oven. As it is perpendicular to the excitaiton beam, our ability to selectivly load the desired isotope (rather than the impurities) is higher than the purity of the oven. Although the beams exit through the front view port, from which the ions? state detection ourescence is collected, the 638 nm , 935 nm and roomlight do not create unwanted signals on the ICCD or PMT as they are ltered. As the waists of these beams is large at the ions, alignment requirements are lax. 64 4.4.5 Iodine Spectroscopy Lock As the linewidth of the S-P transition is 20 MHz , we must stabilize the laser on long time scales to a much narrower center frequency to have consistent detection and cooling. This is accomplished using a spectroscopic feature of iodine narrower than our linewidth, to which an external cavity (part of the MBR-110 laser) is stabilized. An optical ber carrying 40 mW of the MBR-110 output is coupled into a ber EOM, modulated at 13.315 GHz , so that the rst order sideband of this modulation ( 4mW) excites the iodine transition of interest[47, 1]. Three beams are derived from the ber - a reference beam directed at the Nirvana Auto-Balanced Photoreceiver, a probe beam which is sent through the iodine cell and out into the signal photodiode of the Nirvana, and a pump beam transmitted through an 80 MHz Neos AOM into the iodine cell, overlapping with the probe. The iodine cell is insulated with berglass and aluminum foil, and heated to 5000c. This excites the higher energy rovibrational states of the iodine molecules and strengthen the absorption signal. A cold nger maintains the pressure in the cell and prevents pressure broadening of the transitions [48]. The Nirvana detector is designed to subtract the reference from the signal, thus making it a sensitive detector able to remove 50dB of noise. We have encountered two major problems with this setup: Background light from the pump beam can leak onto the signal photodiode. As this background light level can drift, the detector fails to subtract the noise. The Nirvana detector is designed to balance the attenuation of the photodi- 65 odes. However, we have witnessed that it is limited in this ability. As the reference beam is much more intense than the probe beam, it must be atten- uated for the detector to work properly. The 80 MHz modulation of the pump beam is itself modulated at 10 KHz , enabling the use of a lock-in-ampli er to demodulate and amplify the Nirvana error-signal (the derivative of the Doppler free hyper ne transitions in iodine). As there are two strong transitions for the above EOM frequency near to the resonant laser frequency (for the 739 nm , non-doubled light), there will be three erro-signal peaks. We stabilize the MBR-110 to the peak corresponding to the highest frequency of the three at 405.644321 THz , which is 20 MHz above the nearest feature. This can lead to confusion, as the wavemeter displaying the laser frequency drifts in the same range. However, the incorect frequency is close enough to the desired one that ions will in fact be visible on the ICCD if this occurs - but they will appear dim. We route the error-signal from the lock-in-ampli er to a home built PID. The PID output is then routed to the MBR-110 External Lock Input. I have connected an attenuator to this input, as it cannot take an input larger than 10V. This signal is fed to the piezo controlling the length of the MBR-110 external reference cavity (an internal component of the laser, as shown in gure 4.7). The peizo mounted tweeter mirror of the laser locks the length of the laser cavity to the external cavity length. This lock can be stable on long time scales to 2 MHz , which is su cient for the linewidths at hand. 66 Iodine Cell Nirvana Pump Reference P robe Monitor cavit y Figure 4.10: Schematic of iodine setup. The angle between the beams in the cell is exaggerated. 67 4.4.6 Vanguard 355 nm Laser The Vanguard Diode Pumped Solid State Laser is intended to be a turn key laser. The lasing medium is a Nd:YVO4 crystal (Neodymium Doped Yttrium Or- thovanadate) pumped by a solid state diode lasing 12W of 808 nm through a 19 core ber bundle. The crystal (\oscillator") should then lase roughly 5W at 1064 nm . Unfortunately, we are currently overdriving the oscillator to 6W, for reasons to be explained below. The output of the oscillator is then ampli ed by another Nd:YVO4 crystal, pumped by two 20W 808 nm diodes, to roughly 20W of 1064 nm laser light. This light impinges on a Second Harmonic Generator (SHG) that produces an 8W beam of 532 nm . A semiconductor saturable absorber re ector mode-locks the laser at 80 MHz , where the pulses have a 10 ps duration [49]. This mirror has a re ection that decreases with increasing light intensity, thereby allowing the creation of laser pulses at high gain. The 1064 nm and 532 nm beams are mixed on a Third Harmonic Generator (THG) to produce the desired 355 nm light. Although Spectra-Physics is not willing to disclose this information, the SHG and THG are likely type 2 Lithium triborate crystals [50, 49]. The output of the laser according to its speci cations should be 4W, as in gure 4.11. Unfortunately, the power we observe is lower and it is dropping, as the laser is old and bought second-hand. The laser is controlled by a desktop computer via serial connection, using the control program MLUV. After initial warm-up ( 30 min), the power of the laser and its stability may be optimized. As the components of the laser are old, the built in optimization function of MLUV 68 Figure 4.11: MLUV main control window. The laser has two operat- ing modes: one where the laser power is stabilizaed and one where the diode currents are stabilized. In order to run in power stabilization mode there must be some disposable power. As our power is low, we typically run in current stabilization mode. Although we have not found the power to be less stable in this mode on the short time scale, the overall power is slowly dropping. does not work. The initial optimization should focus on the diode currents and temperatures. Lowering diode currents can surprisingly increase power. Usually the most sensitive parameter is the current to \Diode 1", the oscillator pump, as shown in 4.12. These settings can only be changed when the laser is operating in Current Mode. If this procedure fails to raise the power back to its expected level, the next set of parameters to tweak are the ampli er pump diodes currents and temperatures, as 69 Figure 4.12: MLUV oscillator pump current control. shown in 4.13. If these parameters yet again fail to raise the laser power, then it may be attempted to change the oven temperatures of the SHG and THG, as shown in gure 4.14. This is a slow process, as each change must be given time to equilibrate. Each change on the ne knob (roughly in steps of ve units) should be given at least 2 minutes to asses the e ect. If this attempt fails as well, then there is a possibility that the THG and/or the SHG have degraded at the current laser path. It may be necessary to move the THG or SHG. This may be done via MLUV for the SHG and only for the x-direction for the THG, as shown in 4.15. If all the X positions on the THG for a Y position have been degraded, then the Y position may be manually changed via two screws that lock into 6 di erent orientations, via a side panel on the Vanguard. Both screws must be set to the same position, otherwise vertical beam 70 Figure 4.13: MLUV ampli er pump current control. In this gure the individual current controls are grayed out as the mode has not been set to individual diode control. Figure 4.14: Control window for THG and SHG oven temperatures. The ne tunning settings should be used rst. 71 Figure 4.15: MLUV position controls for the SHG and THG. The THG Y-position is only for record keeping - MLUV does not control it. steering will result. Although the methods above have worked in the past to raise the laser power, the power has been dropping constantly and can not be recovered (the experiments in chapters 6 and 7 have been performed with powers ranging from 3.7W to 3.4W). This is likely due to the degrading of the ampli er pump diodes. 4.4.7 Driving Raman and M lmer-S renson Transitions with the Van- guard In 3 we discussed how spin-spin couplings can be created using global beams modulated to drive virtual transitions in the phonon subspace while driving real spin transitions. In the work reported here, this theoretical treatment is realized with the Vanguard DPSS. Although this is a pulsed laser, we will be operating in the weak 72 pulse regime - each pulse pair radiating the ion will contribute a small change to the ion state (also referred to as the weak pulse regime). Thus we will be able to treat the pulse trains as CW lasers [51]. In order to generate the desired beat frequencies, we will interfere the pulses on the ions, absorbing a photon from one frequency component of the pulse train and emitting into another. The beam exiting the laser will be split into two arms, and both arms will be modulated by AOMS. This will give us control of the beat frequency di erence, as schematically represented in gure 4.16. For this scheme, the absolute frequency o set and the absolute phase of the pulse train of each pulse train will have no e ect[51]. We can choose to subtract or add the frequency di erence between AOMs. The beat frequency of interest between teeth in the two o set combs is (80:6(n + 157) + AOM2) (80:6n + AOM1) = 80:6 157 + AOM = 12654 MHz 12 = 12642 MHz , which drives the carrier. A prerequisite for this scheme is an acceptable and stable laser pulse repetition frequency. Our qubit splitting is xed at 12.6428 GHz . As shown in gure 4.17, the 532 nm output of the Vanguard (attenuators not shown) is measured by a Electro- Optics ET-4000 fast photodiode. The output of this photodiode is band pass ltered by a tunable mechanical microwave lter so that only the frequency component close to the qubit splitting is transmitted. This signal is ampli ed and mixed with a microwave frequency source, allowing us to feed forward a frequency o set to AOM1, thus stabilizing the frequency di erence between the Raman arms. An acceptable frequency must satisfy two conditions: 1. It may not drive the carrier transition with just one arm, nor the Zeeman 73 f f frep n n+157 fc+AOM1+? fc+AOM2 fc=C/355nm=844.5THz fBW fBW~1/10ps~100THz fw= frep /N Figure 4.16: Two pulse trains interfering at the ions represented in frequency space. The two AOMs shift the frequency combs, and allow us to control the beat frequency. AOM1 modulates at 225 MHz + , where is a fed forward frequency o set to stabilize the beat fre- quency. AOM2 modulates at 213 MHz when we drive the carrier transition (and a more complex waveform with multiple frequencies for the simulation). The frequency of di erence of 13 MHz is sub- tracted from the beat of of the teeth. The pulse width in frequency space is 1Nfrep , where N is the number of pulses involved in a interac- tion. 74 PID Vanguard AWG Mach - Zender ?/4 ?/4 ?/2 ?/2 12.6GHz f rep ? f AOM1 +? f AOM1 AOM1 AOM2 a c c d d objective v v h1 h2 b X X Figure 4.17: Vanguard with lock and Raman beam setup. Raman arm 1 is modulated by AOM 1, which is driven by an HP8640B with a frequency o set controlled by an error signal from the frequency dif- ference of the microwave reference and the measured repetition rate. Raman 2 is the 0th order beam of AOM 1, and is modulated by AOM 2, with an RF signal created by the AWG. Lenses c(200mm) and d (91mm) image and demagnify the intermediate focus (X) of the verti- cal(lens v,50mm) and horizontal lenses h1(500mm) and h2(150mm). Lens a(200mm) is intended to slowly focus the beam at AOM1 so that is still mostly focused at AOM2. Both Brimrose AOMs func- tion best when the beam is focused in the crystal. Lens b(300mm) recollimates the beam. 75 carrier transition (mf = 0 ! mf = 1) at roughly 7 MHz from carrier (de- pending on magnetic eld strength). 2. It may not drive the M lmer-S renson gate with both Raman beams on (but without the requisite modulations) due to a beat note between the arms that is equally detuned to the blue and red from the carrier. As the pulse width is 10 ps , the bandwidth of the laser is 100 GHz . With a repetition rate of 80 MHz , that indicates 1180 comb lines in the frequency domain. Enforcing these parameters and above conditions, with the more strin- gent requirement that the rate not lie between carrier and Zeeman transition, the following Mathematica code reports allowed repetition frequencies : ranges = Table [frep; fn; 1; 1180g] ; For[n = 1; n 1180; n++; ranges[[n]] = Reduce [ffrep n=2 < (12642:8199 + 7)&&frep n=2Reduc > 12642:8199 7&&frep < 81&&frep > 79g]] ranges = DeleteCases[ranges;False] And the the frequency ranges of the repetition rate frep that satisfy the con- ditions for the possible values between 79 and 81 MHz are reduced to: f80:9988 < frep < 81:; 80:7401 < frep < 80:8295; 80:4829 < frep < 80:5721; 80:2274 < frep < 80:3163; 79:9735 < frep < 80:0622; 79:7213 < frep < 79:8096; 79:4706 < frep < 79:5586; 79:2214 < frep < 79:3092; 79: < frep < 79:0614g Our Vanguard has a measured rep rate of 80.6 MHz , which seems to be just outside one of the permissible ranges (there is no way to control this for this laser, as it depends on laser cavity length that cannot be adjusted). The beat frequency 76 between two comb teeth separated by 157 teeth is closest to the qubit frequency: frep 157 = 80:6 157 = 12654:2 MHz . Our fastest Rabi transitions driven by the Vanguard have been measured to be 1 MHz , which implies the number of pulses needed per Rabi cycle is at least 80. Therefore the width of a comb tooth is no more than 1 MHz when we drive Rabi transitions at full power, and 1 KHz for M lmer- S renson transitions at our typical transition strength. Another consequence of the 10 ps pulse width is that pulse is 3mm long in space. As the pulses must overlap at the ion at the same time, we adjust the Raman2 arm using a beam cube, named in 4.17 as Mach-Zender (the setup is similar to a Mach-Zender interferometer as the ion measures the phase di erence between the arms). As we are utilizing long chains of ions, aligned along the z-axis of the trap (as illustrated in gure 4.18). We are also using global beams to address the entire ion chain at once. Therefore, the most power e cient beam shape is a sheet of light. To approximate this, we use two sets of cylindrical lenses, one vertical and one horizon- tal for each Raman arm. This allows us to compress the vertical beam waist down to 7 m and the horizontal beam waist down to 70 m (Raman2) and 200 m (Raman1) at the ion. In order to measure these beam pro les, we position a mirror in the beam path before the ion, and image the beam at the estimated distance to the ion with a Guppy CCD, as in gure 4.19. Alternatively, one may place the Guppy at the intermediate focus (marked with an X in gure4.17), and estimate the beam waists at the ion based on the demagni cation of the telescope (roughly 1/4). It is important to note that the telescope for Raman2 images AOM2. As AOM2 is modulated by multiple frequencies, the de ection point in the AOM must 77 be imaged so as to refocus all frequencies at the ion chain (the AOM de ects with an angle sin 1(m 2 ), where m is the order of the de ected beam, the wavelength of light, and the sound wave wavelength). This optical setup is far from ideal. The 0th order beam of a driven AOM1 has a poor beam pro le, and the power in the beam depends on how strongly AOM1 is driven, thus coupling the beam strengths of both arms. Also, the beam pro le seems to be coupled to the imaging plane, and we were unsuccessful at changing the beam pro le while maintaining the imaging plane at AOM2. This imaging has also been diagnosed by us to be imperfect, possibly leading to a Stark shift across the chain, that appears as a weak biasing axial eld in quantum simulations. An additional issue in this setup is that Raman2 just barely grazes the objective, and has a re ection o the imaging view port that also passes very close to the objective, as seen in 4.17. This limits our possible numerical aperture (NA) to its current value, as discussed in 4.4.8. 4.4.8 Florescence Collection The state detection via orescence, as discussed in 2.4 is achieved by collecting the spontaneous emission light at 369 nm from the ions using a microscope objective from CVI (UVO-20.0-10.0-355-532). This objective of NA 0.23 is situated roughly 3mm from the imaging port, and 13mm from the ion chain in e ective optical dis- tance, as illustrated in 4.20. This setup is very similar to that described in [52]. With an apex angle of 2 , the solid angle = 2 (1 cos ) = 2 (1 cos(sin 1NA)) = 78 z x y Raman 1 ?k Raman 2 Figure 4.18: Schematic of ion trap apparatus and the geometry of Raman laser beams for the spin-dependent force. Two Raman beams uniformly address the ions, with k along the transverse x-direction. The spin states are de ned with respect to a magnetic eld of 5 G along the y-axis. A photomultiplier tube (PMT) and CCD camera are used to measure the spin state of each ion through standard spin- dependent uorescence. 79 Figure 4.19: Pro le of beam as taken with CCD at estimated ion distance. Our home built Labview program ts for the horizontal (wide) and vertical beam waists 0:168, i.e. only 1:3% of the scattered light. The collected light then passes through a home built 400 m pinhole (a washer thinned by a mill and drilled through) at the image plane to reduce o axis background scatter. The image is further magni ed by a doublet pair to 130 and ltered by a pair of Semrock l- ters (FF01-370/36-25 and Hg01-365-25) that transmit 95% of the light at 369 nm , and block the other lasers and roomlight completly (unfortuantely, they do not block 355 nm ). The light is split into two paths - one for imaging, and one for cool- ing scatter collection. The latter is part of an automatic decrsytalization detection scheme. The imaging path has two possible con gurations - one where a ipper mirror diverts the light to the imaging PMT (Hamamatsu H10682-210), and one where the mirror is folded back and the image proceeds directly to the ICCD. The PMT beam path has a lens to demagnify the image, as long ion chain images are too 80 obje ctive PMT PMT Figure 4.20: Objective collects 1:3% of spontaneous emission from orescing ions. Then collected light passes through pinhole, dou- ble and lters(not shown) that are all mounted in lenstube atop a micrometer positioning stage for alignment. First PMT is for decrys- talization detection. PMT or ICCD imaging is selected via remote- controled ipper mirror, in box(dashed lines). 