POSITION AND TEMPERATURE MEASUREMENTS OF A SINGLE ATOM VIA RESONANT FLUORESCENCE by RICHARD WAGNER JR. A DISSERTATION Presented to the Department of Physics and the Graduate School of the University of Oregon in partial fulfillment of the requirements for the degree of Doctor of Philosophy June 2019 DISSERTATION APPROVAL PAGE Student: Richard Wagner Jr. Title: Position and Temperature Measurements of a Single Atom via Resonant Fluorescence This dissertation has been accepted and approved in partial fulfillment of the requirements for the Doctor of Philosophy degree in the Department of Physics by: Michael Raymer Chair Daniel A Steck Advisor Benjamı́n J. Alemán George Nazin Institutional Representative and Janet Woodruff-Borden Dean of the Graduate School Original approval signatures are on file with the University of Oregon Graduate School. Degree awarded June 2019 ii ©c 2019 Richard Wagner Jr. This work is licensed under a Creative Commons Attribution-NonCommercial-NoDerivs (United States) License. This work may be shared freely with attribution to the author and institution. It may not be edited or used for any commercial purpose. iii DISSERTATION ABSTRACT Richard Wagner Jr. Doctor of Philosophy Department of Physics June 2019 Title: Position and Temperature Measurements of a Single Atom via Resonant Fluorescence The magneto-optical trap (MOT) has been an important tool in quantum optics research for three decades. MOTs allow for hundreds of thousands to millions of atoms to be cooled to micro-Kelvin temperatures for use in a wide variety of experiments. For nearly as long, MOTs with just a single atom have been of some interest to the research community. We have developed an algorithm, based on Bayesian statistics, to carefully measure small numbers of atoms in a MOT. Many techniques have been developed to measure the temperature of atoms in a MOT, including some that can translate to single atoms. We propose a new technique to measure the temperature of a single atom without releasing the atom from the MOT. Temporal modulations in a spatially dependent magnetic field encode information about the position of an atom through associated variation in its fluorescence rate. Measuring this variation reveals the atom’s position distribution and therefore its temperature. The technique is examined for a variety of MOT parameters. Measurements with the technique are an order of magnitude larger than predicted by theory and potential routes for future study are offered. iv CURRICULUM VITAE NAME OF AUTHOR: Richard Wagner Jr. GRADUATE AND UNDERGRADUATE SCHOOLS ATTENDED: University of Oregon, Eugene, OR Miami University, Oxford, OH DEGREES AWARDED: Doctor of Philosophy in Physics, 2019, University of Oregon Master of Science in Physics, 2016, University of Oregon Bachelor of Science in Physics, 2009, Miami University PROFESSIONAL EXPERIENCE: Graduate Research Fellow, Department of Physics, University of Oregon, Eugene OR, Fall 2012, AY 2012-2013, Winter & Spring 2014, Fall 2015, Fall 2016, Spring 2016, Fall 2017, Spring 2016, AY 2017-2018 Science Literacy Program Fellow, University of Oregon, Eugene OR, Winter & Spring 2012, Fall 2014 Graduate Teaching Fellow, Department of Physics, University of Oregon, Eugene OR, AY 2009-2010, AY 2019-2011, Winter & Spring 2015, Winter 2016, Winter 2017 Adjunct Instructor, Lane Community College, Eugene OR, AY 2016-2017, AY 2017-2018 GRANTS, AWARDS AND HONORS: Science Literacy Program Fellow, University of Oregon, Winter & Spring 2012, Fall 2014 Astronaut Scholar, Astronaut Scholarship Foundation, 9/2008 Provost’s Academic Achievement Award, Miami University Office of the Provost, 9/2008 v PUBLICATIONS: Richard Wagner and James P. Clemens. Fidelity of quantum teleportation based on spatially and temporally resolved spontaneous emission. Journal of the Optical Society of America B, 27(6):A73-A80, 2010. N. Souther, R, Wagner, P. Harnish, M. Briel, and S, Bali. Measurements of light shifts in cold atoms using Raman pump-probe spectroscopy. Laser Physics Letters, 7(4):321-327, 2010. Richard Wagner and James P. Clemens. Performance of a quantum teleportation protocol based on collective spontaneous emission. Journal of the Optical Society of America B, 26(3):541-548, 2009. Richard Wagner Jr. and James P. Clemens. Performance of a quantum teleportation protocol based on temporally resolve photodetection of collective spontaneous emission. Physical Review A, 79:042322, 2009. vi For Mom and Dad. Thank you. vii TABLE OF CONTENTS Chapter Page I. INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 1.1. Single Atom MOTs . . . . . . . . . . . . . . . . . . . . . . . 2 1.2. MOT Temperature Measurements . . . . . . . . . . . . . . . 4 1.3. Optical Trap Oscillations . . . . . . . . . . . . . . . . . . . . 7 1.4. Dissertation Outline . . . . . . . . . . . . . . . . . . . . . . 8 II. ATOM OPTICS AND MAGNETO-OPTICAL TRAPS . . . . . . . . . 9 2.1. A Single Atom . . . . . . . . . . . . . . . . . . . . . . . . . 9 2.2. Atoms and Light . . . . . . . . . . . . . . . . . . . . . . . . 11 2.3. Atoms and Magnetic Fields . . . . . . . . . . . . . . . . . . 21 2.4. The Fg = 0→ Fe = 1 Atom . . . . . . . . . . . . . . . . . . 25 2.5. Magneto-Optical Traps . . . . . . . . . . . . . . . . . . . . . 35 III. EXPERIMENTAL SETUP . . . . . . . . . . . . . . . . . . . . . . . . . 44 3.1. Vacuum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44 3.2. Lasers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53 3.3. Magnetic Fields . . . . . . . . . . . . . . . . . . . . . . . . . 66 3.4. Photon Collection . . . . . . . . . . . . . . . . . . . . . . . . 80 IV. OUR SINGLE ATOM MOT . . . . . . . . . . . . . . . . . . . . . . . . 90 4.1. Detecting A Single Atom MOT . . . . . . . . . . . . . . . . 91 4.2. Bayesian Algorithm . . . . . . . . . . . . . . . . . . . . . . . 100 V. ATOMIC FORCES IN A MOT . . . . . . . . . . . . . . . . . . . . . . 114 5.1. 3D and 87Rb Hamiltonians . . . . . . . . . . . . . . . . . . . 114 5.2. Matching Simulation to Experiments . . . . . . . . . . . . . 123 viii Chapter Page 5.3. 3D and 87Rb Calculations . . . . . . . . . . . . . . . . . . . 128 5.4. Escape Channels . . . . . . . . . . . . . . . . . . . . . . . . 134 5.5. Recovering Potential . . . . . . . . . . . . . . . . . . . . . . 143 VI. POSITION AND TEMPERATURE MEASUREMENTS . . . . . . . . 146 6.1. Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146 6.2. Analysis of Photon Arrivals . . . . . . . . . . . . . . . . . . 156 6.3. Measurements . . . . . . . . . . . . . . . . . . . . . . . . . . 165 6.4. Multiple Atom Fluorescence Amplitudes . . . . . . . . . . . 184 6.5. Parametric Resonances . . . . . . . . . . . . . . . . . . . . . 185 6.6. Non-Sinusoidal Waveforms . . . . . . . . . . . . . . . . . . . 198 VII. CONCLUSIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 203 APPENDICES A. HOW A FG = 0→ FE = 1 ATOM BECOMES A V-ATOM . . . . . 208 B. DUAL POLARIZER NOISE REDUCTION . . . . . . . . . . . . . . 212 C. MOT COIL WATER COOLING RATE . . . . . . . . . . . . . . . . 215 D. BAYESIAN EVOLUTION DERIVATION . . . . . . . . . . . . . . . 222 E. GAUSSIAN SAMPLED OSCILLATION AMPLITUDE . . . . . . . 225 F. PARAMETRIC RESONANCE DERIVATION . . . . . . . . . . . . 229 REFERENCES CITED . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233 ix LIST OF FIGURES Figure Page 1.1. The magneto-optical trap . . . . . . . . . . . . . . . . . . . . . . . . . . 2 2.1. 87Rb D2 Transition Level Diagram . . . . . . . . . . . . . . . . . . . . . 22 2.2. Fg = 0→ Fe = 1 Atom Level Diagrams . . . . . . . . . . . . . . . . . . 26 2.3. Optical molasses beam arrangement . . . . . . . . . . . . . . . . . . . . 