81 Figure 4.21: The averaged image of 1000 exposures of 10 bright ions. The pixels are binned in 4 64 super-pixels (hence the vertical com- pression). This image was produced by the script in appendix D for the data taken in 7. large to t the PMT aperture otherwise. In retrospect, we have overly magni ed the image. The only advantage to magni cation is ease of diagnosing the focus. As long as magni cation is enough that adjacent ions do not overlap on the same ICCD pixels the magni cation is su cient. Magnifying beyond that point leads to using more pixels than necessary, slowing down readout, adding readout noise and limiting eld of view. The binned image of 1000 3 ms exposures (produced by the script in appendix D) bright can be seen in gure 4.21. As discussed in the 2.4, the PMT Quantum E ciency (QE) is 30% at 369 nm , and only 20% for the PiMax3:1024i \Super Blue". When we detect the ion state with the PMT for 0.8 s , we collect on average 10 photons for a single bright ion, and this is su cient for detection with 97% delity. Due to the lower QE and readout noise of the ICCD, we must use alonger exposure time of 3 ms for the same detection light power. In addition, we bin the 1024 1024 pixel CCD into 4 64 superpixels to reduce readout time and noise[53]. With our parameters the detection delity for a single ion on the ICCD is 95%. For our magni cation and di raction limit, the ion occupies a Gaussian circle of 250 m on the ICCD (found from an unbinned 12:8 12:8 m pixel size) as seen in gure 4.22. When collecting data via PMT during experiments we t the histograms to simple Poissonians, with an average set by hand. When analyzing such data, we 82 Figure 4.22: High resolution video still of single ion, taken from our control program. Airy ring is somewhat visible. Most of collected light is from cooling. The pinhole is barely visible, outlined by light scattered from the trap and view port. 83 re t the histograms with more realistic distributions[31]. When collecting data with the camera the treatment is more complex. For real time data taking, we collect 1000 images of the ions when they are pumped dark, followed by 1000 images of the ions rotated to bright by a Raman pulse. The bright images are then averaged, and the image discriminated into dark and bright pixels according to a preselected brightness cuto . The bright pixels are used to nd the Regions of Interest (ROI) for each ion. The brightness of the entire region of each shot is then analyzed to derive histograms for the dark and bright state for each ion. A discriminator that gives equal state detection error for both histograms is then determined. The ROIs and discriminators are then returned to the Labview control program to be used in real time state discrimination. The script is presented in appendix D. Due to the spot size of the ions and their typical seperation of 2-5 m , we su er from cross talk. This can be seen in the results returned from the ROI script in table 4.1. Here we see that the detection error is lower at the center of the chain. This is due to the ions higher proximity at the center the chain. The spillover light from neighboring ions causes the bright histogram to be brighter and hence the lower overlap with the dark histogram seemingly reduces . To compensate for this cross talk and bias we post process our data. Using the same 1000 2 exposures used previously, we derive new histograms, this time by tting Gaussians centered at each ion to vertically summed data, as in gure 4.23. By counting frequency of the Gaussian tted amplitudes, we construct a histogram that accounts for ion cross talk as the t is found for the entire ICCD width with xed Gaussian waists and centers. Although this process discards one spatial dimension of the images, it does 84 Y1 Y2 X1 X2 Discriminator 1 5 45 48 7. 10650 1 5 56 59 5. 10650 1 5 65 69 4.9 14450 1 5 74 78 4.7 14350 1 5 82 86 4.3 14600 1 5 90 93 4.9 11100 1 5 98 102 4.8 14850 1 5 107 110 4.7 11300 1 5 116 120 4.6 14800 1 5 127 131 6. 14600 Table 4.1: ROIs and detection error % for 10 ions, as returned by the script in appendix D for the 10 ion data in 7. . not have much of an a ect on the discrimination between signal and background, as there is severe vertical image smearing at our ICCD clock speed setting (16 MHz ). We further enhance our data by correcting for state bias due to the detection error [54]. We construct the matrix M that describes the distortion of the probabil- ities for N ions of each real spin state gi into the detected probability for spin state fj = P2N i=1Mjigi Mji = (1 p0) n0 (1 p0) n1 p 0p 1 (4.19) Where p0 the error probability for one dark ion is equal (or should be) to the error probability p1 for one bright ion, n0 (n1 = N n0) is number of down (up) spins,and ( ) is number of down (up) spin ips to up (down) to misidentify state gi as state fj. We then simply invert this matrix to nd state probabilities closer to the real ones gj = 2NX i=1 M 1ji fi (4.20) 85 2000 4000 6000 8000 200 400 600 800 Histograms Bright and Dark of Ion 1 0 1000 2000 3000 4000 5000 6000 200 400 600 800 Histograms Bright and Dark of Ion 2 1000 2000 3000 4000 5000 6000 200 400 600 800 Histograms Bright and Dark of Ion 3 0 1000 2000 3000 4000 5000 6000 200 400 600 800 Histograms Bright and Dark of Ion 4 0 2000 4000 6000 200 400 600 800 Histograms Bright and Dark of Ion 5 1000 2000 3000 4000 5000 6000 7000 200 400 600 800 Histograms Bright and Dark of Ion 6 0 2000 4000 6000 200 400 600 800 Histograms Bright and Dark of Ion 7 0 2000 4000 6000 200 400 600 800 Histograms Bright and Dark of Ion 8 0 1000 2000 3000 4000 5000 6000 7000 200 400 600 800 Histograms Bright and Dark of Ion 9 0 1000 2000 3000 4000 5000 6000 7000 200 400 600 800 Histograms Bright and Dark of Ion 10 Figure 4.23: The background subtracted Gaussian tted histograms for 10 ions, for the data in chapter 7. The units are arbitrary. The red line is the found discriminator between down and up states, and is somewhat di erent between ions. 86 The algorithm above assumes that we have correctly diagnosed the state detection error, that the error is the same for all ions in the chain, and that the error is independent between ions. All of these assumptions are likely only approximately true, so we can expect improvement but not complete correction of detected state probabilities. 4.5 Arbitrary Waveform Generator (AWG) In order to generate the complex waveforms that drive the simultaneous M lmer- S renson and Raman gates, we drive AOM 2 with a a DA12000-16 from Chase Sci- enti c. This is a standard PCI expansion card that outputs waveform at up 2 GHz and has 16MB of on-board waveform memory. We have setup the AWG to operate at 1 GHz , and it is phase locked to a PTS3200 Direct Digital Synthesizer (DDS). This oscillator itself is phase locked to the same 10 MHz Rubidium clock as AOM 1, thus locking the phase between the modulations on both beams. Unlike all other oscillators used in the experiment, the AWG does not run continuously - it outputs the stored waveform when it receives a TTL trigger, and then stops. This makes the phase jitter on this trigger critical, as the jitter is transformed into noise on the phase relationship between the Raman AOMs. When a trigger is received, the AWG reads each byte in sequence from memory, and converts each byte with a 12 bit Digital to Analog Converter (DAC), such that 2047 is equivalent to 0V, 4096 is the max voltage, and 0 is the minimum voltage. This limitation on the vertical resolution does not create noticeable e ects in the frequency domain as the AWG is 87 bandpassed. However it limits the resolution when dealing with waves of di erent magnitudes with physical e ects of the same order of magnitude, i.e when we drive the M lmer-S renson gate we drive sidebands > 10 times stronger than the wave that drives the Raman coupling. Thus, our resolution for the Raman coupling will be limited to much less than 212. As the resonant frequency of the AOMs is 210MHz, the AWG will typically have only 4-5 points per wave cycle. This is su cient to generate a frequency that can be bandpass ltered of its alias at fclock fsignal 800 MHz . Therefore, we chose to run the AWG at half its possible clock frequency and double the duration of the waves we can generate. At a clock rate of 1 GHz , a byte is converted to an analog voltage every 1ns, allowing us a total waveform time of 16 ms . This is more than su cient for our experiments. Unfortunately, it is not su cient for all of our diagnostic procedures, such as measuring the spin coherence time. Besides the waveform duration limitation, there is also a limitation on wave- form upload speed. For longer experiments (3-4 ms ), the upload time becomes signi cant. The DA12000 has a standard PCI type connector. Naively, this should imply a waveform upload speed of 133MB/s at least, so 120 ms upload time for the entire wave. Unfortunately, we have characterized the upload time to be approx- imately one hundred times slower. As a result, we are currently in the process of replacing this AWG with a more advanced model. The AWG is controlled by functions provided in a DLL by Chase. As we in- terface the AWG via Labview, it was necessary to construct a DLL to interface the function library (da12000 dll.dll) and the card. The chief function of this library 88 is the translation of data structure from Labview into the data structure expected (\SegmentStruct", as de ned in appendix C) by the chapter loading function. Cur- rently, the chapters run consecutively, and only the rst chapter is triggered. A pre- vious attempt to have multiple triggered waveforms results in the occasional missing of a trigger (likely due to a trigger being sent before a chapter was completely gen- erated). This error a ects all following experiments as the AWG chapter execution becomes mismatched with the other triggers/oscillators from the rst AWG chapter being generated. There is also an option for a chapter to loop for a xed number of times. This may be useful for increasing the AWG e ective time (for instance, for experiments where the AWG should generate a short pulse, a long duration of no output, and then another short pulse - i.e. Ramsey experiment). 89 Chapter 5 Experimental Procedure 5.1 Introduction Before each experiment with our apparatus, we tend to perform the same preparatory tests and calibrations. In this chapter I will go through these proce- dures and some of the trouble shooting options when the system is more severely miscalibrated than usual. 5.2 Temperature Before preparing to trap ions, all lasers must be correctly set and stable. Therefore, the rst action of the day is turning on the lasers and allowing them to cool-down (or warm up) and equilibrate. For the DL-100 and Vanguard lasers this takes approximately half an hour. For the Verdi and MBR-110 this may take two hours, so generally we do not turn these lasers o when taking data. The Verdi and MBR require chillers for their operation, so these must be activated before the lasers. For the Vanguard, this is done automatically by MLUV. The MLUV program will also display the percent warm up of the THG 4.11. The Pi:Max ICCD must be cooled down to its operating temperature as well ( 250C). The ICCD is cooled by a water circulating CPU cooler, operating at room 90 temperature. This is due to the crowding on the table preventing proper cooling by the ICCD fan. The ICCD is ready for operation in minutes. RF power to the trap should be set to the operating power used for the ex- periment (typically around 26dBm), as the delivered power changes the resonant frequency of the trap. The set point of the TC200 heater for the resonator should be checked. 5.3 Frequencies Now that the lasers are hopefully stable, their frequencies should be set. As- suming the lasers are well aligned, we set the laser frequencies for 171Yb+ trapping, as indicated in table 5.1. These frequencies are measured by a HighFinesse solid state wavemeter, via optical bers, and displayed by a homebuilt web client 1. If loading fails, it might be necessary to attempt to load 174Yb+, as it does not have hyper ne structure and hence scatters more light, and does not require all the laser modulation needed by 171Yb+ nor can it have a coherent dark state [1]. The wave- lengths for 174Yb+ are listed in table 5.1 as well. When alternating to 174Yb+, the loading rate may be highly increased by switching over to the natural isotope dis- tribution Yb oven, although it is possible to load from our usual high purity 171Yb oven. After setting the laser frequencies, the lasers must be stabilized, as discussed in chapter 4, as well as the wavetrain frequency doubler for the MBR. This often requires slight adjustments of the input coupling lens. 1Created by Peter Maunz. 91 Laser 171Yb+ 174Yb+ MBR-110 405.644318 THz 405.645530 THz 935 nm DL-100 320.56922 THz 320.57190 THz 399 nm DL-100 751.52764 THz 751.52680 THz 638 nm DL-100 469.445-469.442 THz 469.439 THz Table 5.1: Laser frequencies for trapping. The 638 nm can be locked anywhere in above range as long as it is mode-hope free for the entire range. The Vanguard shutter should be opened and emission engaged (via MLUV). The Vanguard frequency cannot be set, rather the feed forward lock on AOM2 is engaged. For the RF resonator, the reference signal is adjusted to close to 0 by slightly adjusting the RF oscillator, and then engaging the RF frequency lock. A moderate gain must be set here (by tuning the HP8640B modulation strength knob) as recrystallization will severely change the resonance frequency. Also, the Labview voltage control program should be checked to see that there is no attenuation of the trap RF power. Once all lasers are locked one may turn on the Yb oven, and after a minute of oven warm up (at 2:4A) one way drop the trapping voltages and RF power using the recrystallization settings, while radiating the trap with the 399 nm light and alternating pulses of 355 nm light, while waiting for an ion or ion cloud to appear in the Pi:Max3 viewing program ( gure 4.22). 5.4 Loading and Alignment If an ion is successfully loaded, then beam alignments must only be slightly adjusted to maximize cooling and detection counts. If loading is slow one may attempt to adjust the 399 nm beam, as that is the most likely to be misaligned. If 92 the detection or cooling seem to be unstable, one may try adjusting the 935 nm beam alignment and polarization. In general, as all 369 beams share the same nal path, it is reasonable to assume that if they are all overlapping then the beams are close to the correct alignment. Another sanity check is the scatter patterns of the trap electrodes. By diverting the beams slightly up or down from the center of the trap, one may nd the gaps between the trap electrodes by viewing the pro le of the beam on exit through the view port. If all beams seem reasonably aligned and loading is not seen on ICCD, then the PMT counts should be viewed instead. If this fails as well, then Yb oven temp should be increased. If loading still fails, then the trap may have shifted relative to the imaging system. In order to realign the imaging system, one must detect a trap corner (as shown in gure 5.1), and thus reorient the imaging system micrometer settings relative to the trap. Vertical alignment of the beams based on the clipping pro le of the beams is trivial. It is important to remember that the beams are entering the trap at a 450 angle, and the horizontal lens alignment of the beams is measuring the displacement of the beams at that angle rather than parallel to the trap. When viewing the corners of the trap, it is most likely the corners of the electrode furthest from the camera that are being viewed. When attempting to load over long periods, it is important to occasionally shut o the trap RF so as to release possible trapped ion clouds. These do not crystallize nor cool e ciently and are thus less visible than single ions. Once the ions are loaded it is important to verify that the cooling and detecting 93 Figure 5.1: (a) Lower left trap corner. (b) Upper right corner. Pin- hole is clearly visible due to high scatter o trap. Burs are due to imperfect drilling. beams are centered on the ion chain. This can be done by moving the ions across the camera (using the trap voltages to push them in the z-direction), but PMT detection counts are more exact. If one intends to take data using the ICCD, it is important to verify that all ions fall on the active PMT surface as the camera and PMT are independently aligned. There seem to be certain regions on the PMT surface that are not active. If it is suspected that the protection beams require alignment, then the iodine lock must be disengaged and the MBR tuned 400 MHz blue so as to bring those beam frequencies closer to transition resonance. 5.5 Raman Beams Rough alignment of Raman beams may be done using an independent Ramsey experiment for each beam using a qubit frequency microwave pulse, or an AC Stark shift measurement. For the Ramsey stark shift measurement, the Vanguard feed 94 forward is unlocked, and an RF switch diverts the microwave reference to via an ampli er to microwave horn (the microwave source frequency is slightly adjusted so as to be exactly on carrier). First, the Rabi oscillation =2 time is measured, i.e. the time it takes for the qubit to reach half of maximum brightness (the qubit is then in the x-y plane of spin space). Second, using the Experimental control program we schedule three experimental pulse sequences (\chapters"): a =2 pulse, a 355 nm pulse, and another =2 pulse. By scanning the duration of the 355 nm pulse, we can see \Ramsey fringes" - oscillations of the qubit due to the Stark shift of the qubit frequency. This method is very sensitive as the pulses can be 10?s of milliseconds long. Once both beams are roughly aligned we can drive Raman transitions with both beams. At that stage maximizing the Rabi frequency is the fastest way to improve alignment. Finally, we return to Stark shift Ramsey measurements on the beams. We can then perform a long Ramsey experiment with no 355 nm light to measure accurately the carrier frequency, as any mismatch between the =2 pulses and the qubit frequency will lead to oscillations where there should be none. Now we use the Ramsey experiment to validate that the beams are truly cen- tered. We shift the ion a distance on the order of the length of long ion chain, and validate that the Ramsey frequency is symmetric, adjusting the horizontal microm- eter to achieve this. Barring suppression by micromotion, coupling to the y-mode, or incorrect polarization our Rabi carrier frequency is now maximized. Both beams should be linearly polarized, with electric elds parallel to table so as to drive the Raman transition with +; + polarization in the ion?s frame of reference. 95 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 Raman beatnote detuning from carrier (MHz) Ram an spectru m for 10 ions, with 355 nm pulse d lase r tilt COM Carrier Figure 5.2: 10 ion Raman spectrum. The x-axis is in MHz detuning from carrier, that is centered at 0 MHz . The dominant motional modes are the x-transverse modes. As can be seen here, their width is only on the order of 100 KHz , vs the carrier which is on the order of MHz . The highest frequency mode is the COM, and second is tilt. The lowest is the zig-zag mode, only a MHz away from carrier for these trap settings. Some of the weaker features are axial and y-modes, which have a weak coupling to our beam geometry. 5.6 Calibration As we have identi ed our qubit frequency exactly, we are now interested in mapping out the motional modes, as can be seen in gure 5.2. If we do not know the general location of the motional modes (due to changes in trapping voltages, for instance), we may load two ions and measure the COM and tilt modes. That is su cient to calculate all other modes. This can be done by scanning the modulation frequency of Raman beam 2 with a xed pulse time. If the pulse duration is long enough and the intensity high enough to drive several oscillation periods for the weakest transition of interest then we can expect to see all resonances. 96 Figure 5.3: Results for two di erent intensities of the automatic side- band nder. The program nds increasingly accurate RSB and BSB frequencies by nding upper and lower limits for decreasing inten- sity and lengthier pulse times. Before nding the RSB, a pulse is applied. Once we identify the modes, we may preemptively apply two sidebands, roughly 100 KHz red detuned from the COM, with the intensity we expect to use for our M lmer-S renson coupling gate. This will ensure that the following measurements will be immune to higher order Stark shifts. However we must shift all frequencies up by the measured Raman 2 Stark shift ( 200 KHz , vs 600 KHz for Raman 1), as they were performed at the lowest power possible. After applying side-band cooling to the COM mode, we may activate the automatic sideband nding functionality of the control program, as shown in gure 5.3. After measuring Rabi oscillations frequencies for the BSB and RSB, we set them equal (equal AWG amplitudes for both frequencies do not necessarily drive equal strength sidebands), and detune them equally from resonance. We then add a carrier frequency, and measure its frequency for three di erent powers, for AWG amplitudes 80 times lower than the sideband amplitudes. Thus, we have cali- brated the AWG amplitudes to real frequencies. The relationship is linear for these low values. We now add a second ion, and remove the carrier. We optimize the spin-spin 97 coupling by alternately scanning the amplitude of one the sidebands or a sideband frequency. This should lead to decent (0.1-1.7 ions bright) contrast, and roughly ve coherent oscillations at our typical operating parameters. If the coherence is bad it is likely due to the vertical alignment of the Raman beams. 5.7 Imaging Calibration We may now load all the ions needed for the experiment. After allowing the RF resonator to settle from the power drop needed for loading, we prepare an experiment sequence where the ions will cooled, then pumped dark, and then detected. The number of experiments per point is set to 1000. Then the camera calibration scan is activated, and we select the \Ions Dark?? option. We add a pulse to rotate all ions to bright after the optical pumping, and repeat the calibration procedure, this time selecting \Ions Bright??. The Mathematica script of appendix D will then automatically start. The control program will display the averaged bright ion images, with ROI boxes and detection parameters. We are now ready to select our experimental chapters and parameters, and to begin experimenting. 98 Chapter 6 Summary of Experiments 6.1 Introduction In this chapter I will summarize the experimental results we have achieved in our group leading up to the work reported in chapter 71. The rst work I will report was our demonstration and control of spin-spin couplings between the pseudo-spins for two and three ions using global Raman beams. Once that ability was mastered, we were able to proceed with the adiabatic quantum simulation experiments: Mea- suring the phase diagram of the ground state Hamiltonian for three spins; Creating a frustrated triangle of AFM coupled spins; Creating a phase transition from the paramagnetic state to FM and AFM states with 16 spin. For most of this work, our Raman beams were driven by an the MBR, detuned only 2.7 THz from the 2S1=2 2 P1=2 transition, and with only 200 mW of UV light. These gates were en- abled by microwave frequency combs of the EOM modulated laser beams, at roughly half the qubit frequency. These beams were then replaced by the 33 THz detuned 355 nm laser beams, with 4W of UV. This scheme o ers the same coupling strengths with much lower unwanted scatter rates. Using the simulated gates I introduced in chapter 3, we will simulate the 1This chapter follows closely parts of [55] 99 Hamiltonian H = X i 10 by rotating the principal Y -axis of motion to be nearly perpendicular to the laser beams, as described in chapter 4.4.6. We direct two Raman laser beams onto the ions to drive spin-dependent forces, with their wavevector di erence aligned along the transverse x-axis of ion motion ( k = k p 2, where k = 2 = ). The Raman beams are detuned 0:5 2:7 THz to the red of the transition at a wavelength of = 369:76 nm . The Raman beams are phase modulated at a frequency nearly half of the 171Yb+ hyper ne splitting with a 6:32 GHz resonant EOM, and each of the two Raman beams have independent AOM shifters in order to select appropriate optical beatnotes to drive Raman transitions [56]. The Raman beams are focused to a waist of approximately 30 100 m with a power of up to 10 20 mW in each beam. When their beatnote is adjusted to drive the carrier transition at the hyper ne transition HF , we observe a carrier Rabi 100 frequency of i=2 1 MHz. For the spin-dependent force, we set i=2 0:4 MHz for each pair of Raman beam and for the transverse eld, i=2 is less than 0.1 MHz. The resonant transition generates the e ective transverse eld by adjusting the phase with respect to the spin-dependent force. For up to nine ions in a chain, we observe that the outer two ions experience 2% lower Rabi frequency, and the variation in the di erential AC-Stark shift in each qubit of < 1%. In the experiment, we rst Doppler laser cool 171Yb+ ions for 3 ms using a laser tuned red of the 2S1=2 2 P1=2 transition at a wavelength of 369:53 nm. We then Raman sideband cool all m modes of transverse motion along x to mean vibrational indices of nm < 1 in about 0:5 ms, well within the Lamb-Dicke limit. Next, the ions are each initialized to the j#i state through standard optical pumping techniques [1]. We then apply the optical spin-dependent force on the ions for a duration by impressing the bichromatic beatnotes at HF . Afterward, the spin states are measured by directing resonant laser radiation having all polarizations on the 2S1=2(F = 1) 2 P1=2(F 0 = 0) transition following standard state-dependent uores- cence techniques [1]. We use a charge-coupled-device (CCD) imager (the detection e ciency is, 95 % per spin). We determine the probability of each spin con guration (for example P""# ) by repeating the above procedure more than 1,000 times. We also measure the probability Pn of having n spins in state j"i by using the PMT, which is useful for higher-e ciency measurements of certain symmetric observables such as entanglement delities and witness operators. 101 J ( k Hz ) (a) (b) (d) (e) (b) (c) (d) (e) ms Pavg Pavg t (ms) 1 p x1 p x 1 p x 1 p x t (ms) (c) t (ms) t (ms) Pavg Pavg Theory Experiment Figure 6.1: (a) Measured coupling J for two ions as a function of detuning overlaid with theory (lines) from Eq.(3.22) with no free t parameters. The detuning is scaled to the axial ( z) and transverse ( 1) COM normal-mode frequencies of motion such that COM, and tilt modes of transverse motion occur at S ( 2 21)= 2 z = 0 and 1, respectively. (b) Time evolution of the average number of ions in the j"i state under the in uence of the bichromatic force in the far-detuned limit, showing the secular oscillation of the Ising spin- spin coupling, where the detuning =2 is 80 KHz from the COM motional sideband. (c) Measurements with 1 = 2 p 3 1 =2 56 KHz . Here the small oscillations from the motional excitation and the coupling to spin states are noticeable on top of the sinusoidal oscillations of the Ising interactions. (d) Measurements at 1 = 2 p 2 1 =2 45 KHz . (e) Measurements at 1 = 2 1 =2 32 KHz . The insets of (b)-(e) show the respective wavepackets in phase space and the areas enclosed are shaded. 6.2.1 Two spin case Figure 6.1(a) shows the theoretical values of J = J1;2 from Eq.(3.22) and measurements at various detunings for two spins with J = 1 2( 21 1 2 22 2 2 2 2 22 ), where 1 = k= p 4m 1 and 2 = k= p 4m 2 are the Lamb-Dicke parameters for COM and tilt modes. The solid theoretical curve is plotted with no adjustable parameters, as the motional mode frequencies and the sideband Rabi frequencies are independently measured. We measure the strength of J by observing the time evolution between j##i z and j""i z after applying the spin-dependent forces on j##i z states. The evolution 102 for the two spins is simply described by U(t) j##i z = exp( iJ (1) x (2) x t) j##i z = cos(Jt) j##i z+i sin(Jt) j""i z when we neglect the couplings between internal states and the motion. As shown in gure 6.1(b), the oscillations of the populations are sinusoidal, with frequency 2J . To detect the sign of the Ising coupling, we applied the same force on the initial state j""i and observe the phase of the oscillations. When the beatnote detuning is close to a vibrational mode, or j 1;2j is within a few sideband linewidths 1;2 , the coupling between motion and spin states modulates the spin state evolution. Figure 6.1(e) shows the spin state evolution at a detuning = 1 = 2 1 [3]. In this case, the motional state is displaced in phase space by no more than j j = 1, and at particular times during the evolution = n= (n = 1; 2:::) the motional degree of freedom is decoupled from the internal state, enabling the generation of pure spin-spin entanglement. As the detuning increases as shown in gure 6.1(b)-(d), the maximum displacement decreases and the evolution approaches a pure sinusoid indicative of pure spin-spin interactions. Typical experiments are performed with 4 , where the largest displacements in a 2% modulation in the spin evolution. 6.2.2 Three spin case For three spins addressed uniformly with the Raman laser beams, we have J1 = J1;2 = J2;3 as the nearest-neighbor (NN) interaction and J2 = J1;3 as the next- nearest-neighbor (NNN) interaction as shown in gure 6.2(a). Since the bandwidth of the transverse mode spectrum is relatively small, all modes can be addressed from 103 J ( k Hz ) ms (a) (b) (c) P0 4 J 1 2J 1 - 2J 2 2J 1 +2J 2 t (ms) Freq uenc y (kHz) J1 J2 Calculation Experiment Figure 6.2: (a) Measured couplings J1 = J1;2 = J2;3; J2 = J1;3 for three ions as a function of detuning overlaid with theory (lines) from Eq.(3.22) with no adjustable parameters. At the scaled detuning S, COM, tilt and zig-zag modes of transverse motion occur at S ( 2 21)= 2 z = 0; 1 and 2:4, respectively. (b) Time evolution of the probability P0 of all spins j#i under the bichromatic force in the far-detuned limit. Here, the two couplings J1 and J2 are clearly visible. The solid line is a t to the time evolutions of Eq.(6.1) with an empirical exponential decay. The measurements are performed at the indicated detuning in (a), or -50 kHz from tilt mode. (c) The Fourier transform of the experimental curve shown in (b), where three peaks originate from the frequency components of 4J1, 2J1 2J2 and 2J1 + 2J2. The orange bars represent the calculated values from Eq.(3.22). The peak near the origin comes from the overall decay of the oscillation due to decoherence with 100 Hz. 104 a single laser beatnote and the signs and the strengths of J1 and J2 are under great control as shown in gure 6.2(a) and Eq.(3.22) [57]. In the region > 1, both have anti-ferromagnetic (AFM) interactions (J1;2 > 0), and in the region of 2 < < 1, both have ferromagnetic (FM) couplings (J1;2 < 0). In the region 3 < < 2, the NN interaction is FM (J1 < 0) and the NNN is AFM (J2 > 0). When is near the tilt mode, J2 overpowers J1 and when is closer to the zig-zag mode, J1 is stronger than J2. Finally, for < 3, all interactions are FM again. We measure the J1 and J2 couplings by observing the oscillations in the pop- ulation of state j###i z after applying spin-dependent force on the three spins. This population oscillates as cos2 J1t cos J2t i sin2 J1t sin J2t, as shown in gure 6.2(b). We use Fourier analysis on the oscillations and nd certain frequency combinations of the couplings 4J1, 2(J1 J2), and 2(J1 + J2), as shown in gure 6.2(c). In this gure, we use theoretical values for the signs of J1 and J2. 6.3 Adiabatic quantum simulation In this section, we describe the adiabatic quantum simulation of the transverse Ising model with three spins, where the exact solution is known, and discuss the criteria for adiabaticity. We then present experimental results for two example Ising interactions strengths and signs. We experimentally investigate this adiabaticity criterion for the two di erent types of NNN coupling that were introduced above. We initialize the spins along the By-direction through optical pumping ( 1 s) and a =2 rotation about the 105 x-axis of the Bloch sphere. Figure 1.3 shows the adiabatic simulation protocol. The simulation begins with a simultaneous and sudden application of both By and J1; J2 where By overpowers J1 (By=jJ1j 10). As the spins are aligned along the y-axis, the sudden turn on will not cause diabatic e ects. A typical experimental ramp of By decays as By(t) = ae t= + b with a time constant of 30 s, varying from a 10 kHz to a nal o set of b 500 Hz after t = 300 s. By varying the power in only one of the Raman beams, this procedure introduces a change in the di erential AC Stark shift of less than 2 Hz. We turn o the Ising interactions and transverse eld at di erent By=jJ1j endpoints along the ramp. We then measure the magnetic order along the x-axis of the Bloch sphere by rst rotating the spins by =2 about the y-axis, and detecting the z-component of the spins. In gure 6.4(a) all interactions are FM and J2=jJ1j 2 4(as in gure. 6.3(a)). The dashed lines in the top panel are the adiabaticity parameter from Eq.(1.6) cal- culated over the trajectory for the two coupled excited states (recall gure 6.3). Due to the 500 Hz nal o set of By, the simulation stops at By=jJ1j 0:5. To examine the behavior extended below this value, we calculate the criteria for an exponential ramp with a 100 s time constant. This pro le was chosen to overlap with experimental parameters for large By=jJ1j and also reach By = 0 in a typical simulation time ( 300 s). The results indicate that Eq.(1.6) is satis ed over the trajectory; _By(t) = 2ge remains much less than one even with a maximum occurring at By=jJ1j 1. To demonstrate the simulation is indeed adiabatic for these parame- ters, we plot the measured probability of occupying a FM state P(FM) = P"""+P### (solid dots) in gure 6.4(a). The black line represents the adiabatic ground state 106 (a) (b) J2 / |J1|~2.1 J2 / |J1|~ - 0.9 |????y|????, |???? |????y |????, |???? Figure 6.3: Energy level diagrams for Eq.(6.1) with two di erent types of spin-spin interactions. For both panels, the NN interactions are FM (J1 < 0). (a) The NNN interaction is FM with J2=jJ1j -2 and (b) AFM with J2=jJ1j 0.9. The arrow in both diagrams in- dicates the trajectory in the simulation, initialized at By=jJ1j 10. Under this condition, the initial ground state is an eigenstate of sec- ond term in Eq.(6.1), a polarized state along By. In both examples, at B J1 some high energy states cross, but the ground state (black solid line) has no level crossings with any excited state. Likewise, the highest energy state does not cross any other levels, allowing one to also adiabatically follow this state. The dotted lines represent excited states which are most strongly coupled to the ground state along the path. In the large eld limit, the energy di erence between ground and excited states ge (here, scaled by q B2y + J 2 1 ) is proportional to By, but as By=jJ1j decreases the spin-spin couplings determine the energy di erence and the form of the ground state. In both (a) and (b), the nal ground state is FM (de ned along the x-axis of the Bloch sphere), however in the case of (a), the minimum ge is 20 times larger. 107 P(F M) (a) (b) J 2 / |J 1 | ~ 2.1 J 2 / |J 1 | ~ - 0.9 B y / |J 1 | B y / |J 1 | Ground State Experimental ramping t =1 00 ms ramping Data Ground State Experimental ramping t =1 00 ms ramping Data Figure 6.4: (a) The theoretical order from the exact experimental ramp with a 35 s time constant and nal o set value given in the text (gray solid line) is in reasonable agreement with the order in the true ground state (black solid line) for By=J1 > 0:5. The dotted line is the expected state evolution for a pure exponential decay ramp with a 100 s time constant, allowing By ! 