32 2.4. V-atom forces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32 2.5. Magneto-Optical Trap Level Diagram . . . . . . . . . . . . . . . . . . . 36 2.6. Atom with Mutlple Ground States . . . . . . . . . . . . . . . . . . . . . 41 2.7. Sisyphus Cooling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42 3.1. Experimental Diagram of Vacuum Chamber . . . . . . . . . . . . . . . . 46 3.2. Installed LIAD system . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49 3.3. LED circuit for LIAD setup . . . . . . . . . . . . . . . . . . . . . . . . . 49 3.4. (Re)baking the vacuum chamber . . . . . . . . . . . . . . . . . . . . . . 52 3.5. Beam Path of MOT Trapping Lasers . . . . . . . . . . . . . . . . . . . . 54 3.6. Beam Path of MOT Repumping Lasers . . . . . . . . . . . . . . . . . . 54 3.7. Pyramid MOT Mirrors . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56 3.8. MOT Frequency Controlling Beam Path . . . . . . . . . . . . . . . . . . 61 3.9. 87Rb D-2 line spectrum . . . . . . . . . . . . . . . . . . . . . . . . . . . 62 3.10. Fiber Polarization Mount . . . . . . . . . . . . . . . . . . . . . . . . . . 65 3.11. Permanent Magnet Magnetic Field Calculation Methods . . . . . . . . . 68 3.12. Permanent Magnet Quadrupole Field . . . . . . . . . . . . . . . . . . . . 69 3.13. Single Atom MOT Magnetic Field Designs . . . . . . . . . . . . . . . . . 70 x Figure Page 3.14. Plumbing Diagram for water cooling the MOT coils . . . . . . . . . . . . 74 3.15. Water Cooled Coils Protection Circuit . . . . . . . . . . . . . . . . . . . 75 3.16. MOT Coils Current Monitoring Circuit . . . . . . . . . . . . . . . . . . 77 3.17. Helmholtz Coil Layout . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78 3.18. Single Photon Collect Lens System . . . . . . . . . . . . . . . . . . . . . 81 3.19. BNC-Header Adaptor for DE2 FPGA . . . . . . . . . . . . . . . . . . . 84 3.20. FPGA implementation for counting photons . . . . . . . . . . . . . . . . 85 3.21. FPGA implementation for timing photons . . . . . . . . . . . . . . . . . 86 3.22. APD Protection Mechanisms . . . . . . . . . . . . . . . . . . . . . . . . 88 4.1. Sample Photon Collection Data . . . . . . . . . . . . . . . . . . . . . . . 92 4.2. Histograms of sample photon collection data . . . . . . . . . . . . . . . . 95 4.3. Variances of sample photon collection data . . . . . . . . . . . . . . . . . 98 4.4. Atom counting with CCD camera . . . . . . . . . . . . . . . . . . . . . . 99 4.5. Sample data for bayesian fluorescence number estimation . . . . . . . . . 104 4.6. Bayesian algorithm flow chart . . . . . . . . . . . . . . . . . . . . . . . . 111 5.1. Excited state populations for 87Rb and V-atom . . . . . . . . . . . . . . 130 5.2. 1D forces on 87Rb and V-atom . . . . . . . . . . . . . . . . . . . . . . . 133 5.3. 3D MOT forces for an assortment of beam phases . . . . . . . . . . . . . 135 5.4. 3D MOT phase dependence . . . . . . . . . . . . . . . . . . . . . . . . . 142 6.1. Sampled oscillation spectrum . . . . . . . . . . . . . . . . . . . . . . . . 159 6.2. Assorted measured spectra . . . . . . . . . . . . . . . . . . . . . . . . . 166 6.3. Single-atom temperatures with release-recapture method . . . . . . . . . 168 6.4. Fluorescence amplitude scaling with current amplitude . . . . . . . . . . 168 6.5. Simulated 1D modulation slopes . . . . . . . . . . . . . . . . . . . . . . 170 xi Figure Page 6.6. Fluorescence amplitude spectrum . . . . . . . . . . . . . . . . . . . . . . 174 6.7. Lorentzian fluorescence amplitude spectrum . . . . . . . . . . . . . . . . 176 6.8. Fluorescence amplitude scaling with field gradient . . . . . . . . . . . . . 177 6.9. Simulated fluorescence amplitude scaling with field gradient . . . . . . . 178 6.10. Fluorescence amplitude as a function of laser detuning . . . . . . . . . . 180 6.11. Fluorescence amplitude with background magnetic fields . . . . . . . . . 182 6.12. Spectra of multiple atoms in a MOT. . . . . . . . . . . . . . . . . . . . . 185 6.13. Images of Single Atom Parametric Resonance . . . . . . . . . . . . . . . 189 6.14. Light Distribution for Parametric Oscillating Atom . . . . . . . . . . . . 191 6.15. Fitted Light Distributions for Parametric Oscillating Atoms . . . . . . . 193 6.16. Light intensity loss due to parametric oscillations . . . . . . . . . . . . . 195 6.17. Power spectra for different waveforms . . . . . . . . . . . . . . . . . . . 199 B.1. Laser power control with polarizers . . . . . . . . . . . . . . . . . . . . . 213 C.1. Water cooling channel dimensions . . . . . . . . . . . . . . . . . . . . . . 219 xii LIST OF TABLES Table Page 3.1. Permanent Ring Magnet Parameters . . . . . . . . . . . . . . . . . . . . 67 3.2. Helmholtz Coil Field Gradients . . . . . . . . . . . . . . . . . . . . . . . 79 3.3. Photon Collection Lens System Collection Efficiency . . . . . . . . . . . 83 5.1. Repumping field and populations of 87Rb D2 energy levels . . . . . . . . 115 5.2. MOT Beam Circular Polarizations . . . . . . . . . . . . . . . . . . . . . 120 6.1. High-gradient MOT parametric resonance conditions . . . . . . . . . . . 187 6.2. Variance compared to amplitude for common waveforms . . . . . . . . . 201 C.1. Water-cooled MOT coil parameters . . . . . . . . . . . . . . . . . . . . . 218 xiii CHAPTER I INTRODUCTION The magento-optical trap (MOT) has become one of the bedrocks for research on the quantum behavior of atoms. The MOT can produce millions of atoms with temperatures on the order of microkelvins. Such small temperatures are necessary to limit atomic motion for studying their classical and quantum dynamics. The magneto-optical trap uses multiple laser fields whose frequencies are often a few MHz smaller (red-detuned) than an atomic resonance of the atomic species being trapped. The light is lower in frequency so that atoms moving towards the laser source sees a Doppler shift moving the light from that laser closer to resonance. This Doppler shift makes the atom more likely to absorb photons from the laser (due the reduced detuning), resulting in radiation pressure that pushes the atom in the propagation direction of the light. With D + 1 lasers for a MOT in D-dimensions, this can result in a cooling force as the lasers damp the motion of the atoms. This process creates what is often referred to as optical molasses and was originally conceived in the 1970s [1, 2] and experimental verified in the following decade [3]. Doppler cooling is only responsible for cooling atoms; it does not trap them. In addition to the laser fields, a MOT requires a (quadrupole) magnetic field which produces a spatially dependent Zeeman shift of atomic energy levels. The Zeeman shifts provide in an additional preferential excitation of the atoms by the laser, creating a restoring force that traps atoms near the location where the magnetic field vanishes. Together with optical molasses, the quadrupole fields impart a force on the atom which causes it behave as a damped harmonic oscillator. There are a number of configurations for a magneto-optical trap, but this work focuses on the trap shown in Figure 1.1. Here, three pairs of counter-propagating lasers push the 1 a) ẑ b) current + ŷ x̂ + + current FIGURE 1.1. The magneto-optical trap. a) Schematic drawing. b) Photograph of our MOT setup around the experiment vacuum cell. For both images, blue arrows show the six-counter propagating MOT beams (with their appropriate polarizations in the schematic drawing). Orange loops are anti-Helmholtz coils that generate the linear magnetic field near the origin. The MOT loads at the origin (red dot). atom towards their intersection point and the magnetic field is generated by a pair of anti-Helmholtz coils—coaxial coils with currents traveling in opposite directions. At the midpoint between the two coils on their central axis, the magnetic field vanishes, establishing the equilibrium position for the atom. 1.1 Single Atom MOTs Not long after the first MOTs were developed [4], they were extended to allow capture of small numbers of atoms, primarily by greatly increasing the strength of the confining magnetic field. This produced traps on order of tens of atoms [5] and quickly down to individual atoms [6]. Since then, single- or few-atom MOTs have largely been used as efficient sources of single atoms for loading into other optical systems [7]. These systems include cavity QED experiments [8, 9], which allow for strong coupling between the atom and cavity optical field modes; optical dipole traps 2 [10, 11], which generally have tighter confinement of atomic motion than MOTs and are far-detuned from atomic excitation; and one-dimensional optical lattices [12, 13], which allow for targeted experimental interaction with multiple atoms. There have been a broad field of research looking at atomic counting and MOT characterization with single (and low-numer) MOTs [14, 15]. Other uses of single atom MOTs have included studies of cross-atomic-species cold-atom interactions [16, 17], rare-isotope separation [18, 19], and detailed studies of the high-gradient MOT loading and loss mechanisms [20, 21]. Besides these studies of loading and loss mechanisms, there is little experimental research on the dynamics of a few atoms in a magneto-optical trap. Additionally, these have investigated loss rates statistically as opposed to the dynamics which causes atomic loss in the traps. For few-atom MOTs, these loses are due to collisions between atoms in the MOT resulting in atoms exiting the trap [15]. Some atomic collisions coincide with atomic energy transitions that provide enough kinetic energy for the atoms to escape the MOT [22]. The only experiments that have looked at the dynamics of a single atom in a magneto-optical trap examined correlations between photons emitted by the atom [23, 24]. These works reveal temporal-correlations that reflect both internal atomic dynamics (Rabi oscillations) and external dynamics (position-dependent electric field intensity and polarization) over several orders of magnitude. The work discussed in this dissertation adds to this little-explored topic by looking at position-dependent oscillations of the atom in a MOT magnetic field. The fluorescence rate of an atom depends on the detuning from resonance of the exciting laser field. Because of the linear magnetic field of the quadrupole, the Zeeman shifted energy levels have detunings that vary spatially. Modulating this magnetic field 3 introduces oscillations in the detuning and thus the rate of fluorescence from the atom. Measuring the fluorescence oscillations can produce a time-averaged position distribution for the atom in the MOT. The position distribution of the atom is very closely related to its potential energy, which can be used, via the equipartition theorem [25–27], to measure the atomic temperature. Therefore, this work functions as an in- situ temperature measurement of the atom in addition to examining the averaged motion of the atom in the MOT. 1.2 MOT Temperature Measurements There are already a few established methods for measuring atomic temperatures in a MOT. Some of the methods, like the method proposed here, compare measurements of the atom(s) in the MOT to models of the MOT in order to extract potential energy information about the atom. Others methods more directly measure the temperature through the atomic kinetic energy, but these methods are lossy, requiring releasing the atom(s) from the MOT and reloading new atoms for experiments. One such lossy method is the release-recapture method which was used to estimate the MOT temperature in the first successful MOT publication [4]. This method turns off the trapping fields in order to allow the trapped atoms to expand from the trap ballistically. Turning the trap on again, after a given amount of time, recaptures just a fraction of the atoms. Comparing trap-off times with recapture fraction gives an estimate of average atomic velocity, and hence temperature in the MOT. In addition to measuring the temperature of traps with large numbers of atoms, this technique has be used to measure the temperature of a single atom both in a MOT and in a dipole trap [16, 28, 29]. 4 The time of flight technique releases atoms from a MOT and allows them to fall under gravity through a near-resonant laser field [30]. As the atoms pass through the field, the light they scatter is measured. Observing how this fluorescence changes as a function of time from release (atoms moving directly downward initially pass through the laser before atoms that were initially moving directly upward) gives an estimate of the average velocity of the atoms, and hence their temperature. This technique can be used for other systems including atoms in an optical lattice [31]. Other configurations of the time of flight technique use an additional laser beam to push the atoms in some direction where the probe beam has been located [30]. Pushing the atoms vertically upward takes advantage of converting kinetic energy to gravitation potential energy in order to measure a maximum height for the MOT atoms to reach, giving a measure of their initial kinetic energy. Pushing the atoms horizontally lets gravity drag the atoms downward, under the probe beam, to measure a travel distance for the atoms—and therefore a maximum horizontal velocity distribution. Modified time-of-flight techniques have been used to measure the temperature of single atom [29], although due to the difficulty of imaging scatter from a single atom, they are less commonly used than release-recapture methods. For the single-atom measurement, instead of detecting the light of an atom as it passes through a nearby beam, the position of the atom is detected on a CCD after a resonant imaging pulse. Repeating the test provides information of the spatial distribution of the atom after release, allowing velocity and temperature to be estimated for an atom initially inside the MOT. Another lossy method adiabatically reduces the optical potential in which an atom resides [16, 28, 32]. Measuring the probability that the atom remains in the trap at a given potential energy gives an estimate of its kinetic energy. This 5 method is similar in concept to evaporative cooling techniques used, as an example, in Bose-Einstein condensates [33]. While evaporative cooling is used to decrease the temperature and increase the density of an atomic cloud, the adiabatic lowering technique just probes the temperatures of a small number of atoms. For temperature measurements, this technique is mostly used in dipole traps where the potential can be easily reduced by lowering the intensity of the trapping laser [4, 16, 34]. One method that preserves the number of atoms in the MOT looks at the frequency spectrum of photons emitted by the atoms. The motion of atoms in the trap will broaden the wavelength of the emitted light via the Doppler effect. Measuring this broadening allows for an estimate of the velocity of the atoms [35]. This is also used in measuring the temperature of trapped ions [36]. Ion traps can also make use of quantized motion to measure spectra and temperature [36]. Spectra from ions (including single ions [37]) reveal sidebands of resonance peaks. The number and relative intensities of the sidebands are related to the average vibrational mode of the ion in the trap, and hence to temperature. This technique can also be used to measure temperatures of neutral atoms in optical lattices [38]. Another number-preserving method takes advantage of the harmonic-oscillator model of the MOT and the equipartition theorem. This method uses an external force (created either an oscillating, external uniform magnetic field [26] or additional laser beam [27]) to drive oscillations in the center of mass of the MOT. Measuring the amplitude responses at different frequencies gives the natural frequency, and thus the spring constant, of the restoring force in the MOT. With a measurement of the RMS radius of the MOT via pictures of it, the average potential energy of the atoms is revealed. As a temperature estimate, this method is similar to our proposed technique 6 as it measures temperature via external oscillations, but this method is not applicable to single-atom MOTs where there is not a well defined size of the atomic cloud. 