0. (b) The data also matches well to theory, as we avoided the regions where diabatic transitions are expected for By=J1 1. According to the calculations, the duration of three-spin experiments near the special point should be on the order of milliseconds. and the grey line is the theoretical expected probability including the experimental ramp. The dotted line in this gure is the theoretical state evolution using a By- eld ramp that reaches zero. The predicted evolution does not signi cantly deviate from the ideal ground state and the data is in good agreement with all three theory curves. Figure 6.4(b) presents the case when theNNN interaction is AFM and J2=jJ1j 0.9 (as in gure 6.3(b)). When By=jJ1j 1, _By(t) = 2ge reaches a maximum value of 0:6, indicating that the probability for excitations will likely increase. This di er- ence is because in this case the gap ge at the ?critical? point is 15 times smaller than that in gure 6.3(a). In contrast to the FM J2 case, the theoretical probability curves shown in the lower panel of gure 6.4(b) predict signi cant diabatic e ects 108 when using this By- eld pro le for simulations near the critical point. In fact, to successfully evolve to the true ground state near By = 0, the simulation time (as- suming same initial conditions and an exponential ramp of By) should be at least a factor of ten longer. Because all the data lies outside of the region where the energy gaps are small, the diabatic excitations are minimal, but further experimental study is needed to precisely quantify this e ect. One method to probe excitations, which may also be useful as N 1, is to perform and then reverse the experimental ramp and measure the probability of returning to the initial state [58]. 6.4 Phase diagram Assuming adiabaticity as described above, we can generate an experimental phase diagram for the transverse Ising model. We will rst describe this for the three spin case, and then discuss speci c features and scalability in sections 6.6 and 6.7. For the three spin Hamiltonian in Eq. (6.1), the competition between the two spin-spin couplings and the transverse eld gives rise to a rich phase diagram. Here we label the 23 possible spin con gurations as the two FM states, j"""i and j###i , two symmetric AFM states, j#"#i and j"#"i , and four asymmetric AFM states, j""#i , j"##i , j#""i and j##"i , all de ned along the x-axis of the Bloch sphere. In gure 6.5(a), we plot a part of the theoretical phase diagram where the nearest-neighbor interactions are always FM (J1 < 0). The order parameter is the probability of occupying a FM state, P(FM) = P""" + P###. For regions where 109 B y / |J 1 | B y / |J 1 | B y / |J 1 | J 2 / |J 1 | P(FM) P(F M) J 2 / |J 1 | J2 / |J 1 | (a) (b) (c) (d) P(FM ) (c) (d) Figure 6.5: (a) Theoretical phase diagram for Eq.(6.1). The color scale indicates the amplitude of the FM order parameter, P (FM) = Pj"""i +Pj###i . Here, J1 is always negative, yielding FM order in that coupling. In the region where J2=jJ1j <0, there is a crossover to FM order as By=jJ1j is lowered. When J2=jJ1j >0, the AFM and FM interactions compete. When J2=jJ1j =1 and By =0 the ground state is comprised of 6 states: four asymmetric AFM and two FM states. This creates a special sharpened point where all lines of equiprobable FM order converge. (b) Experimental measurements of the phase diagram for Eq. 6.1 (solid bars) compared to the theoretical pre- diction from Fig. 6.5 (surface) . The vertical amplitude is the FM order parameter P(FM)= Pj"""i + Pj###i . The ratio of By=jJ1j was varied from 10 to 0.1 for J2=jJ1j values of -1.3,-2.0, -3.6, 4.2, 2.0, 1.3, 0.92, 0.74, and 0.62. J1 < 0 for all traces. (c) As By=jJ1j ! 0 in the region where J2=jJ1j < 1 , we observe a smooth crossover to FM order. The lled circles and solid line are the data and theory for J2=jJ1j = 1:3, respectively. (d) When changing J2 for a xed and small value of By=jJ1j the system undergoes a sharp transition. The data ( lled circles) shown is for a scan of By=Jy = 0:57. The average deviation per scan of By=jJ1j from the exact ground state is 0:09. 110 By=jJ1j 1, the ground state is polarized along By with P(FM)= 2=2N = 1=4, as all states in the x-basis are equally populated and there are two FM states. As By=jJ1j decreases, di erent magnetic phases arise. When the NNN interaction is also FM (J2 < 0), and By=jJ1j 1 the ground states are the two degenerate FM states. In the region where the NNN interaction is AFM and J2 overpowers J1 (J2=jJ1j > 1), the asymmetric AFM states are lowest in energy. A special point appears at J2=jJ1j = 1 and By = 0, where all the contours of constant FM order meet. Here, the ground state will be a superposition of the FM and asymmetric AFM states. This e ect arises because the pairwise interaction energy cannot be minimized individually, leading to a highly degenerate, or frustrated, ground state [59]. To be clear, when J1 = J2, both the FM and AFM ground states are degener- ate. As can be seen in gure 6.3(b), as the transverse eld is reduced the energy gap will go to zero, and adiabaticity will be impossible. For the FM case, there is a two fold degeneracy. For the AFM case, there is a six fold degeneracy. However, for our experiments we always have a slight transverse eld remaining and the degeneracy is lifted, creating the non-degenerate, superposed ground states we discuss in this paper. This procedure is performed for nine di erent combinations of J1 and J2 de- termined by the beatnote detuning from Eq.(3.22). In gure 6.5(b) we present the results as a 3D plot of the FM order parameter, with the theoretical phase diagram (surface) from gure 6.5(a) superimposed on the data. The data is in good agree- ment with the theory (average deviation per trace is 0:09) and shows many of the 111 essential features of the phase diagram. As By=jJ1j decreases, a smooth crossover from a non-ordered state to FM order occurs in the region where J2=jJ1j < 1 (Fig. 6.5(c)). The data (e.g. gure 6.5(c) show small amplitude oscillations in the initial evolution due to the sudden application of the spin-spin interaction, which is held constant during the simulation to minimize variation in the di erential AC stark shifts. As the number of spins increases, this is an example of a quantum phase transition. A rst order transition due to an energy level crossing is apparent ( g- ure 6.5(d) when changing J2 for a xed and small value of By=jJ1j = 0:57. This transition is sharp, even in the case of three spins [60]. 6.5 Spin frustration and entanglement We also study the properties of the ground states in the case of a frustrated Ising Hamiltonian. Frustration in spin systems occurs when spins cannot nd a ground state that minimizes the energy of each pairwise interaction [61, 12]. As shown in gure 6.6(a), this can be simply illustrated by three spins with AFM inter- actions on a triangular lattice [62]. The situation gives rise to a large ground state degeneracy, leading to magnetic analogues of liquids and the crystal arrangement of ice [63, 64]. For quantum spins, the frustrated ground states are expected to be directly related to entanglement [65, 66]. We realize the textbook example of spin frustration in a unit triangular cell with AFM interactions by setting the Raman beatnotes to the blue side of COM mode. For comparison, we also study the ground state property of all FM interac- 112 J>0 1 2 3 J>0 J>0 1 2 3 J1 J1 J2 (a) (b) (c) (d) t (ms) t (ms) B y / J rms B y / J rms |???? |???? |???? |????|???? |???? |????|???? P opula tions P opula tions Ground State Experimental ramping Ground State Experimental ramping Figure 6.6: (a) Simplest case of spin frustration in a triangular ge- ometry with AFM interactions. (b) Image of three trapped atomic 171Yb+ ions in the experiment, taken with an intensi ed CCD cam- era. The spins in the linear ion string have nearest-neighbor (J1) and next-nearest-neighbor (J2) interactions mediated through the collec- tive vibrational modes. (c),(d) Evolution of each of the eight spin states, measured with a CCD camera, plotted as By=Jrms is ramped down in time, with each plot corresponding to a di erent form of the Ising couplings. The dotted lines correspond to the populations in the exact ground state and the solid lines represent the theoretical evolution expected from the actual ramp, including nonadiabaticity from the initial sudden switch-on of the Ising Hamiltonian. (c) All in- teractions are AFM. The FM-ordered states vanish and the six AFM states are all populated as By ! 0. Because J2 0:8J1, a population imbalance also develops between symmetric and asymmetric AFM. (d) All interactions are FM, with evolution to the two ferromagnetic states as By ! 0. 113 tions by setting to the red side of COM mode. A linear chain of three ions can have NN and NNN interactions through the collective normal modes discussed in section 3.4 ( gure 6.6(b)). The experimental procedure to prepare the ground states of the Hamiltonian with all AFM interactions and all FM interactions is the same as the description in the section 4. Figure 6.6(c) shows the time evolution for the Hamiltonian frustrated with nearly uniform AFM couplings and gives almost equal probabilities for the six AFM states (three-quarters of all possible spin states) at By 0. Because J2 < 0:8J1 for this data, a population imbalance also develops between symmetric and asymmetric AFM states. Figure 6.6(d) shows the evolution to the two ferromagnetic states as By ! 0, where all interactions are FM. The adiabatic evolution of the ground state of Hamiltonian (6.1) from By Jrms to By Jrms should result in an equal superposition of all classical ground states and therefore carry entanglement. For instance, for the FM case, we ex- pect a GHZ ground state j###i - j"""i . For the isotropic AFM case, we expect the ground state to be j##"i + j"##i + j#""i - j""#i - j#"#i - j"#"i . We char- acterize the entanglement in the system at each point in the adiabatic evolution by measuring particular entanglement witness operators [67]. This is accomplished by performing various global rotations to the three spins before measurement, and combining the results of many identical experiments. When the expectation value of such an operator is negative, this indicates entanglement of a particular type de ned by the witness operator. For the FM case, we measure the expectation of the symmetric GHZ witness operator WGHZ = 9I^=4 J^ 2x (1) y (2) y (3) y [68, 67], where I^ is the identity operator and J^i 12( (1) l + (2) l + (3) l ) is proportional to 114 (a) (b) B y /J rms B y /J rms ?W GHZ ? ?W W ? t (ms ) t (ms ) Gro u n d Sta te Exp e ri m e n ta l ra m p i n g In c l u d i n g s p i n - m o ti o n c o u p l i n g Gro u n d Sta te Exp e ri m e n ta l ra m p i n g In c l u d i n g s p i n - m o ti o n c o u p l i n g Figure 6.7: (a) Entanglement generation through the quantum simu- lation for the all FM interactions. We measure the entanglement by observing the expectation of a particular operator that indicates the presence of entanglement, called the entanglement witness [67]. For the case of all FM interactions we use the GHZ witness [67], sensitive to the state (j###i j"""i )= p 2. We nd that entanglement occurs when jByj=Jrms < 1. (b) Entanglement generation for the case of all AFM interactions. Here, we use the symmetric W-state witness, sen- sitive to the state (j##"i +j#"#i +j"##i j#""i j"#"i j""#i )= p 6 and we nd that entanglement emerges for By=Jrms <1.1. In both (a) and (b) the error bars represent the spread of the measured expec- tation values for the witness, likely originating from the uctuations of experimental conditions. The black solid lines are theoretical wit- ness values for the exact expected ground states, and the black dashed lines describe theoretically expected values at the actual ramps of the transverse eld By. The blue lines reveal the oscillation and suppres- sion of the entanglement due to the remaining spin-motion couplings, showing better agreement to the experimental results. Note that the residual spin-motion couplings do not appear to impact on the FM order of each state, as shown in Fig. 6.6. In the theoretical curves we do not include other possible errors such as state detection in- e ciency or errors due to spontaneous scattering or uctuations in control parameters. 115 the lth projection of the total e ective angular momentum of the three spins. For the AFM (frustrated) case, we measure the expectation of the symmetric W state witness, WW = (4 + p 5)I^ 2(J^ 2x + J^ 2 y ) [67]. In both cases, as shown in gure 6.7, we nd that entanglement of the corresponding form develops during the adiabatic evolution. In macroscopic systems, the global symmetry in the Ising Hamiltonian (6.1) is spontaneously broken, and ground-state entanglement originating from this symme- try is expected to vanish for the non-frustrated FM case [6]. However, for the frustrated AFM case, the resultant ground state after symmetry-breaking (e.g., j""#i + j"#"i + j#""i ) is still entangled. While spontaneous symmetry-breaking does not occur in a small system of three spins, we can mimic its e ect by adding a weak e ective magnetic eld Bx P i (i) x to the Hamiltonian during the adiabatic evolution[69, 70]. We experimentally observed that the frustrated ground state car- ries entanglement even after global symmetry is broken by using appropriate witness in the Ising model, and thereby establishes a link between frustration and an extra degree of entanglement [59]. In the presence of transverse eld, however, the disentanglement between mo- tional states and spin states becomes imperfect and is accumulated during the adi- abatic evolution. Fortunately, the residual entanglement does not have an in uence on the probabilities of spin product states measured in the direction of the Ising model axis [37]. Therefore we do not see the e ects on the experiments generat- ing the phase diagram shown in section 6.4. The in uence of spin-motion coupling becomes noticeable in the witness measurements, since the motional degrees of free- 116 dom are traced out during the spin state detections. As shown in the blue curves of gure 6.7, the entanglement of the spin states is suppressed because of the remaining spin coupling to motions. 6.6 Scalability of the quantum simulation As the number of spins N grows, the technical demands on the apparatus are not forbidding [59, 57]. In particular, the expected adiabatic simulation time for the spin models is inversely proportional to the ?critical? gap in the energy spectrum; for instance, in a fully-connected uniform ferromagnetic transverse Ising model in a nite-size system, this gap decreases as N 1=3 [71]. Scaling this system to accom- modate long ion chains will allow the investigation of critical behavior depending on the the system size, which is intractable in classical numerical simulation. We perform a benchmarking experiment where all interactions Ji;j are ferro- magnetic regardless of number of spins in the system by tuning the Raman beatnote detuning close to the COM mode. We carefully investigate deviations of experimen- tal simulations from theoretical predictions as the system size increases and discuss possible solutions overcoming the limitations. We observe a crossover from para- magnetic to ferromagnetic spin order, and the crossover sharpens as the number of spins is increased, prefacing the expected quantum phase transition [6] in the thermodynamic limit. We nd that particular order parameters of the system can be quite insensitive to the imperfections of the quantum simulations, and the ex- traction of intensive variables such as the magnetization are much less susceptible 117 to decoherence compared with full tomographic characterizations of the resulting quantum state. In the experiment, we produce the strength of Ji;j close to 1 kHz by setting the Raman beatnote detuning 1 + 4 1 , where 1 are the Lamb-Dicke param- eter and the frequency of COM motional mode, respectively. The strength of the couplings are pretty uniform among the pairs, since the COM mode dominates the interactions. The non-uniformity in the Ising couplings arises from other vibrational modes, which produce around 30% di erences in the strength at the most. We note that for larger detunings, the range of the interaction falls o even further with distance, approaching the limit Ji;j 1=ji jj3 for 1 [81, 72]. The experiment is performed according to the adiabatic quantum simulation protocol, as described in section 1.3. We initially start with strong e ective trans- verse eld By 5NJrms (N = 2, 3, ..., 9) after preparing the ground state of the By Hamiltonian. We transfer it to the Ising Hamiltonian with weak transverse eld by exponentially ramping down By with time constant = 80 s. We observe the evolution of state step by step as we proceed the experiment. For the measurements, we use the PMT and obtain the probability Ps of having s spins in state j"i from a histogram of uorescence counts, constructed by the more than 1000 N times repetition of the experiments [73]. The nal states of the adiabatic evolution are the superposition of two perfect FM states j"" "i and j## #i , called GHZ state, since we implement FM couplings for all the pairs of spins and begin with the ground state of the transverse eld. Therefore we measure the density matrix of the GHZ state as the simulation evolves. We also use other observables such as 118 B y / NJ rms P(F M) (a) (b) P a rity ? (rad) N=2 N=3 N=4 N=5 N=2 N=3 N=4 N=5 N=6 Figure 6.8: Experimental results of adiabatic quantum simulation depending on the system size. Here all pairs of spins have FM in- teractions. (a) The FM order P (FM) evolutions as the system size increases. Initially P (FM) starts 2=2N , since P (FM) is the prob- abilities of two states Pj"" "i and Pj## #i over equally distributed 2N states. As the spin-spin interactions J overpowers By (By ! 