1.3 Optical Trap Oscillations In addition to measuring MOT spring constants with oscillations, oscillations of laser fields have been used to measure properties of particles in other forms of optical traps. For beads in an optical dipole trap (see Section 2.2.4), oscillating the power of the laser which confines the atom can excite resonances in the bead [39]. As with the MOT spring measurements above, Imaging the bead’s motion with these oscillations reveals the trapping strength on the bead. Similar experiments have probed the trapping potential for atoms in an optical lattice, in which the interference of multiple laser fields creates a periodic lattice of positions where atoms are trapped [40, 41]. Here, modulations of the lattice’s trapping beam power also modulates the trapping potential for the atoms. Measuring the populations of atoms still present in the lattice after being driven at various frequencies can reveal the vibrational states of the lattice [40]. Additional, the modulations can be seen in changes in power measured from beams diffracted by the atoms arranged in the lattice [41]. For these two purely optical traps, a magnetic field is not necessary, requiring that modulations be driven by oscillations in laser power. Attempting to detect oscillations in fluorescence from the trapped particles, then, would be difficult as the signal would be swamped by oscillations in background fluorescence levels. Instead, these experiments (save for the diffraction experiment in [41]) directly imaged the particles in the trap with a camera to observe oscillations. This can be challenging for a single atom, although our experiment does reveal similar oscillations for the single atom when imaged with a CCD camera (see section 6.5). Without modulating 7 the beam power, as in our experiment, we instead detect oscillations directly through measurements of photon arrivals from an atom. 1.4 Dissertation Outline The layout of this dissertation is as follows. Chapter II has a theoretical description of the interaction between atoms, light and magnetic fields, building to a description of the functioning of a MOT. Chapter III describes the experimental apparatus, focusing on relevant changes made for the single-atom experiments described in later Chapters. Chapter IV discusses our single-atom MOT and examines a new technique for monitoring and controlling experiments based on a single atom. Chapter V expands on the theory in Chapter II to look more closely at the behavior of atoms with the complete electronic structure of the D2 transition for rubidium. Chapter VI looks at position and temperature measurement experiments performed on our single atoms. Finally, Chapter VII draws conclusions for our experiment and briefly lays out the path forward for future investigation. 8 CHAPTER II ATOM OPTICS AND MAGNETO-OPTICAL TRAPS In this Chapter, the interaction of atoms with light and magnetic fields is sketched using a standard semiclassical picture in which the atom’s internally energy is quantized but its external motion and external fields are treated classically. Calculations are done in one-dimension with the atom is treated as having a single ground state and a small number of excited states when appropriate. Details of the calculation are given with an eye toward the three-dimensional picture in Chapter V with a full D2 transition of 87Rb. After examining atomic interactions with electric and magnetic fields individually, an atom inside a magneto-optical trap is discussed. 2.1 A Single Atom Until the broader discussion of magneto-optical traps in Section 2.5, the atom will be treated as a qubit with energy separation ~ω0. Under this assumption, the atomic Hamiltonian should have the form ĤA = ~ω0 |e〉〈e| (2.1) with the quantum state of the atom in the form |ψ〉 = ce|e〉+ cg|g〉 (2.2) where |e〉 and |g〉 are the atomic excited and ground states, respectively. In this definition, the ground state energy is defined to be zero. Rather than working with the atomic wavefunction, later calculations are simplified by using the atomic density 9 operator [42]. defined by     |c |2 c c∗ ρ ρ ρ = |ψ〉〈ψ|  e e g  e,e e,g=   =   . (2.3) c ∗ 2gce |cg| ρg,e ρg,g s In the Schrödinger picture, the density operator evolves under the equation ( ) ( ) d d ρ = |ψ〉 〈ψ|+ | dψ〉 〈ψ| dt dt dt i i = − Ĥ |ψ〉〈ψ|+ |ψ〉〈ψ| Ĥ ~ [ ] ~d i ρ = − Ĥ, ρ . (2.4) dt ~ Note that this equation is identical to the time evolution of an operator in the Heisenberg picture. The use of density matrices must be implemented to look at mixed states—quantum states which cannot be simple written as a linear superposition of eigenstates of a Hamiltonian [43]. Such states appear, for example, in analysis of entanglement [44], teleportation [45], and when looking at quantum trajectories [46, 47]. The density operator also is beneficial as operator expectation values are calculated simply by tracing the atomic states over the product of the operator applied to the atomic density operator as [ ] ∑ 〈Â〉 = Tr Âρ = 〈n|Âρ|n〉 (2.5) n where |n〉 form a complete basis to describe eigenstates of the system. For analysis of an atom in the MOT, the use of the density operator is important in modeling spontaneous emission of photons from the atom through the Lindblad 10 superoperator [48] defined as 1 ( )L [σ̂] ρ = σ̂ρσ̂† − σ̂†σ̂ρ+ ρσ̂†σ̂ (2.6) 2 where σ̂ = |g〉〈e| and σ̂† = |e〉〈g| are the atomic raising and lowering operators respectively. For the two level atom, the superoperator simplifies to L [σ̂] ρ = ρe,e|g〉〈g| − 1 (|e〉〈e|ρ+ ρ|e〉〈e|) . (2.7) 2 Including this operator, the evolution of an atom which undergoes spontaneous emission follows d − i [ ] ρ = ĤA, ρ + ΓL [σ̂] ρ, (2.8) dt ~ where Γ is the decay rate of the atom. 2.2 Atoms and Light A few simplifying assumptions are made to analyze the interaction of an atom with light, following the methodology of [48]. The light, for now, is treated as a linearly polarized electric field of a single mode. The atom is treated as small enough that the spatial variation of the electric field can be ignored. Thus the light field has the form E~ (t) = ̂E0 cos(ωt+ φ), (2.9) where ω is the frequency of the light, assumed to be close to the transition frequency of the atom, ω0, and ̂ is the polarization direction of the light. These assumptions are sufficient for finding a form for the interaction Hamiltonian between an atom and 11 light, but removing some of these assumptions leads to interesting results, which are discussed when appropriate. 2.2.1 Interaction Hamiltonian The interaction between the atom and electric field is treated as a dipole interaction with an atomic dipole operator ( ) d̂ = 〈e|d~|g〉 σ̂† + σ̂ , (2.10) where d~ = e~r is the dipole moment of the atom (~r is the position operator for the atom’s electron). This definition of d̂ derives from treating the electron position operator as that of a harmonic oscillator, where r̂ is proportional to the sum of the oscillator raising and lowering operators [49], which here correspond to atomic raising and lowering operators, σ̂† and σ̂. The expectation value of the dipole operator is ( ) 〈d̂〉 = 〈e|d~|g〉 〈σ̂†〉+ 〈σ̂〉 . (2.11) With the atom treated as a dipole, we can find an interaction Hamiltonian by comparison tot the energy of an electric dipole interacting with a field. This provides an interaction Hamiltonian ĤAF = −E~ · d̂. (2.12) In the absence of the electric field, the excited state population evolves as e−iω0t and 〈σ̂〉 evolves in this same way. With this, the dipole operator can be written in the form d̂ = d̂+ − d̂+ (2.13) 12 with d̂− ≡ 〈e|d~|g〉σ̂† ∼ eiω0t and d̂+ ≡ 〈e|d~|g〉σ̂ ∼ e−iω0t. Writing the electric field in terms of complex exponentials gives ~ E0 [ ] E(t) = ̂ e+iωt+φ + e−iωt−φ ≡ E~− + E~+ (2.14) 2 and is used to write the interaction Hamiltonian as ( ) HAF = − d̂− · E~− + d̂− · E~+ + d̂+ · E~− + d̂+ · E~+ . (2.15) In terms of the exponentials, these four terms are proportional to ei(ω0+ω)t, ei(ω0−ω)t, e−i(ω0−ω)t and e−i(ω0+ω)t respectively. In the limit where the frequency of light is very close to that of the atomic energy, the terms with the frequency differences oscillate much more slowly than the terms with their sum (the first and last term). The rotating wave approximation ignores these quickly oscillating terms, so that the interaction Hamiltonian becomes H = −d̂− · E~+AF − d̂+ · E~−. (2.16) Written in terms of atomic raising and lowering operators, the Hamiltonian is ~ ( ∗ − )HAF = Ω σ̂e iωt + Ωσ̂†e+iωt , (2.17) 2 where 〈g|̂ · d~|e〉E e−iφ0 Ω = (2.18) ~ 13 is named the Rabi frequency. The excited state population for any two level system at rest and interacting with a non-resonant electric field will oscillate with this frequency [42, 50, 51]. In an alternate view of the rotating wave approximation, the electric field is quantized as a harmonic oscillator. Under this view, the (single frequency, polarization and mode) electric field follows E~ ∝ f~(~r)â+ f~∗(~r)â†, (2.19) where f~(~r) are the spatial mode functions of the field, and ↠is the photon creation operator for this mode (expanding to a general electric field requires summing over this term for each frequency, mode, and polarization). With this picture, the interaction Hamiltonian is Ĥ ∝ âσ̂ + âσ̂† + â†σ̂ + â†AF σ̂†. (2.20) The first term of this equation removes a photon from the field and lowers the atom from excited to ground state. The last term adds a photon to the field and raises the atom into the excited state. Both of these are non-energy conserving and can be dropped1. These two terms are the same as the quickly rotating terms which were dropped previously. 1These terms, while ignorable here, can be viewed as atomic interaction with the quantum vacuum [52]. This interaction leads to phenomenon such as the Casimir-Polder affect [53] and the Lamb shift [42, 54]. 14 2.2.2 Optical Bloch Equations The evolution of the atomic wavefunction in the Schrödinger picture is d | iψ〉 = − (H +H ~ A AF ) |ψ〉 (2.21) dt which produces equations iωt ċg = − iΩe c2 e (2.22) − iΩ∗ iωtċe = iω e0ce + c .2 g These can be simplified in a rotating frame where ce is transformed to c̃ee −iωt to produce ċ = − iΩg c2 e (2.23) ċ = −i(ω − ω)c + iΩ∗e 0 e c2 g (note, here I’ve dropped the c̃e notation). Eventually moving to this frame is what motivated writing the dipole operator in terms of complex exponentials in Equation 2.13. These same equations could have been derived from an atomic Hamiltonian of the form ĤA = −~∆ |e〉〈e| (2.24) and interaction Hamiltonian ~ ( ∗ †)ĤAF = Ω σ̂ + Ωσ̂ , (2.25) 2 where ∆ = ω − ω0 is the detuning of the electric field from the atomic resonance frequency. These are the forms of the atomic and interaction Hamiltonians that will be used throughout this text. 15 From these forms, the time evolution of the atomic density operator, calculated from Equation 2.8, produces the optical bloch equations [42] ρ̇e,e = −Γρ ie,e − (Ωρ ∗g,e − Ω ρ2 e,g) ( ) ρ̇ = − Γ iΩe,g + i∆ ρe,g − (ρ2 2 g,g − ρe,e) ( ) (2.26)∗ ρ̇ Γ iΩg,e = − − i∆ ρg,e + (ρ2 2 g,g − ρe,e) ρ̇ i ∗g,g = Γρe,e + (Ωρg,e − Ω ρe,g) .2 In the steady state, these have analytic solutions ρSS = |Ω| 2/Γ2 e,e 1+(2∆/Γ)2+2|Ω|2/Γ2 ρSS = − iΩ 1+2i∆/Γe,g Γ 1+(2∆/Γ)2+2|Ω|2/Γ2 (2.27) SS iΩ∗ρ = 1−2i∆/Γg,e Γ 1+(2∆/Γ)2+2|Ω|2/Γ2 1+(2∆/Γ)2ρSSg,g = .1+(2∆/Γ)2+2|Ω|2/Γ2 Recall that ρe,e is the population of the excited state ρ 2 e,e = |ce| from Equation 2.2. With this definition, the rate that an atom scatters photons is R = Γρe,e. (2.28) 16 2.2.3 Radiation Pressure The force acting on an atom is the time evolution of the atom’s (classical) momentum p~. Following the Heisenberg picture, this is ~ i [ ] F = Ĥ, p~ = −∇~ Ĥ. (2.29) ~ To look at the force of light on the atoms, we use the Hamiltonian 2.25. This gives ~ −~ ( ) F = ∇~ Ω∗σ̂ +∇~ Ωσ̂† . (2.30) 2 From the definition of the Rabi frequency in Equation 2.18, Ω ∝ E iφ0e , where E0 is the field magnitude and φ is its phase. Both of these can depend on space, but for now consider the electric field to be a plane wave propagating in the +ẑ direction so that the electric field strength has the form E0(~r) = E +ikz 0e . (2.31) Using this in Equation 2.30 gives a straightforward equation for the force ( ) ~ −~ 2dE0F = ∇~ e−ikz 2dE0σ̂ + ∇~ e+ikzσ̂† 2 ( ~ ~− )~ 2dE e ikz 2dE e+ikz = − − 0 0ikẑ σ̂ + ikẑ σ̂† 2 ~ ~ ( F~ ik~ = − −Ω∗ ) σ̂ + Ωσ̂† ẑ. (2.32) 2 17 Taking the expectation value of the force gives 〈~ 〉 −ik~ [( ) ] F = Tr −Ω∗σ̂ + Ωσ̂† ρ ẑ 2 −ik~= (−Ω∗ρe,g + Ωρg,e) ẑ (2.33) 2 after using Equation 2.5. The evolution for the excited state population from the optical bloch equations in 2.26 can be solved in steady state to get i ( ) ΓρSS = − −Ω∗ρSSe,e + ΩρSS .2 e,g g,e Comparing this to the expectation value for the force, the steady-state force on the atoms is 〈F~ 〉SS = ~kΓρSSe,e ẑ. (2.34) This is an insightful equation as Γρe,e is just the scatter-rate of photons by the atom as in Equation 2.28, and ~kẑ is the momentum carried by each photon. Thus the force felt by atom is just the average rate it absorbs momentum from scattered photons. The momentum change from emitting photons goes to zero in the limit of many absorption-emission events as the emissions have random directions. This force due to absorbed photon momentum is commonly referred to as radiation pressure. In addition to being the basis for optical molasses (see Section 2.4.1 below), radiation pressure has been used experimentally to launch atoms in atomic clocks [55, 56], cool micromechanical resonators [57, 58], and even macroscopic objects such as the first solar sail successfully flown by the Japan Aerospace Exploration Agency [59, 60]. 