0), P (FM) are developed. Here J is the average strength of all inter- actions. The red, orange, yellow, green and blue dots represent the experiments for the total number of spins N=2, 3, 4, 5, and 6, re- spectively. Ideally P (FM) should be close to 1 at the end. However, P (FM) clearly reduces as the number of spins increases in the sys- tem from 2 to 6. The sources and amounts of errors are discussed in the text. (b) The parity oscillations for the nal states of the simula- tion depending on the number of spins, obtained from the population di erence between the even number of j"i state and the odd num- ber of j"i state after applying analysis =2 pulse and swipping its phase . The contrast of the oscillations provide the lower bound of the coherence, the o -diagonal element of the density matrix for the GHZ state (j"" "i j## #i )= p 2. The coherences decrease to 0.8, 0.47, 0.35, 0.27 much faster than P (FM) as the number of spins increase from 2 to 5, because of spin-motional couplings during the simulation as discussed in the text. 119 magnetization to characterize the time evolutions and the phase transitions. We analyze the reliability of the simulation depending on the system size by using the delity of GHZ state, jh SIM jGHZi j2 = 12(P## # + P"" ") + jC## #;"" "j. Here P (FM) = P## # + P"" " and the GHZ coherence jC## #;"" "j is the coe cient of the j## #i h"" "j in the density matrix [68]. We measure coherence of GHZ state by observing the contrast of the oscilating parity signals [Figure 6.8(b)], obtained by applying analysis =2 pulse with di erent phase and taking the dif- ferences in populations of the even number bright states and odd number bright states (P (0) + P (2) + ::: P (1) P (3) :::). Figure 6.8 shows the experimental results of the quantum simulation as the number of spins increases in the system. Initially P (FM) starts 2=2N , since P (FM) is the total probabilities of two states (j"" "i , j## #i ) and the ground state of By P (j)y , j## #i y are equally distributed in a total 2N states in the x-basis. Ideally P (FM) should be close to 1 at the end of the simulation. However, we observe that the nal states are increasingly deviated from the ideal situation as the number of spins grows as shown in gure 6.8(a). We also observe that the GHZ coherence decreases much more rapidly than P (FM) shown in gure 6.8(b). After 6 spins, we did not measure any signi cant GHZ entanglement for the nal state due to the large suppression of the coherence compared to the populations. In the following subsections, we discuss the reason of these deviations and we summarize the expected experimental imperfections in quantum simulations as the system size grows. 120 6.7 Scaling of imperfections 6.7.1 Spin-motion coupling As discussed in section 3.4, the e ective spin-spin coupling Hamiltonian that we assume to describe the system is only valid when the detuning of the beatnote is much larger than the sideband strength. As our detuning is limited by the coher- ence time of the system, i.e. the need for strong coupling strengths, the spin-motion entanglement does not completely vanish during the simulation. In fact, the spin- motion coupling increases in the presence of the transverse eld. The transverse eld mixes the spin states along the axis where the spin-dependent force is applied, therefore this coupling induces phonon excitations, modifying the nal state from the ideal GHZ state to (j"" "i j i j## #i j i )= p 2. Here j i is a mo- tional coherent state and hni = j j2 increases as the amount of state mixing grows. According to numerical calculations corresponding to the experiment, hni increases to 0:5 for the ve spin experiment. The e ective phonon excitation occurs pri- marily in the early stages of the simulation evolution, where the strength of the transverse eld is much larger than the frequency of spin rotations. In the exper- imental conditions, the Raman beatnote is detuned 4 1 from the COM mode, which results in an increment of j j, because of the required large initial transverse eld. However, the population P (FM) and the evolution is not sensitive to these spin-motion couplings. 121 6.7.2 Diabaticity The nite ramping speed of parameters in the Hamiltonian leads to excitations out of the ground state and can lead to oscillations in the observed order parameter. This diabaticity in the evolution, along with errors in the initialization to the original ground state is estimated to suppress the nal value of P (FM) by 4% for N = 5 shown in the orange line of gure 6.9(a). As discussed in section 6.7.2, the diabaticity is related to the minimum gap over the trajectory of the Hamiltonian as well as the ramping time. We note that the gap between the ground and rst-excited state of the fully-connected uniform FM model scales as N 1=3, implying that the simulation time by a factor of ten when the number of spins grows by a factor of one thousand. 6.7.3 Spontaneous Emission One of major error sources is the spontaneous emission from Raman beams which amounts to a 10% spontaneous emission probability per spin in 1 ms for a detuning of 2.7 THz [74]. Spontaneous emission dephases and randomizes the spin state, and thus introduces entropy into the system. In addition, each sponta- neous emission event populates other states outside of the Hilbert space of each spin with a probability of 1=3. Spontaneous emission errors grow with increasing system size, which also suppresses P (FM) order with increasing N , as seen in gure 6.9. We theoretically estimate the suppression of P (FM) due to spontaneous emission by averaging over quantum trajectories and solving density matrix equations to be 5% for N = 2 spins and 13% for N = 5 spins. The error contributions are the 122 same for both populations and coherences. 6.7.4 Intensity uctuations Intensity uctuations on the Raman beams during the simulation induce an uctuating AC Stark shift on the spins. The AC stark shifts produce an imbal- ance on the blue and red sideband detuning, which give rise to imperfections in the spin-spin interactions. According to our numerical calculations, the imbalance uc- tuations of 150 Hz ( 1:5% intensity uctuations in our experimental conditions) can explain the suppressions of P (FM) at the end of the experimental simulations shown in gure 6.9. We investigate the non-uniformity of the laser beams that in- duces di erent AC shifts on the spins by measuring the AC shift on each location, which is introducing position dependent Bz- eld with at most 200 Hz di erence. According to the numerical calculation, the additional small Bz- eld does not no- ticeably suppress P (FM) . 6.7.5 Detection Errors Imperfect spin detection e ciency contributes 5 10 % uncertainties in P (FM). Fluorescence histograms for P (0) and P (1) have a 1% overlap (in detection time of 0.8 ms) due to o -resonant coupling of the spin states to the 2P1=2 level. This prevents us from increasing detection beam power or photon collection time to sep- arate the histograms. In the experiment, the average photon number from a single bright spin is 12. The uncertainty in tting the observed uorescence histograms to 123 B y / NJ rms P(F M) (a) (b) (c) Ground State Experimental ramping + Spontaneous emissions + Intensity fluctuations Data N=5 P(FM) Coherence N=2 N=3 N=4 N=5 Errors Data Diabaticity Spontaneous Emission Intensity Fluctuations Spin - Motion coupling Errors Data Diabaticity Spontaneous Emission Intensity Fluctuations N=2 N=3 N=4 N=5 Figure 6.9: (a) The comparison of experimental results of P (FM) to the theoretical expectations including various imperfections of ex- periments for the 5 spins. The black line represents the evolution of the perfect adiabatic evolution, and the dashed line shows the actual time evolution including the actual ramping of the transverse eld with the imperfections of initial state preparations. The red line is obtained from the theory with spontaneous emissions and the green line is calculated by adding intensity uctuations of 1.5 %. The over- all populations of P (FM) are in agreement with the green theoretical curve including all the above mentioned imperfections. The horizon- tal shift By=J comes from the inaccuracy of the calibration that is not fully understood yet. (b) The amount of errors in P (FM) (the diagonal parts of the density matrix) from the ideal ground states for the case of N=2,3,4 and 5. (c) The amount of errors in coherence (the o -diagonal parts). In both (b) and (c), the green squares represent the total errors measured in the experiment and the bars illustrate the numerically estimated errors from the various sources of experi- mental imperfections. The black, red and green bars show the error amounts from non-perfect adiabatic evolutions, spontaneous emis- sions and intensity uctuations. The blue bars stand for deviations from the spin-motion couplings. The spin-motion coupling reduces the coherence signi cantly as the system size grows, while it dose not have any in uence on the P (FM). determine P (s) increases. The histograms are also a ected by the intensity uctua- tions of detection laser beams and the nite widths. This problem can be eliminated by detecting each spin individually with an imaging detector. Figure 6.9 (a) shows the experimental results and the theoretical calculations including a few steps of imperfections discussed above for N=5 spin case as an exam- ple. In the population P (FM) also shown in gure 6.9 (b), the main deviations come 124 from the spontaneous emissions and the intensity uctuations. As discussed, small amounts of intensity uctuations degrade the performance of the experimental sim- ulations. One of the solutions for those imperfections is to implement a high power laser with a detuning far from the 2P energy levels, which would minimize sponta- neous emission while maintaining the same level of Ising couplings. This would also allow versatility in varying the Ising interaction (together with the e ective external eld) during the simulation, as the di erential AC stark shift between spin states is negligible for a su ciently large detuning. The coherence time increases in the absence of spontaneous emission, allowing for a longer simulation time necessary to preserve adiabaticity as the system grows in size. Recently Raman transitions have been driven using a mode-locked high power pulsed laser at a wavelength of 355 nm, which is optimum for 171Yb+ wherein the ratio of di erential AC Stark shift to Rabi frequency is minimized and spontaneous emission probabilities per Rabi cycle are < 10 5 per spin [51, 75]. Figure 6.9(c) shows the measured and numerically estimated errors in the mag- nitude of coherence, jC## #;"" "j, at the nal state of the quantum simulation. We can clearly see that the errors of the coherence much more rapidly increase than those of P (FM). According to the numerical study, the dominant source for the errors of jC## #;"" "j is the non-vanishing spin-motion couplings in the experimental simulation, shown as the blue area of the gure 6.9(c). In principle, we can elim- inate the e ect of the spin-motion coupling by alternating the transverse eld and Ising interactions [76]. The adiabatic evolution can be discretized by the Trotter 125 expansion written by UTIM( ) = T exp i Z 0 dt0 [HI(t 0) +HB(t 0)] [exp ( iHI( =N)) exp ( iHB( =N))] N ; (6.2) where HI and HB are the Ising Hamiltonian and the transverse eld Hamiltonian, respectively. In the experiment, we can choose =N as the special duration (1/ ), where the spin-motion couplings vanish. We can also reduce the errors of the Trotter expansion by increasing the Raman beatnote detuning from the motional mode. In condensed matter, phase transitions are typically described in terms of order parameters or correlations instead of the density matrices of particular states. We use a absolute magnetization hjmji = P m jm=N jP (m) per site along the Ising direction. Actually we rescale the magnetization hjmji from 0 to 1 regardless of the number of spins to make a fair comparison even for small size systems. We nd that the deviation between experiment and theory for this order parameter does not grow substantially as the system is scaled up in size. Figure 6.10 shows the scaled magnetization, hjmSji for N = 2 to N = 9 spins, showing a nal value of 80% ( gure 6.10) regardless of number of spins. Moreover as shown in gure 6.10(a) we observe the sharpening of the crossover curves from paramagnetic to ferromagnetic spin order with increasing system size. The continued sharpening of this transition is of great interest to the understanding of nite size e ects in phase transitions and can be used to compare various numerical techniques in studying critical phenomena. 126 B y / NJ rms B y / NJ rms ?| m S |? ?|m S |? (a) (b) N=2 N=5 N=9 N = 2 Data N = 9 Data N = 2 Theory N = 9 Theory Figure 6.10: (a) Scaled average absolute magnetization per site, < jmSj > vs B=NJrms is plotted for N = 2 to N = 9 spins. As B=NJrms is lowered, the spin ordering undergoes a crossover from a paramagnetic to ferromagnetic phase. The crossover curves sharpen as the system size is increased from N = 2 to N = 9, anticipating a QPT in the limit of in nite system size. The oscillations in the data arise due to imperfect initial state preparation and non-adiabaticity due to nite ramping time. (b) Magnetization data for N = 2 spins (circles) is contrasted with N = 9 spins (squares). The data deviate from unity at B=NJrms by 20%, predominantly due to spontaneous emissions in Raman transitions and intensity uctuations of Raman laser beams, as discussed in the text. Here, the theoretical time evo- lution curves (red line for N = 2 and black line for N = 9 spins) are calculated by averaging over 10,000 quantum trajectories. 127 6.8 Conclusion and Outlook Trapped atomic ions represent a promising platform for the quantum sim- ulation of intractable Hamiltonian systems. There have been several theoretical proposals in this direction, largely following schemes in the realm of quantum com- puting, and recent experiments have shown that the system can be scaled to a degree where all classical simulations become impossible [59, 70, 77]. This paper has shown both theoretically and experimentally that the quantum simulation of quantum magnetism and the emergence of spin order can be controlled through external laser beams up to 9 spins and can further be scaled to much larger num- bers of spins. The stable con nement of larger numbers of ions may require novel ion trap architectures such as anharmonic axial potentials [78] for a linear chain or two-dimensional trap geometries [79, 80], but there are no known fundamental limitations in this scaling. As discussed here, for a xed level of total laser power, the errors associated with decoherence from spontaneous Raman scattering from the lasers is expected to grow only as N1=3 for the linear chain, holding errors from phonon creation and diabatic transitions to excited states at xed levels. Alterna- tively, all of these errors can be held at a xed value independent of N so long as the laser power increases by N1=3. In either case the required time for adiabatic ramping grows as N1=3, so slowly drifting errors such as (real) magnetic elds and motional heating of the ions must be kept under control for very large N as discussed in the supplementary information of Ref. [59] This system can also be extended to Heisenberg or XYZ spin models [81] or 128 spin-1 systems by adding a few more laser beams. As the system grows, the trans- verse motional modes that mediate the Ising couplings can give rise to higher levels of frustration and complex phases of magnetic ordering. For instance, by preparing a ground state of a highly frustrated collection of trapped ion spins, it should be possible to create localized topological excitations and guide their transport through the system [82]. This example of topological matter is of great interest for the ro- bust representation and manipulation of quantum information [83, 84]. But more generally, the trapped ion system is poised to be the rst to determine ground state features of Hamiltonians where no solution can be obtained otherwise. 129 Chapter 7 Quantum Simulation of the Devil?s Staircase 7.1 Introduction Recent e orts in quantum simulation, motivated by its promise as a tool for advancing understanding of condensed matter systems, solving di cult optimization problems, and simulating other physical systems while providing an in-situ insight previously unattainable have seen signi cant advances. These include simulation of the nearest -neighbor anti-ferromagnetic Ising model [85], and the long-range anti-ferromagnetic Ising model. Recent theoretical interest in the simulation of the complete Devil?s Staircase [86] using cold, trapped ion quantum simulators [87] have motived us to simulate this system for six and ten spins. In this chapter I report a quantum simulation experiment1, similar to those in chapter 6. In this experiment a linear chain of trapped, cold ions are irradiated by lasers far detuned from resonant transitions to simulate and control an Hamiltonian of AFM coupled Ising spins with long range, frustrated interactions. We add to this e ective interaction a combined axial and transverse simulated elds, and by scanning the strength of the axial eld we explore N/2 phase transitions in the ground state of the zero transverse eld spin chain of N spins. Site resolved imaging, as described in section 4.4.8 allows us to extract ground state order parameters and 1Manuscript in preparation. 130 state probabilities, ampli ed over the background by the post-processing techniques of section 4.4.8. We compare the experimental results to theoretical predictions. 7.2 Overview We perform a quantum simulation of the AFM Ising model with a tunable axial biasing eld Ba: H = NX i;j Jij i x j x +Bx NX i ix +By NX i iy (7.1) Where H is the simulated Hamiltonian of the system, ix the Pauli spin operator on spin i, and Ji;j is the spin-spin coupling between spin i and spin j as de ned in 3.22. The j"i and j#i states are de ned as presented in 2.4, and the ions are radially con ned with a transverse collective COM mode frequency !1 = 2 4:863 MHz , and a tilt frequency !2 = 2 4:813 MHz . This Hamiltonian has been Identi ed as a physical model for several real world systems, such as graphite intercalation compounds, and exhibits the much investigated complete Devil?s staircase behavior for a macroscopic number of spins [88]. Just as in the experiments described in chapter 6, we will initialize the spin system in the ground state of a trivial Hamiltonian, where the magnetic elds over- whelm the J component of the Hamiltonian. Unlike the simulations of chapter 6, we cannot simply rotate the spins to align with transverse eld By, rather we will have to apply a Raman pulse that rotates the spins into the superposition of Bx + By. Thus, for each value of Bx the phase of the initialization pulse relative to the simu- 131 lation frequency phases will be di erent. For our selected simulation pulse phases, the phase of the initialization pulse will be = 1800 tan 1(Bx=By) (7.2) As all components of J are positive for our red beatnote detuning from the transverse COM, initializing in the ground state (along the total B eld) will simulate AFM coupling. After initializing, we will apply all components of the Hamiltonian simultane- ously, and proceed to ramp down By with an exponential envelope (with no o set). The duration of the time constant of the envelope is a compromise - the ideal time constant would satisfy the adiabatic condition [89] with a large safety margin. How- ever, we are constrained by the coherence time of the ions, which for us was found to be roughly 3 ms . Therefore, we chose our time constant so as to achieve as low a B- eled as possible in the available time without causing severe oscillations or excitations towards the end of the simulation. Finally, we apply a Raman rotation to measure the spin state in the mea- surement basis (the z-axis of the Bloch sphere). Repeating this experiment with di erent Bx values allows us to map the phase diagram of this Hamiltonian. 7.3 Experiment As described in section 3.4, we control the dynamics [57] by applying far detuned global Raman beams with a wavenumber di erence k along the transverse 132 Coupling Strength J1;2 660 Hz J1;3 368 Hz J1;4 235 Hz J1;5 158 Hz J1;6 106 Hz J2;3 621 Hz J2;4 373 Hz J2;5 242 Hz J2;6 158 Hz J3;4 609 Hz J3;5 373 Hz J3;6 235 Hz J4;5 621 Hz J4;6 368 Hz J5;6 660 Hz Table 7.1: Calculated spin-spin couplings for 6 ions with our trap, detuning, and laser intensity parameters (perpendicular to chain alignment - \x" direction) direction of the ion chain. To create the the spin-spin coupling matrix J, we detune two Raman beams 80 KHz to the blue and red of the COM mode, so that the beatnote is red of the COM mode. This is done while verifying that the strengths of the beams drive sideband transitions on a single ion with equal strength, are symmetrically detuned from the carrier transition, and the beams are centered on the center of the spin chain (the position of a single ion in the trap). To diagnose the intensity of our 355 nm center wavelength 80 MHz pulsed laser Raman beams, we measure a two ion spin-spin coupling strength of 66.5 KHz . As- suming equal intensity across our ion chain, the spin-spin coupling matrix J is calculated and presented in table 7.1. The single ion RSB or BSB strength for this intensity on resonance is calculated to be 23 KHz , so our detuning 3 . 133 2 3 4 5 6 0.0 0.1 0.2 0.3 0.4 0.5 0.6 Ion Index i J 1 , iHKH zL Spin-Spin Coupling Decay Figure 7.1: Power law t for the for the rst ion. The curve was tted with only the exponent as a parameter. As in our other experiments where we tuned to the red of the COM, the couplings are decaying according to a power law with exponent Ji;j Ji;i+1 ji jj (7.3) This exponent depends only on trap conditions - it is a feature of the motional mode bandwidth. For our current trap conditions and six ions, = 0:976, as shown in gure 7.1. In previous work, we have explored the e ect of changing this parameter on the correlations of the nal AFM state for Bx = 0.2 characterizes the range of the spin-spin coupling, and for longer range AFM interactions we can expect to see more frustration in the ground state with smaller energy gaps. 2Accepted for publication in \Science", R. Islam et al Emergence and Frustration of Magnetism with Variable-Range Interactions in a Quantum Simulator 134 The characteristic power law decay arises due to the non-negligible contribu- tion of the tilt mode as well as smaller contributions from the remaining trans- verse x-modes. y-modes are suppressed by beam alignment and control of the trap principal-axes, so that the y-mode sidebands are at least 10 times suppressed com- pared to the x-modes. We further suppress the y COM mode by lowering its fre- quency relative to the x COM mode, by adjusting trap anisotropy. As described in 4.4.6, to generate the Raman beams we use two lasers, mod- ulated by two AOMs. The rst AOM is driven by an oscillator stabilized by an error signal from the beating of a Microwave source and the frequency component of the pulsed laser near the qubit frequency. The second AOM is driven by an AWG. During the simulation, the AWG drives the AOM with three di erent frequencies: one to drive the qubit Rabi oscillations, and two symmetrically detuned from this central frequency to drive the M lmer-S renson spin-spin coupling gate. It is only when these frequencies interfere on the ion with the light modulated by the rst AOM that transitions are driven. The AWG and the driver of the stabilized AOM are phase locked to a Rubidium clock, outputting a reference signal at 10 MHz . We select and de ne the Pauli vector component via the relative phase of the di erent beatnotes. We set the phases as elucidated in table 7.2 As demonstrated in section 3.4 it is only the phase di erence of the BSB and RSB im that a ects the M lmer-S renson spin-spin coupling. Therefore, we arbitrarily set the phase of the RSB to 0, and set only the phase of the BSB to control the axis of the interaction. As im is equal to half the di erence of BSB and 135 Pulse Phase Result Initialization Bx xi +By y i Bx 900 xi By 00 y i RSB 00 Combined with BSB we get BSB 1800 Jij xi x j Final 900 yi Table 7.2: Pulses for axial simulation experiment. Initial pulse is de ned in 7.2. For example, with no axial eld = 1800. The nal rotation of the spins into the measurement basis can alternatively be set to 900, as this will only result in ipping the spins in the nal state and does not change the physics. RSB phases, setting to 1800 sets the interaction axis to be the x-axis. The control panel for the AWG chapter phases, frequencies and amplitudses is shown in gure 7.2. All the lasers used in the experiment are global - we do not use individual addressing. This simpli es optical setup and reduces sensitivity to beam steering, however this prevents us from measuring the elements of the density matrix that require operations that are not a global rotation. The experiment proceeds as follows. The ions are Doppler cooled for 3 ms , and optically pumped to the j#i state by a 8 s laser pulse resonant with the 2S1=2 jF = 1i ! 2P1=2 jF = 1i transition, as shown in gure 2.3. 30 cycles of alternating red-sidebands resonant with the x COM mode and 3 s long optical pumping pulses reduce the occupation number of this phonon mode to 0:1 phonons, so that state is initialized in the spin as well as motion sub-spaces. We then rotate our spins to the trivial ground state of a Hamiltonian domi- nated by a transverse magnetic eld, where our initial transverse eld By is 5 times 136 Figure 7.2: Control panel for quantum simulation AWG waveform in the Labview control program. The phase of J is the phase of the BSB. 137 larger than the largest spin-spin coupling JMAX . We ramp down the transverse eld By with an exponential time constant of 600 s . This value was found by numerically evolving the Schr odinger equation (SE) for our experimental parameters and maximizing the resemblance of the mag- netization order parameter of the evolving state to that of the adiabatic perfect ground state of Hamiltonian 7.1 at time t of the simulation. The magnetization order parameter is de ned here as M = 1=2 + nj"i nj#i 2N . When the ramping is complete, we rotate our spins back to the measurement basis, and measure each spin using a resonant beam with a Princeton PI-MAX 3 ICCD with a 3ms exposure time, as described in 4.4.8. Thus, we are able to measure the distribution of spin eigenstates for this process, with a 93% readout delity per ion. This delity is reduced from the theoretically attainable delity of 99:5% by electronic readout noise and the 0.23 numerical aperture of the imaging system. In order to correct the biasing of the state probabilities, we redistribute them as described in 4.4.8. For 6 ions, we map out a phase diagram by stopping the simulation at three di erent nal times of 306 s , 960 s and 3 ms corresponding to By=JMAX values of 3:00; 1:01 and 0.034. We measure the nal state 4000 times for 61 equally spaced values of Bx=JMAX between 0 and 4. Figure 7.3 displays the phase diagram of the magnetization order parameter. The surface plot is the perfect adiabatic order parameter for the ground state, whereas the points are experimental results. The theoretical plot shows three rst-order [88] phase transitions at By=JMAX = 0. We have also attempted to repeat the experiment for points close to the phase transition 138 0 0.5 1 Magnetization Figure 7.3: Magnetization Phase Diagram for 6 Ions. 139 once our measurements indicated their location. However, our control of the system parameters did not seem to be stable enough over the duration of the data taking. Our dominant measured states at By=JMAX 0 are displayed in gure 7.4 and further investigated. Comparing only the dominant states to theoretical evolution of Schr odinger equation (solid lines) we see the data for the dominant states closely follows. However, the simulation is not su ciently adiabatic and the data is too noisy to clearly show the sharpness of the phase transitions. This is accentuated when we compare the magnetization parameter for the adiabatic solution, the evolved solution and the data, as in gure 7.5. The lack of adiabaticity at the phase transition can be investigated theoreti- cally from multiple aspects. One aspect is that of the adaibaticity parameter, which for our system roughly corresponds to the inverse of the minimum energy gap be- tween the ground state and the rst coupled excited state for the trajectory of the simulation [22]. For the Hamiltonian 7.1, there are eigen-energy crossings zero at the phase transitions, as can be seen in gure 7.6 and 7.7. Another interesting feature of the phase transitions is its relation to an in- verse pseudo-temperature. Currently, there is ongoing theoretical and experimental investigation of thermalization in closed, coherent quantum systems [90]. Accord- ing to the currently developing Eigenstate Thermalization Hypothesis (ETH)3 [91], some classes of closed quantum systems display steady-state observables that corre- spond to the micro-canonical expected observables for analogous classical systems. For the Hamiltonian at hand, we will see that in the low axial eld limit the state 3I would like to thank Chao Shen for introducing me to this work. 140 0 1 2 3 4 0.0 0.2 0.4 0.6 0.8 1.0 Bx JMAX Probabilit y States vs Bx Probabilit y fl ? fl fl ? fl fl fl fl fl fl fl fl ? fl ? fl ? + ? fl ? fl ? fl fl fl ? fl ? fl + fl ? fl ? fl fl fl fl fl ? fl fl + fl fl ? fl fl fl fl fl fl fl fl ? + ? fl fl fl fl fl Figure 7.4: Results and simulation of dominant states. The phase transitions in the adiabatic limit are marked with dashed lines. The non-adiabaticity causes the states to bleed out of their phase in the adiabatic case. Some excited states are populated signi cantly. The red and green curves and data are noticebaly excited states for their respective regimes. The rst excited state for each stair is a kink in the ground state - the spins are all ipped relative to the ground state at a point in the spin chain that requires the least energy. Typically this is close to the edge chain, as the couplings there are weakest. 141 0 1 2 3 4 0.0 0.2 0.4 0.6 0.8 1.0 Bx JMAX Magnetizatio n Figure 7.5: 6 ion staircase. Black solid line is perfect adiabatic magne- tization, red solid line is numerically integrated Schr odinger equation, and dots are calculated from data. 142 0 1 2 3 4 - 10 - 5 0 5 10 15 20 Bx JMAX EHKH zL Figure 7.6: Complete eigen-energy spectrum for our Hamiltonian. On the left hand side, the pure AFM case, the lowest energy states are the AFM states, which are degenerate, and the highest states are the FM states. As the axial bias is increased, the FM states degeneracy is lifted. The FM state counter aligned to the eld increases in energy and remains the highest, while the FM state aligned with the eld becomes the lowest energy state at high axial eld limit. At the high axial eld limit, the energy states converge into 6 degenerate levels, corresponding to number of spins aligned with the eld. (a) (b) 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5- 8 - 7 - 6 - 5 - 4 - 3 - 2 Bx JMAX EHKH zL 0 1 2 3 4 0.0 0.1 0.2 0.3 0.4 0.5 Bx JMAX DE HKH zL Figure 7.7: (a) Low lying energy levels and and crossing points. The dashed lines mark the energy crossings. (b) Energy di erence between the two lowest lying states. These lead to the rst-order transitions seen in 7.5 143 probabilities will be occupied with a probability that is roughly proportional to an exponential function of the energy of the eigenstate: Pi = e (Ei E0)T (7.4) Where E is the energy of the state, E0 is a normalizing factor, and T is the inverse pseudo-temperature. By tting to the simulated histograms, we can extract T for the nal states at all Bx values and By = 0, as shown in gure 7.8. For six ions, this treatment is not very convincing, as only the rst peak aligns itself with the rst phase transition. Nevertheless, I will use this treatment to de ne a pseudo- temperature lowering lter. The lter will succeed to amplify the dominant ground state in the regime where the pseudo-temperature picture is not valid as well, as long as the most dominant state is the true ground state - a reasonable assumption when By = 0. We de ne an amplifying parameter A, which for A=1 leaves the probabilities unchanged, and for increasing A will suppress the weaker states compared to the most dominant states. For each state probability for a given Bx and By, we de ne a new probability P 0i = P A i . Then PN i=1 P 0 i is renormalized to 1. The e ect can be compared to removing excess energy from the system, as can be seen in gure 7.10. In this gure, we compare the e ect of increasing A for a simulated nal evolved state to increasing the total simulation time while keeping the ratio of simulation time to exponential decay time constant xed. Starting with a simulated state from the evolved Schr odinger equation with 0.8 s time constant and our experimental 144 (a) (b) -2 0 2 4 10-7 10-6 10-5 10-4 0.001 0.01 0.1 EHKHzL Probabilit y -10 -5 0 5 10 15 20 10-8 10-6 10-4 0.01 EHKHzL Probabilit y (c) 0 1 2 3 4 0 2 4 6 8 Bx JMAX Invers e Temperatur eH10 -3 K B- 1 L Figure 7.8: (a) Exponential t of probability of state as a function of state energy, for Bx = 0. The t slope is the inverse pseudo- temperature. (b) Exponential t for Bx=JMAX = 4. (c) The temper- ature coe cient for all ts. The t fails rather quickly after showing a correspondence to the rst phase transition. Phase transitions are marked with dashed lines. J, we gradually increase A, and then compare the individual states for increasing Alpha and decay time. Scaling the A axis for all states and values of Bx equally, we see that for most values of Bx the ampli ed and the slowed evolution states match well. Applying the lter to our data with increasing ampli cation (or \cooling"), we can retroactively slow down the simulation. Figure 7.9 demonstrates this e ect. We repeat the same experiment, this time for 10 ions, and only for a nal simulation time of 3 ms . Our initial value of By is set to 3Jmax, and we perform the 145 0 1 2 3 4 0.0 0.2 0.4 0.6 0.8 1.0 Bx JMAX Magnetizatio n A=1 0 1 2 3 4 0.0 0.2 0.4 0.6 0.8 1.0 Bx JMAX Magnetizatio n A=2 0 1 2 3 4 0.0 0.2 0.4 0.6 0.8 1.0 Bx JMAX Magnetizatio n A=3 0 1 2 3 4 0.0 0.2 0.4 0.6 0.8 1.0 Bx JMAX Magnetizatio n A=4 Figure 7.9: Power lter with increasing strength on 6 ion magnetiza- tion for Bx = 0. When lter strength is A=1, the state probabilities are unchanged. When A=4, the adiabatic magnetization staircase emerges from the data. adiabatic simulation procedure for Bx=JMAX = 0! 5:5, in step of 0.05, and repeat each simulation 4000 times. As J is of the same order of magnitude and range as before, as shown in gure 7.11, we can expect to see smaller energy gaps. As we cannot extend simulation time, the expected e ect will be much worse diabaticity, i.e. stronger excitation of low lying energy states. We can also expect a lower ratio of signal to noise, as decoherence and detection error for overall state will increase for a larger number of spins. This can be seen in a comparison of our detected states to simulated states in gure 7.12. Here we must rescale the y-axis for the data to compensate for di erent scales for data and theory. Interestingly, for 10 ions the pseudo-temperature treatment works far better. As seen in gure 7.13, the exponential t lies next to many more points, improving 146 1 2 3 0.30 0.35 0.40 0.45 0.50 1 2 3 t HmsL Probabilit y o fStat e A fl ? fl ? fl ? , Bx JMAX = 0 1 2 3 0.00 0.02 0.04 0.06 0.08 1 2 3 t HmsL Probabilit y o fStat e A fl ? fl ? ? fl , Bx JMAX = 0 1 2 3 0.20 0.22 0.24 0.26 1 2 3 t HmsL Probabilit y o fStat e A fl fl fl ? fl fl , Bx JMAX = 2.1 1 2 3 0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 1 2 3 t HmsL Probabilit y o fStat e A fl fl ? fl ? fl , Bx JMAX = 2.1 Figure 7.10: Figure above demonstrates e ect of power lter versus improved adiabaticity, for two states for two values of Bx. Red line is power lter on 0.8 s evolved state. Blue line is evolved state with increasing exponential decay constant , with xed ratio of Tfinal . For some values of Bx this procedure will fail, as close to a phase transition there may be population oscillations. 147 (a) (b) 2 4 6 8 10 0.0 0.1 0.2 0.3 0.4 0.5 0.6 Ion Index i J 1 , iHKH zL Spin-Spin Coupling Decay 0 5 10 i j 0.0 0.2 0.4 0.6 Ji, j KHz Figure 7.11: (a) Power law decay for 10 ions, shown for spin 1. (b) The J coupling matrix. Note that it is symmetric across the chain and across the diagonal. The diagonal elements are meaningless. the validity of the t. When we extract all the pseudo-temperature coe cients for the Bx domain of the simulation, we see more temperature cusps corresponding to the adiabatic regime phase transitions. Applying the same ltering method to our 10 ion magnetization data, we see how the staircase emerges again. However, for this data the ltering is much more aggressive, as could be expected for a more diabatic simulation. It is less successful as well, which is likely due to increased overall noise and decoherence [92]. 