18 2.2.4 Dipole Traps Returning to Equation 2.30, if the assumption is made that the electric field magnitude, rather than the phase, depends on space, the equation for force can be written as ( − )iφ iφ F~ = −~ 2de ∇~ 2deE0σ̂ + ∇~ E σ̂†0 2 ( ~ ~ ) −~ 2dE e −iφ 0 ∇~ E0 2dE0eiφ∇~ E0 = σ̂ + σ̂† 2 ~ E0 ~ E0 ~ ( ) = − Ω∗∇~ log[E0]σ̂ + Ω∇~ log[E ]σ̂†0 . (2.35) 2 As done previously, the expectation value for the force is 〈F~ 〉 −~= ∇~ log[E0] (Ω∗ρe,g + Ωρg,e) . (2.36) 2 This equation is real, as the sum of the complex terms produces just 2Re [Ωρg,e]. From the steady state Bloch equations, this is ( ) 4∆ |Ω|2∗ ss ss /Γ2Ω ρe,g + Ωρg,e = (2.37) 1 + (2∆/Γ)2 + 2 |Ω|2 /Γ2 where, again, ∆ is the detuning of the electric field from the atomic energy, Ω is the Rabi frequency and Γ is the atomic decay rate. Here, it is convenient to introduce a saturation parameter defined as 2 |Ω|2 /Γ2 s = (2.38) 1 + (2∆/Γ)2 19 so that ( ∗ )ss ss 2∆sΩ ρe,g + Ωρg,e = . (2.39)1 + s Then, the force becomes 〈~ 〉 − ~∆sF = ∇~ log[E0]. (2.40) 1 + s The definition of s is proportional to the field intensity, which goes as E20 . As field intensity is a more common parameter for experiments than field magnitude (the (integrated) electric field intensity of a laser is proportional to beam power), writing ∇~ log[E0] in terms of the saturation parameter comes from ∇~ s = 2s∇~ log[E0]. (2.41) This produces a force equation just in terms of the saturation parameter ~ 〈~ 〉 − ~∆∇s −~∆F = = ∇~ log[1 + s]. (2.42) 2(1 + s) 2 For large detunings where there is little excitation of the atom by the field, s  1. This simplifies the force to ~ 〈F~ 〉 ≈ − ~∆∇s = −~∆∇~ s, (2.43) 2(1 + s) 2 so that the force on the atom is directed towards regions of high intensity (large s). A common field where the beam intensity depends on position is a pair of tightly focused gaussian beams. This field has electric field intensity [61] [ ] [ 2P ] −(2 (x 2 + y2) I(~r) = exp ) (2.44) πw20 1 + (z/z ) 2 0 w20 1 + (z/z0) 2 20 where P is the total beam power, w0 is the beam waist, the beam radius at its focus, and z0 is the Rayleigh range of the beam, which gives a measure of the length of the focus size along the beam axis and is defined as the distance where the beam radius √ grows from its minimum by a factor of 2. For this field, the force is given by [ ( ) ] 〈~ 〉 2~∆s w 2 2(x2 + y2) F = xx̂+ yŷ + 0 − zẑ . (2.45) w20 [1 + (z/z 2 0) ] z2 z20 0 [1 + (z/z 2 0) ] With the laser detuned below resonance, ∆ < 0, this is clearly a restoring force. This specific arrangement of focused gaussian beams is often called a dipole trap. 2.3 Atoms and Magnetic Fields The interaction of an atom with a magnetic field is a magnetic dipole interaction based on the total angular momentum of the atom, including the orbital angular momentum L~ , spin S~ and nuclear angular momentum I~. The combination of these give the hyperfine structure of the atom, quantized with F~ = L~ + S~ + I~. (2.46) The D2 transition of 87Rb, on which our MOT is based, has transitions between the 52S1/2 ground state and the 5 2P3/2 excited states. Both of these states have S = 1/2. The ground state has L = 1/2 and the excited state has L = 3/2. Together with 87Rb’s nuclear spin of I = 1/2, the excited state can have total angular momentum F with values between 0 and 3, and the ground state can have values of either 0 or 1 [62]. This is shown in the level diagram of Figure 2.1. The frequency splitting between the hyperfine levels is on the order of 102 MHz while energy shifts due to the magnetic field (Zeeman shifts) of the MOT is on the order of MHz. With such 21 |E; F = 3i 52P3/2 |E; F = 2i |E; F = 1i |E; F = 0i |G; F = 2i 52S1/2 |G; F = 1i FIGURE 2.1. 87Rb D2 Transition Level Diagram. Labeled at the transitions address for our MOT. a small shift of the levels due to the magnetic field, the total angular momentum F~ is a fair quantum number to use to study the interaction between the atom and the magnetic field [62]. Here, we’ll treat the |G;F = 2〉 and |E;F = 3〉 levels as the only two levels of our atom. A fuller picture of the hyperfine atom is discussed in Chapter V. The atom is treated as a magnetic dipole that is aligned with its angular momentum µ~ = µ ~BgFF , (2.47) where µB is the Bohr magneton. Much like the electric dipole interacting with the electric field of light, the interaction between the magnetic dipole and magnetic field is then [63] Ĥz = −µ~ ·B~ . (2.48) 22 Trapping Repump Working in one dimension, assuming that B~ lies along the z-axis, the interaction simplifies to Ĥz = −µBgF F̂zBz when the z-axis is also the angular momentum quantization axis of the atom. F̂z is the projection operator of the atomic angular momentum along the quantization axis. Writing this explicitly gives ∑∑∑ Ĥz = −µBBz gFmF |s;F,mF 〉〈s;F,mF | (2.49) s=e,g F mF where s is the excited or ground state of the atom. From this, it is clear that the magnetic field is responsible for shifts of the energy levels of atoms, the Zeeman shifts [64]. The quantities gF are Landé g-factors [62]. They result from perturbative approximations made when the shift in atomic energy levels due to the magnetic field is much smaller than the hyperfine splitting. In this case, the total atomic angular momentum F~ , as defined in 2.46, serves as a good quantum number. In 87Rb, the hyperfine splitting is on the order of gigahertz for the ground states and hundreds of MHz for the excited states. The Zeeman splitting (per Gauss) is on the order of 1MHz/G. With magnetic fields in the MOT on the order of tens of Gauss, this condition is met,which is good since all For 87Rb’s D2 transition, all excited states have gFe = 2/3 and the ground states have gFg=2 = 1/2 and gFg = −1/2 [62]. When a 3D model of the atom is discussed in Chapter V, this assumption about the magnetic interaction will remain true—the atom’s quantization axis will align with the magnetic field direction at every location in space. This will require rotation of the lab frame to align the z-axis of the lab frame with the magnetic field. This is 23 the largest complication of the 3D model as the polarization directions for each of the six MOT beams must be correctly written in a circular and linear polarization basis. This rotation is discussed in Section 5.1.3. 2.3.1 Magnetic Trapping A particle with a magnetic dipole momentum µ~ in a magnetic field will have a potential energy U = −µ~ · B~ . If this energy is spatially varying, the particle will experience a magnetic force F~ = ∇~ (µ~ ·B~ ) (2.50) As done above (and as will be assumed while discussing the MOT), when the dipole ∣∣ ∣∣ moment is aligned with the field the interaction energy is U = µ ∣B~ ∣. Under these circumstances, the force is then ~ − ∇~ ∣ ∣ ∣ F = µ ∣B~ ∣∣ , (2.51) so that the force is zero where the magnetic field strength vanishes. This force must also apply to an atom with magnetic dipole moment [65, 66] and is critical to evaporative cooling for Bose-Einstein condensates [67]. As done for radiation pressure, the force is examined quantum mechanically by Equation 2.29 with the Hamiltonian 2.49. An atom in the steady state experiences a force ( ) 〈~ 〉 ∂B ∑∑∑ z ∣∣ ∣∣2F = µB ẑ g ssFmF cs,m , (2.52)∂z F s=e,g F mF where the energy level populations are written as their wavefunction coefficients rather than density operator elements (strictly for simplification of subscripts). The 24 relationship to the classical magnetic force of 2.51 is a bit unclear as it appears uniform. However, the steady state populations have spatial dependence (see Figures 5.1 for a graph of the populations as a function of position), which allows for magnetic trapping. For a linear magnetic field B~ = −B′zzẑ, as will be discussed for the MOT, the force is ∑∑∑ ′ ∣∣ ∣F~ = −µ B ẑ g m css ∣2B z F F s,m . (2.53)F s=e,g F mF For states where gF > 0, when z < 0 (so that B > 0), the energy levels with mF < 0 will be preferentially populated as their energies will be reduced by Zeeman shifting. This results in the double sum producing a negative value, given an overall force in the positive z-direction. When z > 0, the mF > 0 energy levels are preferentially populated and produce a force in the negative z-direction. Thus, the overall force is to locate the atom near z = 0, just as in the classical case. This results holds true when gF < 0, with the preferential population switching positive and negative values for mF . It is important to note here that this magnetic trapping is distinct from the magnetic confinement discussed in Section 2.5.1. The trapping here results from the minimization of the magnetic dipole energy. The confinement trapping results from spatially preferential photon absorption due energy level Zeeman splitting. 