7.4 Conclusion Despite the technical challenges faced by this experiment, the sharp phase transitions at near zero transverse eld can be observed. Here we rely on the fact that when our transverse eld is extinguished, the dominant state will be the ground state. As can be seen in gure 7.14, the staircase signal can be extracted from the 148 0 1 2 3 4 50.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0 0.1 0.2 0.3 Bx JMAX Pro bHSimulatio nL States vs Bx Pro bHDat aL Pro bHSimulatio nL fl ? fl ? fl fl ? fl ? fl fl fl ? fl fl fl fl ? fl fl fl fl fl fl fl fl fl fl fl fl ? fl ? fl ? fl ? fl ? fl + fl ? fl ? fl ? fl ? fl ? fl ? fl fl ? fl ? fl ? fl + fl ? fl ? fl ? fl fl ? fl fl ? fl fl fl ? fl fl ? fl + fl ? fl fl ? fl fl fl ? fl fl fl ? fl fl ? fl fl ? fl + fl ? fl fl ? fl fl ? fl fl fl fl fl ? fl fl fl ? fl fl + fl fl ? fl fl fl ? fl fl fl fl fl fl fl fl ? fl fl fl fl + fl fl fl fl ? fl fl fl fl fl fl fl fl fl fl fl fl fl fl ? + ? fl fl fl fl fl fl fl fl fl Figure 7.12: Results and simulation of dominant states. The phase transitions in the adiabatic limit are marked with dashed lines. The non-adiabaticity causes the states to bleed out of their phase in the adiabatic case. Dashed lines mark phase-transitions, found from en- ergy crossings. Here the y-axis of the data and theory are to di enret scales, as the data is suppressed by noise and increased decoherence [92] 149 (a) (b) -2 0 2 4 6 10-6 10-4 0.01 EHKHzL Pro b Prob VS EHKHzL, Bx = 0. -5 0 5 10 10-10 10-8 10-6 10-4 0.01 EHKHzL Pro b Prob VS EHKHzL, Bx = 2.2 (c) (d) -15 -10 -5 0 5 10-10 10-8 10-6 10-4 0.01 1 EHKHzL Pro b Prob VS EHKHzL, Bx = 4.95 0 1 2 3 4 5 0 2 4 6 8 Bx JMAX Invers e Temperatur eH10 -3 K B- 1 L Figure 7.13: (a) Exponential t of probability of state as a function of state energy, for Bx = 0. The t slope is the inverse pseudo- temperature. (b) Exponential t for Bx=JMAX = 2:2. The higher energy states are beginning to separate into energy bands. (c) Ex- ponential t for Bx=JMAX = 4:95. The higher energy states are now strongly separated into gaps, corresponding to the number of spins aligned with the axial eld. This causes the psuedo-temperature pic- ture to fail.(d) The temperature coe cient for all ts. This time the t shows reliably the phase transition points, and only followng the second to last transition. Phase transitions are marked with dashed lines. 150 0 1 2 3 4 5 0.0 0.2 0.4 0.6 0.8 1.0 Bx JMAX Magnetizatio n A=1 0 1 2 3 4 5 0.0 0.2 0.4 0.6 0.8 1.0 Bx JMAX Magnetizatio n A=6 0 1 2 3 4 5 0.0 0.2 0.4 0.6 0.8 1.0 Bx JMAX Magnetizatio n A=11 0 1 2 3 4 5 0.0 0.2 0.4 0.6 0.8 1.0 Bx JMAX Magnetizatio n A=16 Figure 7.14: 10 ion data power ltered with power A, increasing from no ltering to A=16. A=16 plot has a solid black curve that is the adiabatic ground state solution. Dashed lines are phase transition location for adiabatic ground state. 151 data. However, this lter will fail to amplify the correct states close to a phase transition, as the exact ground state will not dominate for all simulation times and diabatic parameters. As quantum simulators are scaled to higher numbers of spins and more complex Hamiltonians, extraction of the desired signal[93] will be challenged by increased de- coherence [92], readout error for the total state, and diabatic e ects. However, these issues may be overcome by data processing methods catered to the Hamiltonian sim- ulated and guided by theoretical understanding of certain classes of Hamiltonians. This work was supported by the US Army Research O ce through funds from the DARPA Optical Lattice Emulator Program and the IARPA MQCO Program; the NSF Physics at the Information Frontier Program, and the NSF Physics Frontier Center at JQI. 152 Chapter 8 Simulating the Ising Model with Arbitrary Control of the Couplings 8.1 Introduction In previous chapters I have discussed work were we applied global spin- dependent optical dipole forces to generate trivial forms of the spin couplings, such as a uniformly decaying ferromagnet or anti-ferromagnet. Now I will present my pro- posal1 for how to tailor optical forces to generate arbitrary fully-connected networks of N spins that uniquely specify each of the N(N 1)=2 pairwise interactions. The scheme is independent of the spatial geometry of the ion crystal and is compatible with one-dimensional arrays of trapped ions used in current experiments. We start with the arbitrary fully-connected Ising Hamiltonian on N spins, H = X i> i;m i), so that the phonon states can be adiabatically eliminated, leaving the pure spin-spin coupling above [3, 33, 57]. 8.2 Control of an Arbitrary Lattice Hamiltonian The above expression has N +1 control parameters in the set of Rabi frequen- cies f ig and the global beatnote detuning . In order to generate an arbitrary Ising coupling matrix Ji;j however, it is necessary to have at least N(N 1)=2 independent controls [103]. Additional control parameters can be introduced by adding multiple spectral beatnote detunings to the Raman beams, one near each motional mode (see Fig. 8.1), with a unique pattern of spectral components on each ion. There are several ways to achieve this, all involving some form of individual ion addressing. For simplicity, we retain the same set of N Raman beatnote detunings m on each 154 ? 2 ? N ? 1 frequency ? N ? 1 ?i,N ?i,1 Figure 8.1: Spectrum of transverse mode frequencies !m for N = 10 ions in an anisotropic harmonic linear trap (solid lines), with the highest mode frequency !1 corresponding to center-of-mass motion. Raman beatnote detunings m from the qubit frequency !s are de- noted by the N = 10 dashed lines, with each spectral feature near a given motional sideband. The height of the dashed lines represents the intensity of each beatnote for ion i. In general each ion will be illuminated with a di erent set of intensities. ion and allow the spectral amplitude pattern to vary between ions, all characterized by the N N Rabi frequency matrix i;n of spectral component n at ion i. Note that the relative signs of the Rabi frequency matrix elements can be controlled by adjusting the phase of each spectral component. This individual spectral amplitude addressing provides N2 control parameters, and the general Ising coupling matrix becomes Ji;j = NX n=1 i;n j;n NX m=1 i;m j;m!m 2n !2m (8.2) NX n=1 i;n j;nFi;j;n; (8.3) where Fi;j;n characterizes the response of Ising coupling Ji;j to spectral component n. An exact derivation of the e ective Hamiltonian given a spectrum of spin-dependent forces gives rise to new o resonant cross terms, which can be shown to be negligible 155 in the rotating wave approximation, as long as the bandwidth is xed and the transverse center of mass mode is set to a frequency smaller than twice the transverse zig-zag mode frequency. Thus sums and di erences of beatnotes do not directly encroach any sideband features in the motional spectrum of the crystal [101]. We tune each beat note frequency near a unique normal mode so that Fi;j;n has independent contributions for each n. Given a desired Ising coupling matrix Ji;j, we use standard constrained nonlinear optimization to nd the corresponding Rabi frequency matrix i;n, while minimizing the total beam intensity. The deviation between the desired and the attained coupling was less than typical round-o errors. However, Eq. 8.3 depends nonlinearly on the Rabi frequencies, and it is not clear that a solution exists or how the resulting total optical power scales with N . We rst describe a formal method to invert Eq. 8.3, showing the existence of a solution. For simplicity we force the control matrix to be lower triangular, or set i;n = 0 for i < n, with i;i !s (this still leaves N(N 1)=2 independent parameters). Isolating the n = i term in the sum of Eq. 8.3, we nd j;i = Ji;j P n 4 #include 5 #include 6 #include "AWG dll . h" 7 #include " da11000 d l l impor t . h" 8 9 const double num=2 3.14159265358979323846/1000; / f o r speed , f requnecy in MHz / 10 11 struct SegmentStruct //need to b u i l d an array o f t h e s e 12 f 13 DWORD SegmentNum ; // Current Segment Number 14 unsigned short SegmentPtr ; // Pointer to curren t user segment 15 DWORD NumPoints ; // Number o f po in t s in segment 16 DWORD NumLoops ; // Number o f t imes to repea t segment ( a p p l i e s to next segment ) 17 DWORD BeginPadVal ; // Pad va lue f o r beg inn ing o f t r i g g e r e d segment 18 DWORD EndingPadVal ; // Pad va lue f o r ending o f t r i g g e r e d segment 19 DWORD TrigEn ; // I f > 0 then wai t f o r t r i g g e r b e f o r e going to next 170 segment . 20 DWORD NextSegNum ; // Next segment to jump to a f t e r complet ion 21 // o f curren t segment a c t i v i t i e s 22 g ; 23 24 struct wave 25 f 26 double f ; 27 double phase ; 28 double amp ; 29 g ; 30 31 struct l inear param 32 f 33 34 double f1 , p1 , f2 , p2 , f3 , p3 ; 35 f loat a1 , a2 , a3 ; 36 double s l ope ; 37 g ; 38 39 struct sb param 40 f 41 unsigned long modes ; 42 double f ; 43 double durat ion ; 44 unsigned long OPtime ; 45 unsigned short p u l s e s ; 171 46 g ; 47 48 int make64x (unsigned long s i z e ) 49 f 50 d i v t r e s u l t=div ( s i z e , 6 4 ) ; 51 i f ( r e s u l t . rem>0) 52 return(64 r e s u l t . rem) ; 53 else 54 return (0 ) ; 55 g 56 57 d e c l s p e c ( d l l e x p o r t ) long s i d e b a n d c o o l i n g s e q u e n t i a l (unsigned short a , unsigned long modes , double f , 58 unsigned long duration , unsigned long OPtime , unsigned short p u l s e s ) 59 f 60 unsigned long i , j ,m, t =0; 61 62 for ( i =0; i

0 then wai t f o r t r i g g e r b e f o r e going to next segment . 199 DWORD NextSegNum ; // Next segment to jump to a f t e r comple t ion 200 // o f curren t segment a c t i v i t i e s 201 g ; / 202 203 204 for ( i =0; i <(unsigned long ) ( S) >NumOfSegments ; i++) 205 f 206 for ( j =0; j <(unsigned long ) ( ( S) >segment [ i ] . SegmentPtr ) > WaveSize ; j++) 207 f 208 a [ t ]=( ( S) >segment [ i ] . SegmentPtr ) >wave [ j ] ; 209 t++; 210 g 211 g 212 return ( (unsigned long ) ( S) >NumOfSegments ) ; 213 g 214 215 d e c l s p e c ( d l l e x p o r t ) unsigned long BuildDA11000Array ( SegmentHdl S , SegmentStruct a ) 216 f 217 unsigned long i ; 218 / 219 s t r u c t SegmentStruct //need to b u i l d an array o f t h e s e 178 220 f 221 DWORD SegmentNum ; // Current Segment Number 222 unsigned shor t SegmentPtr ; // Pointer to curren t user segment 223 DWORD NumPoints ; // Number o f po in t s in segment 224 DWORD NumLoops ; // Number o f t imes to repea t segment ( a p p l i e s to next segment ) 225 DWORD BeginPadVal ; // Pad va lue f o r beg inn ing o f t r i g g e r e d segment 226 DWORD EndingPadVal ; // Pad va lue f o r ending o f t r i g g e r e d segment 227 DWORD TrigEn ; // I f > 0 then wai t f o r t r i g g e r b e f o r e going to next segment . 228 DWORD NextSegNum ; // Next segment to jump to a f t e r comple t ion 229 // o f curren t segment a c t i v i t i e s 230 g ; / 231 for ( i =0; i <(unsigned long ) ( S) >NumOfSegments ; i++) 232 f 233 a [ i ] . BeginPadVal=( S) >segment [ i ] . BeginPadVal ; 234 a [ i ] . EndingPadVal=( S) >segment [ i ] . EndingPadVal ; 235 a [ i ] . NextSegNum=( S) >segment [ i ] . NextSegNum ; 236 a [ i ] . NumLoops=( S) >segment [ i ] . NumLoops ; 237 a [ i ] . SegmentNum=( S) >segment [ i ] . SegmentNum ; 238 a [ i ] . TrigEn=( S) >segment [ i ] . TrigEn ; 239 a [ i ] . NumPoints=( ( S) >segment [ i ] . SegmentPtr ) >WaveSize ; 240 a [ i ] . SegmentPtr =( ( S) >segment [ i ] . SegmentPtr ) >wave ; 241 g 242 return ( (unsigned long ) ( S) >NumOfSegments ) ; 243 g 244 179 245 d e c l s p e c ( d l l e x p o r t ) bool LoadSegmentHdl ( SegmentHdl S) 246 f 247 bool e r r o r=fa l se ; 248 SegmentStruct a= new SegmentStruct [ ( S) >NumOfSegments ] ; // don ? t f o r g e t d e l e t e [ ( S) >NumOfSegments ] 249 BuildDA11000Array (S , a ) ; 250 // INITIALIZE BOARD 251 da11000 SetTriggerMode (1 , 0 , 0 ) ; 252 da11000 CreateSegments (1 , 1 , ( S) >NumOfSegments , a ) ; 253 254 delete [ ( S) >NumOfSegments ] a ; 255 return ( e r r o r ) ; 256 g 180 Appendix D Mathematica ROI and Discriminator Selection Script The following script is called after taking 1000 shots of the ions pumped dark and then 1000 shots of the ions rotated to bright. It is called from the Labview control program automatically after the shots have been taken, run as a command line script with the following parameters - \dark.csv" - \bright.csv" - Nions - Nimages - cuto (0 - 1) The script (\ROI auto selection.m")is created and run from the directory of the data les. example : math - script \ROI auto selection.m" \dark Shots.csv" \bright Shots.csv" 4 1000 0.5 DarkDataT = Transpose[Import[$CommandLine[[4]]]];ort[$CommandLine[[4]]]] dataT = Transpose[Import[$CommandLine[[5]]]]; Nions = ToExpression[$CommandLine[[6]]];oExpression[$CommandLine[[6]]]oExpression[$CommandLine[[6]]] MaxDataT = Max[dataT]; Nimages = ToExpression[$CommandLine[[7]]];oExpression[$CommandLine[[7]]] width = Dimensions[dataT][[2]]=Nimages; height = Dimensions[dataT][[1]]; This section sums all the frames and then normalizes the image. It also cal- culates the average value of both the dark and bright pixels. NormImageDark = Sum[DarkDataT[[All; widthx+ 1;;(x+ 1)width]]; fx; 0; Nimages 1g]; NormImage = Sum[dataT[[All; widthx+ 1;;(x+ 1)width]]; fx; 0; Nimages 1g]; 181 NormImageSubtracted = NormImage NormImageDark; NormImageSubtracted/=Max[NormImageSubtracted];NormImageSubtracted/=Max[NormImageSubtracted]NormImageSubtracted/=Max[NormI IonImage = NormImageSubtracted/.x /;x < ToExpression[$CommandLine[[8]]]->0; Here we nd the vertical Region for ion1, VertStart and VertEnd that de ne the vertical edges. We nd VertStart by starting at 0, and then scan the horizontal pixels, looking for brightness > threshold. If no non-zero pixels are found, VertStart is shifted by 1 and we scan again ....VertEnd starts from the height, and we scan along the horizontals, if no bright pixel is found we shift to height - 1 and scan again and so on ... VertStart = 0; VertEnd = height; For[n = 1; n < height + 1; n++; For[i = 1; i < width + 1; i++; If[IonImage[[n; i]] > 0;VertStart = n; i = width + 1;n = height + 1]]] For[n = height; n > VertStart; n{; For[i = 1; i < width + 1; i++; If[IonImage[[n; i]] > 0;VertEnd = n; i = width + 1;n = 1]]] If[VertEnd > height;VertEnd = height];t] VertStart = If[VertStart > 0;VertStart; 1]; Export[\uncuto image " <> DateString[f\Year";\Month";\Day"g] <> \.jpg"; Image[IonImage[[All; All]]]] HEdge = Join[ff1; 0gg;Table[f1; 1g; fi; 1; Nionsg]] center = VertStart + Floor[(VertEnd VertStart)=2];ertStarertStar 182 For[i = 1; i<=Nions; i++; For[j = HEdge[[i; 2]] + 1; j < width; j++; If[Count[Table[N[IonImage[[v; j]]]; fv;VertStart;VertEnd; 1g]; 0:] < (3(VertEnd VertStart))=4; HEdge[[i+ 1; 1]] = j; For[k = j; k < width; k++; If[Count[Table[N[IonImage[[v; k]]]; fv;VertStart;VertEnd; 1g]; 0:] > (3(VertEnd VertStart))=4; HEdge[[i+ 1; 2]] = k 1; k = width; j = width] ]]; ]] HEdge = Delete[HEdge; 1] dataFrames = Table[dataT[[VertStart;;VertEnd; widthx+ 1;;(x+ 1)width]];ertEnrtEn fx; 0; Nimages 1; 1g]; darkFrames = Table[DarkDataT[[VertStart;;VertEnd; widthx+ 1;;(x+ 1)width]];ertEn fx; 0; Nimages 1; 1g]; HistogramDataBright = Table[Total[Flatten[dataFrames[[j; All; HEdge[[i; 1]];;otal[Flatten[dataotal[Flatten[data HEdge[[i; 2]]]]]]; fi; Nionsg; fj; 1; Nimagesg]; HistogramDataDark = Table[Total[Flatten[darkFrames[[j; All; HEdge[[i; 1]];;able[ HEdge[[i; 2]]]]]]; fi; Nionsg; fj; 1; Nimagesg]; Export[\Histograms " <> DateString[f\Year";\Month";\Day"g] <> \.pdf"; Table[Histogram[fHistogramDataDark[[i]]; HistogramDataBright[[i]]g;t[ f500g; ImageSize->300]; fi; Nionsg]//TableForm] Export[\Histogram tables dark" <> DateString[f\Year";\Month";\Day"g] <> \.xls"; Table[Table[f#[[1; i]]; #[[2; i]]g; fi; Length[#[[2]]]g]& [HistogramList[HistogramDataDark[[j]]]]; fj; Nionsg]] Export[\Histogram tables bright" <> DateString[f\Year";\Month";\Day"g] <> \.xls";Table[Table[f#[[1; i]]; 183 #[[2; i]]g; fi; Length[#[[2]]]g]&[HistogramList[HistogramDataBright[[j]]]]; fj; Nionsg]] ErrorThreshold = Table[f0; 0g; fi; Nionsg]; Steps = 50; For[i = 1; i < Nions + 1; i++; DL = HistogramList[HistogramDataDark[[i]]; fStepsg];HistogramList[HistogramDataDark[HistogramList[HistogramDataDark[ BL = HistogramList[HistogramDataBright[[i]]; fStepsg];HistogramList[HistogramDataBrigHistogramList[HistogramDataBrig We match the histogram dimensions Di L = Min[BL[[1]]] Min[DL[[1]]]; Di U = Max[BL[[1]]] Max[DL[[1]]];Max[BL[[1]] If[Di L>=0; fBLNew1 = Join[Table[Min[BL[[1]]] Abs[Di L] + (n 1) Steps;able[Min[BL[[1]] fn; 1; Floor[Abs[Di L]=Steps]g]; BL[[1]]]; BLNew2 = Join[Table[0; fn; 1; Floor[Di L=Steps]g]; BL[[2]]]; BLNewInt = Join[fBLNew1g; fBLNew2g]; DLNewInt = DL; g; fDLNew1 = Join[Table[Min[DL[[1]]] Abs[Di L] + (n 1) Steps;able[Min[DL[[1]] fn; 1; Floor[Abs[Di L]=Steps]g]; DL[[1]]]; DLNew2 = Join[Table[0; fn; 1; Floor[Abs[Di L]=Steps]g]; DL[[2]]]; DLNewInt = Join[fDLNew1g; fDLNew2g]; BLNewInt = BL; g]; If[Di U>=0; fDLNew3 = Join[DLNewInt[[1]];Table[Max[DLNewInt[[1]]] + n Steps; fn; 1; Floor[Di U=Steps]g]]; DLNew4 = Join[DLNewInt[[2]];Table[0; fn; 1; Floor[Di U=Steps]g]]; DLNew = Join[fDLNew3g; fDLNew4g]; BLNew = BLNewInt; g;BLNewI 184 fBLNew3 = Join[BLNewInt[[1]];Table[Max[BLNewInt[[1]]] + n Steps; fn; 1; Floor[Abs[Di U]=Steps]g]]; BLNew4 = Join[BLNewInt[[2]];Table[0; fn; 1; Floor[Abs[Di U]=Steps]g]]; BLNew = Join[fBLNew3g; fBLNew4g]; DLNew = DLNewIntg; ]; TDL = Table[Total[DLNew[[2; 1;;j]]]; fj; 1; Dimensions[DLNew[[2]]][[1]]g]; TBL = Table[Nimages Total[BLNew[[2; 1;;j]]]; fj; 1; Dimensions[BLNew[[2]]][[1]]g]; Di = TDL TBL; Now we nd the unbiased thresholds and export our ndings for the control program to use x = Position[Abs[Di ]; Min[Abs[Di ]]][[1; 1]];osition[Abs[Di osition[Abs[Di Error = N[(Sum[BLNew[[2; n]]; fn; 1; xg]=Nimages) 100]; ErrorThreshold[[i]] = fVertStart 1;VertEnd; HEdge[[i; 1]] 1; HEdge[[i; 2]]; Error; BLNew[[1; x]]g; ] Export[\ROI Error and Thresholds.csv"; ErrorThreshold]an ErrorThreshold = Prepend[ErrorThreshold; f\Lower Vertical Edge";\Upper Vertical Edge";EdgeEdg \Left Edge"; \Right Edge"; \State detection error"; \Threshold"g] Export[\ROI Error and Thresholds " <> DateString[f\Year";\Month";\Day"g]an <> \.xls"; ErrorThreshold] 185 Bibliography [1] S. Olmschenk, K. C. Younge, D. L. Moehring, D. N. Matsukevich, P. Maunz, and C. Monroe. Manipulation and detection of a trapped yb[sup +] hyper ne qubit. Phys. Rev. A, 76(5):052314, 2007. [2] M. J. Madsen, W.K. Hensinger, D. Stick, J.A. Rabchuk, and C. Monroe. Planar ion trap geometry for microfabrication. Appl. Phys. B, 78:639, 2004. [3] A. S rensen and K. M lmer. Quantum computation with ions in thermal motion. Phys. Rev. Lett., 82(9):1971{1974, 1999. [4] Richard Feynman. Simulating physics with computers. Intl J Theo Phys, 21(6/7):467, 1982. [5] S. Lloyd. Universal quantum simulators. Science, 273(5278):1073{1078, 1996. [6] S. Sachdev. Quantum Phase Transitions. Cambridge University Press, 2011. [7] K. Binder and A. P. Young. Spin glasses: Experimental facts, theoretical concepts, and open questions. Rev. Mod. Phys., 58:801{976, 1986. [8] L. Balents. Spin liquids in frustrated magnets. Nature, 464:199{208, 2010. [9] B A Cipra. The ising model is np-complete. SIAM News, 33(6):1{3, 2000. [10] Satish Damaraju, Varghese George, Sanjeev Jahagirdar, Tanveer Khondker, Robert Milstrey, Sanjib Sarkar, Scott Siers, Israel Stolero, and Arun Subbiah. A 22nm ia multi-cpu and gpu system-on-chip. In Solid-State Circuits Con- ference Digest of Technical Papers (ISSCC), 2012 IEEE International, pages 56{57. IEEE, 2012. [11] M. A. Nielsen and I. L. Chuang. Quantum Computation and Quantum Infor- mation. Cambridge University Press, Cambridge, UK, 2000. [12] R. Moessner and A. P. Ramirez. Geometrical frustration. Phys. Today, 59:24, Feb 2006. [13] Leon Balents. Spin liquids in frustrated magnets. Nature, 464:199, 2010. [14] Tru Moriya and Kazuo Ueda. Antiferromagnetic spin uctuation and super- conductivity. Reports on Progress in Physics, 66(8):1299, 2003. [15] S. Sachdev. Order and quantum phase transitions in the cuprate supercon- ductors. Rev. Mod. Phys., 75:913{932, 2003. [16] Jacob D. Biamonte and Peter J. Love. Realizable hamiltonians for universal adiabatic quantum computers. Phys. Rev. A, 78(1):012352, 2008. 