2.4 The Fg = 0→ Fe = 1 Atom For the remainder of this Chapter, the focus will be on a Fg = 0→ Fe = 1 atom, that is one having no ground state angular momentum and an excited state with a total angular momentum of 1. With the angular momentum formalism in Equation 25 |E; mF = 1 i |E; mF = 0 i |E; mF = +1 i |i |0 i |+ i |i |+ i ⌦0 ⌦ ⌦ + + |G; mF = 0 i |g i |g i (a) (b) (c) FIGURE 2.2. Fg = 0 → Fe = 1 Atom Level Diagrams. (a) Full labeling for states of the atom. (b) Simplified naming conventions used in the text. The electric fields coupling excited and ground states are labeled with their Rabi frequencies as give in Equation 2.56. (c) V-atom reduction when there is no electric field to excite the |0〉 state. 2.46, this could correspond to a spin-1/2 atom (S = 1/2), orbital angular momentum L = 1/2 and no nuclear spin (I = 0). With no nuclear spin, the angular momentum vector J~ = L~ + S~ is often used rather than F~ , but F~ used here for more directed comparison to the full atomic energy levels discussed in Chapter V. The level diagram for such an atom is shown in Figure 2.2a. For this atom, the levels will be labeled as shown in Figure 2.2b. The atom has density operator    ρ−− ρ−0 ρ−+ ρ−g    ρ0− ρ  00 ρ0+ ρ0g ρ =     . (2.54) ρ+− ρ+0 ρ++ ρ+g  ρg− ρg0 ρg+ ρgg Each excited states has its own raising operator defined, the three being defined as σ†− = |−〉〈g| , σ†0 = |0〉〈g| , and (2.55) σ†+ = |+〉〈g| . 26 The excited and ground states are coupled via electric fields E~ , E~+ 0 and E~−, named relative to the excited states they couple (note this names are opposite the naming conventions in Equation 5.9). These fields are circularly polarized or linearly polarized. Following the derivation of Equation 2.25, this leads to an atom-field coupling Hamiltonian ~ ( ) ~ ( ) ~ ( ) ĤAF = − Ω∗ †−σ̂− + Ω−σ̂− − Ω∗ † ∗ †2 2 0σ̂0 + Ω0σ̂0 − Ω σ̂+ + Ω+σ̂ . (2.56)2 + + It is also assumed that the electric fields are all detuned from resonance ω0 by some amount. Following the derivation in 2.24, the atomic Hamiltonian can be written as ĤA = −~∆+ |+〉〈+| − ~∆0 |0〉〈0| − ~∆− |−〉〈−| . (2.57) The magnetic field Hamiltonian is given directly from Equation 2.49 and for the Fg = 0→ Fe = 1 atom is Ĥz = µBgFBz (|+〉〈+| − |−〉〈−|) , (2.58) as evident from the mF values shown in Figure 2.2a. Defining ∆B = µBgFBz/~, this can be written as Ĥz = ~∆B |+〉〈+| − ~∆B |−〉〈−| . (2.59) Each of the excited states can spontaneously decay and it assumed this occurs at the same rate, Γ, for each state. The evolution of the atomic density operator follows the equation d i [ ] ρ = − ĤA + ĤAF + Ĥz, ρ + ΓL [σ̂−] ρ+ ΓL [σ̂0] ρ+ ΓL [σ̂+] ρ (2.60) dt ~ 27 based on the density operator evolution definition of 2.8. The form of the Lindblad superoperator of Equation 2.7 reveals that spontaneous emission from one excited state of the atom depends only on the populations of that excited state. With these separate decay paths, the impact on the evolution of the state of the atom from each state is independent as written above. The time evolution of each density operator element ρi,j is listed in Appendix A.2. Our magneto-optical trap consists of pairs of counter-propagating, circularly polarized lasers, as shown in Figure 2.5. The counter-propagating beam have the same circular polarization in their reference frame, but in the reference frame of an atom (with the beams moving towards it from different directions) the two beams have opposite polarizations, corresponing to only the E~− and E~+ electric fields. Matching the model to our experiment sets E~0 → 0. As discussed in Appendix A, doing this effectively simplifies the atom to the V-atom, shown in Figure 2.2c. Without the linear field component to excite the |0〉 state, the x‘x‘population of this state will decay quickly to zero and can be ignored. The remainder of this Chapter will assume this simplification of the atomic structure. The equations of motion for the internal structure of the V-atom are given in equations A.4. In these equations, for simplicity, the excited state energies Zeeman energy shifts are left out. They can be returned to the solutions by allowing ∆± → ∆±±∆B. These solutions are not particularly enlightening other than it is nice that there is a analytic solution and we can use the two-level atom steady-state solution to check their validity. To do this, remove one electric field by letting Ω∓ → 0 and ∆∓ → 0, then the population in the |±〉 state is ss |Ω̃±|2ρ±,± = (2.61) 1 + 4δ2± + 2|Ω̃±|2 28 where, as defined in Appendix A, Ω̃± = Ω±/Γ and δ± = ∆±/Γ. These agree with the population of the excited state from the two-level atom optical Bloch equations 2.27, as they should: as was the case for the |0〉 state, when there is no field coupling an excited state to the ground state, that excited state can be ignored as its population will decay to zero. Thus, with only one laser, the atom should behave as a two-level atom excited by just one field and the solutions should agree exactly with the optical bloch equations. The quantity Ω̃ = Ω/Γ is often written in terms of the ratio of the electric field intensity relative to a saturation intensity as I ∣∣ ∣∣ ∣ 2 2 = 2 Ω̃∣ 2 |Ω|= . (2.62) I 2sat Γ This is the saturation parameter, Equation 2.38, with ∆ = 0. For the electric field with amplitude E0, the intensity of the field is I = 0cE 2 0/2. Together with the definition of Ω in Equation 2.18, the saturation intensity is  c~2Γ20 Isat = ∣∣ ∣2 . (2.63) 4 ∣〈g|̂ · ∣d~|e〉∣ This is referred to as the saturation intensity, as when I  Isat (i.e. |Ω| >> Γ), the two-level atom excited state population in Equation 2.27 saturations to 1/2. This quantity clearly has different values for different electric field polarizations ̂ (the dominator will be changed). When the saturation intensity is referred to throughout this text, it is in reference to circular polarized light exciting MOT trapping transition which has a value 1.669mW/cm2 for the atomic species of rubidium used in our experiments [62]. 29 2.4.1 Optical Molasses The idea of radiation pressure can be extended further to look at the effects of identical counter-propagating plane waves interacting with an atom. These two fields are defined as E+(~r) = E −ikz 0e and (2.64) E−(~r) = E e +ikz 0 . These are defined so the field propagating in the negative z-direction excites only the σ+ transition and the field propagating in the positive z-direction excites only the σ− transition, as shown in 2.3. For each of these beams, atoms moving against their propagation direction should see the frequency of the light shifted to a higher frequency due to the Doppler effect. An atom with velocity v then sees each beam having velocity ω+ = ω + kv (2.65) ω− = ω − kv, where ω is the rest frequency of the light. The atom then sees light that is detuned from the resonant frequency ω0 by ∆+ = ω+ − ω0 = (ω − ω0) + kv = ∆L + kv (2.66) ∆− = ω− − ω0 = (ω − ω0)− kv = ∆L − kv, where ∆L is the (red) detuning of the laser from the atomic resonance. Following the atomic force derivation in Section 2.2.3, the force operator is ( ) ~ −ik~ ∗ − † − ik~ ( ) F = Ω σ̂ ∗ †+ + Ω+σ̂+ ẑ −Ω2 2 −σ̂− + Ω−σ̂− ẑ, (2.67) 30 which calculated in steady-state gives F ss(v) = [−~kΓρ(v)+.