186 [17] D. Aharonov, W. van Dam, J. Kempe, Z. Landau, S. Lloyd, and O. Regev. Adiabatic quantum computation is equivalent to standard quantum compu- tation. SIAM Review, 50(4):755{787, 2008. [18] John M Martinis, M Ansmann, and J Aumentado. Energy decay in super- conducting josephson-junction qubits from nonequilibrium quasiparticle exci- tations. Physical review letters, 103(9):097002, 2009. [19] J Majer, JM Chow, JM Gambetta, Jens Koch, BR Johnson, JA Schreier, L Frunzio, DI Schuster, AA Houck, A Wallra , et al. Coupling superconduct- ing qubits via a cavity bus. Nature, 449(7161):443{447, 2007. [20] Edward Farhi, Je rey Goldstone, Sam Gutmann, and Michael Sipser. Quan- tum computation by adiabatic evolution. arXiv preprint quant-ph/0001106, 2000. [21] Anders W. Sandvik. Ground states of a frustrated quantum spin chain with long-range interactions. Phys. Rev. Lett., 104:137204, Mar 2010. [22] E. Farhi and S. Gutmann. Analog analogue of a digital quantum computation. Physical Review A, 57(4):2403{2406, 1998. [23] DM Tong, K Singh, LC Kwek, and CH Oh. Su ciency criterion for the validity of the adiabatic approximation. Physical review letters, 98(15):150402, 2007. [24] D F. James and J Jerke. E ective hamiltonian theory and its applications in quantum information. Canadian Journal of Physics, 85(6):625{632, 2007. [25] WC Martin and WL Wiese. Atomic, molecular and optical physics handbook. American Institute of Physics: Woodbury, New York, 1996. [26] S A King, R M Godun, S A Webster, H S Margolis, L A M Johnson, K Szy- maniec, P E G Baird, and P Gill. Absolute frequency measurement of the 2 s 1/2 - 2 f 7/2 electric octupole transition in a single ion of 171 yb + with 10 -15 fractional uncertainty. New Journal of Physics, 14(1):013045, 2012. [27] Kazuhiko Sugiyama and Jun Yoda. Production of ybh+ by chemical reaction of yb+ in excited states with h2 gas. Phys. Rev. A, 55:R10{R13, Jan 1997. [28] D. J. Wineland, R. E. Drullinger, and F. L. Walls. Radiation-pressure cooling of bound resonant absorbers. Phys. Rev. Lett., 40:1639{1642, Jun 1978. [29] Q. A. Turchette, Kielpinski, B. E. King, D. Leibfried, D. M. Meekhof, C. J. Myatt, M. A. Rowe, C. A. Sackett, C. S. Wood, W. M. Itano, C. Monroe, and D. J. Wineland. Heating of trapped ions from the quantum ground state. Phys. Rev. A, 61:063418, May 2000. [30] D. J. Berkeland and M. G. Boshier. Destabilization of dark states and optical spectroscopy in zeeman-degenerate atomic systems. Phys. Rev. A, 65:033413, Feb 2002. 187 [31] M Acton, K.-A. Brickman, P. C. Haljan, P. J. Lee, L. Deslauriers, and C. Mon- roe. Near-perfect simultaneous measurement of a qubit register. Quantum Info. Comput., 6(6):465{482, September 2006. [32] P J Lee, K-A Brickman, L Deslauriers, P C Haljan, L-M Duan, and C Monroe. Phase control of trapped ion quantum gates. Journal of Optics B: Quantum and Semiclassical Optics, 7(10):S371, 2005. [33] A. S rensen and K. M lmer. Entanglement and quantum computation with ions in thermal motion. Phys. Rev. A, 62:022311, 2000. [34] J. I. Cirac and P. Zoller. Quantum computation with cold trapped ions. Phys. Rev. Lett., 74:4091{4094, 1995. [35] J. J. Garc a-Ripoll, P. Zoller, and J. I. Cirac. Speed optimized two-qubit gates with laser coherent control techniques for ion trap quantum computing. Phys. Rev. Lett., 91:157901, 2003. [36] D. F. V. James. Quantum dynamics of cold trapped ions, with application to quantum computation. Appl. Phys. B, 66:181{190, 1998. [37] C.-C. Joseph Wang and J. K. Freericks. Intrinsic phonon e ects on analog quantum simulators with ultracold trapped ions. Phys. Rev. A, 86:032329, Sep 2012. [38] D. L. Moehring, P. Maunz, S. Olmschenk, K. C. Younge, D. N. Matsukevich, L.-M. Duan, and C. Monroe. Entanglement of single-atom quantum bits at a distance. Nature, 449:68, 2007. [39] P. Maunz, D. L. Moehring, S. Olmschenk, K. C. Younge, D. N. Matsuke- vich, and C. Monroe. Quantum interference of photon pairs from two remote trapped atomic ions. Nature Physics, 3(8):538{541, June 2007. [40] Wolfgang Paul. Electromagnetic traps for charged and neutral particles. Rev. Mod. Phys., 62:531{540, Jul 1990. [41] S. Earnshaw. On the nature of the molecular forces which regulate the con- stitution of the luminiferous ether. Trans. Camb. Phil. Soc, 7:97{112, 1842. [42] D. J. Berkeland, J. D. Miller, J. C. Bergquist, W. M. Itano, and D. J. Wineland. Minimization of ion micromotion in a paul trap. Journal of Applied Physics, 83(10):5025{5033, 1998. [43] HG Dehmelt. Radiofrequency spectroscopy of stored ions i: Storage. Advances in Atomic and Molecular Physics, 3:53{72, 1967. [44] DJ Wineland, C. Monroe, WM Itano, D. Leibfried, BE King, and DM Meekhof. Experimental issues in coherent quantum-state manipulation of trapped atomic ions. Journal of Research of the National Institute of Stan- dards and Technology, 103(3), 1998. 188 [45] R.R.A. Syms, T.J. Tate, M.M. Ahmad, and S. Taylor. Design of a microengi- neered electrostatic quadrupole lens. Electron Devices, IEEE Transactions on, 45(11):2304{2311, 1998. [46] A. Zverev and H. Blinchiko . Realization of a lter with helical components. Component Parts, IRE Transactions on, 8(3):99{110, 1961. [47] Andrew Chew. Doppler-Free Spectroscopy of Iodine at 739nm. Undergraduate honors thesis, The University of Michigan, 2008. [48] Pierre Dub e and Michael Trinczek. Hyper ne-structure splittings and absorp- tion strengths of molecular-iodine transitions near the trapping frequencies of francium. JOSA B, 21(6):1113{1126, 2004. [49] Michael Mond, Holger Schoene, Andreas Diening, Guenter Hollemann, and Wolf Seelert. Highly stable uv-mode-locked lasers with an output power of 35 w at 355 nm. In Proc. SPIE, volume 7193, page 71930W, 2009. [50] Mingwei Li D. James Kafka. Method and apparatus for laser micro-machining of polymeric articles using a mode-locked laser. Patent, (EP 1490195 A1), 12 2004. [51] D. Hayes, D. N. Matsukevich, P. Maunz, D. Hucul, Q. Quraishi, S. Olmschenk, W. Campbell, J. Mizrahi, C. Senko, and C. Monroe. Entanglement of atomic qubits using an optical frequency comb. Phys. Rev. Lett., 104(14):140501, 2010. [52] Louis Deslauriers. Cooling and heating of the quantum motion of trapped Cd+ ions. PhD thesis, University of Michigan., 2006. [53] Aaron Christopher Lee. Ytterbium Ion Qubit State Detection On An ICCD Camera. Undergraduate honors thesis, University of Maryland, 2012. [54] C Shen and L-M Duan. Correcting detection errors in quantum state engi- neering through data processing. New Journal of Physics, 14(5):053053, 2012. [55] K Kim, S Korenblit, R Islam, EE Edwards, MS Chang, C Noh, H Carmichael, GD Lin, LM Duan, CC Joseph Wang, et al. Quantum simulation of the trans- verse ising model with trapped ions. New Journal of Physics, 13(10):105003, 2011. [56] P. J. Lee, B. B. Blinov, K.-A. Brickman, L. Deslauriers, M. Madsen, R. Miller, D. L. Moehring, D. Stick, and C. Monroe. Atomic qubit operations with an electro-optic modulator. Opt. Lett., 28:1582, 2003. [57] K. Kim, M.-S. Chang, R. Islam, S. Korenblit, L.-M. Duan, and C. Monroe. Entanglement and tunable spin-spin couplings between trapped ions using multiple transverse modes. Phys. Rev. Lett., 103(12):120502, 2009. 189 [58] H T Quan and W H Zurek. Testing quantum adiabaticity with quench echo. New Journal of Physics, 12:093025, 2010. [59] K. Kim, M.-S. Chang, S. Korenblit, R. Islam, E. E. Edwards, J.K. Freericks, G.D. Lin, L.-M. Duan, and C. Monroe. Quantum simulation of frustrated ising spins with trapped ions. Nature, 465:590{593, 2010. [60] G-D Lin, C Monroe, and L-M Duan. Sharp phase transitions in a small frus- trated network of trapped ion spins. Physical Review Letters, 106(23):230402, 2011. [61] P Lecheminant and HP Diep. Frustrated spin systems, 2005. [62] GH Wannier. Antiferromagnetism. the triangular ising net. Physical Review, 79(2):357, 1950. [63] MJ Harris, ST Bramwell, DF McMorrow, Th Zeiske, and KW Godfrey. Ge- ometrical frustration in the ferromagnetic pyrochlore ho f2g ti f2g o f7g. Physical Review Letters, 79(13):2554{2557, 1997. [64] Steven T Bramwell and Michel JP Gingras. Spin ice state in frustrated mag- netic pyrochlore materials. Science, 294(5546):1495{1501, 2001. [65] Christopher M Dawson, Henry L Haselgrove, and Michael A Nielsen. Noise thresholds for optical cluster-state quantum computation. Physical Review A, 73(5):052306, 2006. [66] Bruce Normand and Andrzej M Ole s. Frustration and entanglement in the t f2gg spin-orbital model on a triangular lattice: Valence-bond and generalized liquid states. Physical Review B, 78(9):094427, 2008. [67] Otfried Ghne and Gza Tth. Entanglement detection. Physics Reports, 474(16):1 { 75, 2009. [68] 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, 404:256, 2000. [69] H Schmitz, A Friedenauer, Ch Schneider, R Matjeschk, M Enderlein, T Huber, J Glueckert, D Porras, and T Schaetz. The arch of simulating quantum spin systems with trapped ions. Applied Physics B, 95(2):195{203, 2009. [70] E. E. Edwards, S. Korenblit, K. Kim, R. Islam, M.-S. Chang, J. K. Freericks, G.-D. Lin, L.-M. Duan, and C. Monroe. Quantum simulation and phase diagram of the transverse- eld ising model with three atomic spins. Phys. Rev. B, 82:060412, 2010. [71] Tommaso Caneva, Rosario Fazio, and Giuseppe E Santoro. Adiabatic quantum dynamics of the lipkin-meshkov-glick model. Physical Review B, 78(10):104426, 2008. 190 [72] X. L. Deng, D. Porras, and J. I. Cirac. E ective spin quantum phases in systems of trapped ions. Physical Review A, 72(6):063407, 2005. [73] Dietrich Leibfried, Emanuel Knill, Signe Seidelin, Joe Britton, R Brad Blakestad, John Chiaverini, David B Hume, Wayne M Itano, John D Jost, Christopher Langer, et al. Creation of a six-atom schr odinger catstate. Na- ture, 438(7068):639{642, 2005. [74] R Ozeri, WM Itano, RB Blakestad, J Britton, J Chiaverini, JD Jost, C Langer, D Leibfried, R Reichle, S Seidelin, et al. Errors in trapped-ion quantum gates due to spontaneous photon scattering. Physical Review A, 75(4):042329, 2007. [75] W. C. Campbell, J. Mizrahi, Q. Quraishi, C. Senko, D. Hayes, D. Hucul, D. N. Matsukevich, P. Maunz, and C. Monroe. Ultrafast gates for single atomic qubits. Phys. Rev. Lett., 105:090502, 2010. [76] Craig R Clark, Tzvetan S Metodi, Samuel D Gasster, and Kenneth R Brown. Resource requirements for fault-tolerant quantum simulation: The ground state of the transverse ising model. Physical Review A, 79(6):062314, 2009. [77] R. Islam, E.E. Edwards, K. Kim, S. Korenblit, C. Noh, H. Carmichael, G.-D. Lin, L.-M. Duan, C.-C. Joseph Wang, J.K. Freericks, and C. Monroe. Onset of a quantum phase transition with a trapped ion quantum simulator. Nat Commun, 2(377), 2011. [78] Lin, G.-D., Zhu, S.-L., Islam, R., Kim, K., Chang, M.-S., Korenblit, S., Mon- roe, C., and Duan, L.-M. Large-scale quantum computation in an anharmonic linear ion trap. Euro. Phys. Lett., 86:60004, 2009. [79] Roman Schmied, Janus H Wesenberg, and Dietrich Leibfried. Optimal surface- electrode trap lattices for quantum simulation with trapped ions. Physical review letters, 102(23):233002, 2009. [80] Robert J Clark, Tongyan Lin, Kenneth R Brown, and Isaac L Chuang. A two-dimensional lattice ion trap for quantum simulation. Journal of Applied physics, 105(1):013114{013114, 2009. [81] D. Porras and J. I. Cirac. E ective quantum spin systems with trapped ions. Phys. Rev. Lett., 92(20):207901, 2004. [82] Y.-J. Han, R. Raussendorf, and L.-M. Duan. Scheme for demonstration of fractional statistics of anyons in an exactly solvable model. Phys. Rev. Lett., 98(15):150404, 2007. [83] A. Kitaev. Fault-tolerant quantum computation by anyons. Ann. Phys., 303:2{ 30, 2003. 191 [84] Chetan Nayak, Steven H Simon, Ady Stern, Michael Freedman, and Sankar Das Sarma. Non-abelian anyons and topological quantum compu- tation. Reviews of Modern Physics, 80(3):1083, 2008. [85] Jonathan Simon, Waseem S. Bakr, Ruichao Ma, M. Eric Tai, Philipp M. Preiss, and Markus Greiner. Quantum simulation of antiferromagnetic spin chains in an optical lattice. Nature, 472:307{312, 2011. [86] Per Bak. The devil?s staircase. Physics Today, 39:38{45, 1986. [87] Philipp Hauke, Fernando M Cucchietti, Alexander M uller-Hermes, Mari- Carmen Ba~nuls, J Ignacio Cirac, and Maciej Lewenstein. Complete devil?s staircase and crystal{super uid transitions in a dipolar xxz spin chain: a trapped ion quantum simulation. New Journal of Physics, 12(11):113037, 2010. [88] Per Bak and R Bruinsma. One-dimensional ising model and the complete devil?s staircase. Physical Review Letters, 49(4):249{251, 1982. [89] Zhaohui Wei and Mingsheng Ying. Quantum adiabatic computation and adi- abatic conditions. Phys. Rev. A, 76:024304, Aug 2007. [90] Marcos Rigol, Vanja Dunjko, and Maxim Olshanii. Thermalization and its mechanism for generic isolated quantum systems. Nature, 452(7189):854{858, 2008. [91] JM Deutsch. Quantum statistical mechanics in a closed system. Physical Review A, 43(4):2046, 1991. [92] T. Monz, P. Schindler, J. T. Barreiro, M. Chwalla, D. Nigg, W. A. Coish, M. Harlander, W. H ansel, M. Hennrich, and R. Blatt. 14-Qubit Entanglement: Creation and Coherence. Physical Review Letters, 106(13):130506, April 2011. [93] Alejandro Perdomo-Ortiz, Neil Dickson, Marshall Drew-Brook, Geordie Rose, and Al an Aspuru-Guzik. Finding low-energy conformations of lattice protein models by quantum annealing. Scienti c Reports, 2, 2012. [94] 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. [95] D. J. Wineland, C. Monroe, W. M. Itano, D. Leibfried, B. E. King, and D. M. Meekhof. Experimental issues in coherent quantum-state manipulation of trapped atomic ions. J. Res. Nat. Inst. Stand. Tech., 103:259{328, 1998. [96] D. Wineland and R. Blatt. Entangled states of trapped atomic ions. Nature, 453:1008{1014, 2008. 192 [97] J. M. Taylor and T. Calarco. Wigner crystals of ions as quantum hard drives. Phys. Rev. A, 78:062331, 2008. [98] R. J. Clark, T. Linb, K. R. Brown, and I. L. Chuang. A two-dimensional lattice ion trap for quantum simulation. J. Appl. Phys., 105:013114, 2009. [99] Roman Schmied, Janus H. Wesenberg, and Dietrich Leibfried. Optimal surface-electrode trap lattices for quantum simulation with trapped ions. Phys. Rev. Lett., 102:233002, 2009. [100] P. C. Haljan, K.-A. Brickman, L. Deslauriers, P. J. Lee, and C. Monroe. Spin-dependent forces on trapped ions for phase-stable quantum gates and entangled states of spin and motion. Phys. Rev. Lett., 94:153602, 2005. [101] Shi-Liang Zhu, C. Monroe, and L.-M. Duan. Trapped ion quantum computa- tion with transverse phonon modes. Phys. Rev. Lett., 97, 2006. [102] D. Leibfried, R. Blatt, C. Monroe, and D. Wineland. Quantum dynamics of single trapped ions. Rev. Mod. Phys., 75:281{324, 2003. [103] J. J. Garca-Ripoll, P. Zoller, and J. I. Cirac. Coherent control of trapped ions using o -resonant lasers. Phys. Rev. A, 71:062309, Jun 2005. [104] R. Moessner and A. P. Ramirez. Geometrical frustration. Phys. Today, 59:24, Feb 2006. [105] C. Knoernschild, X. L. Zhang, L. Isenhower, A. T. Gill, F. P. Lu, M. Sa man, and J. Kim. Independent individual addressing of multiple neutral atom qubits with a micromirror-based beam steering system. Appl. Phys. Lett., 97(13):134101, 2010. [106] C. Knoernschild, T. Kim, P. Maunz, S. Crain, and J. Kim. Stable optical phase modulation with micromirrors. arXiv:1108.5782, 2011. [107] F. Shaikh, A. Ozakin, J. M. Amini, H. Hayden, C.-S. Pai, C. Volin, D. R. Deni- son, D. Faircloth, A. W. Harter, and R. E. Slusher. Monolithic microfabricated symmetric ion trap for quantum information processing. arXiv:1105.4909, 2011. [108] A. Kitaev. Anyons in an exactly solved model and beyond. Annals of Physics, 321(1):2{111, 2006. [109] M. M. Rams, M. Zwolak, and B. Damski. A quantum phase transition in a quantum external eld: The formation of a Schroedinger magnet. arXiv:1201.1932, January 2012. [110] B. P. Lanyon, C. Hempel, D. Nigg, M. M uller, R. Gerritsma, F. Zhringer, P. Schindler, J. T. Barreiro, M. Rambach, G. Kirchmair, M. Hennrich, P. Zoller, R. Blatt, and C. F. Roos. Universal digital quantum simulation with trapped ions. Science, 2011. 193 [111] Michael J. Biercuk, Hermann Uys, Aaron P. Vandevender, Nobuyasu Shiga, Wayne M. Itano, and John J. Bollinger. High- delity quantum control us- ing ion crystals in a penning trap. Quantum Information & Computation, 9(11):920{949, 2011. 194