+ + ~kΓρ(v)−,−] ẑ, (2.68) where the density operator elements are written to note that they explicitly depend on the velocity of the atom through the Doppler shifted detuning of the lasers. Allowing √ Ω+ = Ω− = Γ/ 2 and ∆L = −Γ, the force (in units of ~kΓ) is plotted as a function of velocity (in units of |∆L| /k) in Figure 2.4a in red. This result is easy to explain. An atom moving towards one of the fields, sees a Doppler-shifted beam (at higher frequency) that is closer to resonance than the beam it is travel along with. A smaller detuning (recall, ∆L < 0) allows the opposing beam to more easily excite the atom. This is shown with the different length arrows coupling the ground and excited states in the level diagram of Figure 2.3. The photons absorbed from the in-tune beam apply a larger force than the opposing beam, giving a net force pushing the away in the opposite direction of its motion. With counter- propagating beams, each beam slows atoms moving towards it. This gives an overall decrease in speed, and thus temperature of the atom. This arrangement of lasers is known as optical molasses and are a well studied method to cool atoms [1–3]. For the D2 transition of 87Rb, the force scale is 3.2×10−20 N and the velocity scale is 4.7 m/s in Figure 2.4a. A speed of 4.7 m/s corresponds to an atomic temperature of 78mK [62]. Typically MOT temperatures are order of 10 µK, corresponding to velocities of order 10 cm/s (This difference in expected and measured temperatures is discussed in Section 2.5.3). The graph of Figure 2.4b rescales the axes to 10−22 N and cm/s. Clearly from this graph, in the range of velocities for atoms in the MOT, the force is very close to linear. Expanding Equation 2.68 to first order in v gives a 31 + |i |+ i |g i FIGURE 2.3. Optical molasses beam arrangement. An atom moving towards a beam see its frequency shifted closer to resonance. F [~k] (a) 0.2 v [|L| /k] 0.1 " or # ~ |L| z µ g @BB F @z -3 -2 -1 1 2 3 -0.1 -0.2 (b) ⇥ (c) ⇤ ⇥ ⇤F 10 22N 30 F 10 22N 5 20 10 v [cm/s] z [µm] -40 -20 20 40 -40 -20 20 40 -10 -5 -20 -30 FIGURE 2.4. V-atom forces. (a) The red curve shows forces of a 1-dimensional V-atom and the blue curve shows the force on the extended two-level atom. Graph shown is for both atom in optical molasses as a function of velocity or atom in a MOT as a function of position. Peaks occur where Doppler shift (molasses) or Zeeman shift (MOT) match the laser detuning, ∆L. (b) Damping force for velocities in range of atoms in a MOT. (c) Restoring force for positions on scale of atomic motion in a MOT. 32 damping force F~ (v) = F~0ẑ − βvẑ where ( ) ~kΓ |Ω̃ 2p| − |Ω̃m|2 F0 = (2.69) 1 + 4δ2L + 2|Ω̃ |2m + 2|Ω̃ 2p| and { 2 2 2 ~ 2 | | × (|Ω̃m| +|Ω̃ 2 p| )(|Ω̃ 2 2m| −[|Ω̃p| ) +4(|Ω̃ 2m| +|Ω̃ 2p| )β = 8 k δL ]+ [1+4δ2 2 2 2 2 2 2 L+2|Ω̃m| +2|Ω̃p| ] 16δL+(2+|Ω̃m| +|Ω̃p| )} (2.70) 4(|Ω̃ |2+|Ω̃ |2)(1+[ 4δ2 )+16|Ω̃ |2m p L m |Ω̃p|2 ] 2 . [1+4δ2L+2|Ω̃m|2+2|Ω̃p|2] 16δ2L+(2+|Ω̃ |2m +|Ω̃p|2) The sign of this constant force F0 depends just on the two field’s intensities (recall, the Rabi frequencies are proportional to the field strength). So, this constant force is just an overall force caused by one beam having more power, and thus pushing harder, on the atom. For most cases, this is dropped by balancing the beams, as plotted in Figure 2.4a. In this case, the damping force reduces to F~ = −βV-atomvẑ with ∣∣ ∣∣2 16k2~ |δL| ∣Ω̃∣ βV-atom = −[ ∣ ∣ ] [ ] . (2.71)∣∣ ∣∣ 2 2 2 ∣ ∣ ∣2 ∣ ∣∣ ∣ ∣ ∣ ∣ ∣ 4 1 + 4 Ω̃ + 4δL 1 + 4δL + 2 Ω̃ + Ω̃∣ A common alternative derivation of this damping is done by using the two-level atom result for ρsse,e in Equation 2.27 for both excited state populations in Equation 2.68. This is taking the two equations of 2.61 and treating them both as accruate, while they are in only true in the presence of just one excitation field. With these, and making sure to use the the appropriate detunings from Equation 2.66, as done in [25, 48], the atomic damping force is [ ] F = −~kΓρSS (∆ + kv) + ~kΓρSSe,e L e,e (∆L − kv) ẑ. 33 The results of this extended two-level atom is shown in Figure 2.4 as the blue curves. This calculation does not rely on oppositely polarized fields. Two fields with linear polarization will work for the extended two-level model. The extended two-level method overestimates the force from one beam. It assumes too large of an excited state population, as it ignores effects on the atomic population due the opposite beam. With the V-atom, an individual excited state population is reduced from its two-level atom population, as some of that population is shifted into the other excited state. At large speeds, the agreement between methods is better. In this case, the fast-moving atoms are much closer to resonance with one of the beams, allowing it to dominate the atomic state. The V-atom then behaves much like the two-level atom at large speeds. In the small speed range, the force calculated by [25, p. 88] for the extended two-level atom is ∣ ∣2 16~k2 | ∣ ∣δL| ∣Ω̃∣ F~ (v) = −[ ∣ ] vẑ. (2.72) ∣∣ ∣∣2 2 1 + 2 Ω̃∣ + 4δ2L Compared to the V-atom in the Equation 2.71, the damping coefficient β for the extended two-level atom is slightly larger. While the damping of the motion should drive the atom to rest, this damping force is balanced by the random emission of photons by the atom. Each photon absorption, which damps the motion is followed by an emission, which gives the atom a momentum bump of magnitude ~k in the opposite direction of the photon direction. With a detuning ∆ = −Γ/2, to maximum β for both models in equations 2.71 and 2.72, analyzing the diffusion of the atomic velocity distribution [2, 25] results in a minimum atomic energy of ~Γ Umin = . (2.73) 4 34 With the 1D equipartion theorem, this results in what is called the Doppler temperature, the lower limit on temperatures for atoms in an optical molasses. ~Γ TD = . (2.74) 2kB 2.5 Magneto-Optical Traps As noted above, atoms in optical molasses are still free to diffuse [68] without being confined to any one location. Confining the atom to make an actual trap can be done with magnetic fields. 2.5.1 Magnetic Confinement Placing an atom into a region with a spatially variying magnetic field will give the excited states of the atoms a spatially varying Zeeman shift. In particular, adding a linear magnetic field with gradient −B′z (here, assume B′z > 0) shifts the excited state energies (for the V-atom) by µBgF ∆ = (−B′B z) , (2.75)~ z following the convention from Equation 2.59. The spatial energy shifts of the |±〉 states are shown in Figure 2.5. From this Figure, an atom located at z > 0 has its |+〉 excited state energy shifted down in energy. This reduces the detuning of the laser from resonance, ∆L (shown in Figure by the thin, red line), improving the atomic interaction with σ+ light. With σ+ light traveling form the z > 0 direction, photons absorbed from that field push the atom back toward z = 0. The same analysis holds true for an atom located at z < 0 with the |−〉 level and the σ− field. This results 35 z = 0 |i L |+ i + |g i FIGURE 2.5. Magneto-Optical Trap Level Diagram. An atom displaced from z = 0 has one excited state energy Zeeman shifted closer to frequency of the red detuned laser. Correctly matching the polarization of a laser to the direction of the magnetic field applies a restoring force on the atom towards z = 0. in the magnetic field imposing a restoring force from the lasers onto the atom. This behavior is the position-space analog to damping in optical molasses. This interaction is quantified following similar steps to those in Section 2.4.1. The detuning of the atoms from resonance follow the equations ∆+ = ω+ − ω0 = ω − (ω0 −∆B) = ∆L + ∆B (2.76) ∆− = ω− − ω0 = ω − (ω0 + ∆B) = ∆L −∆B, where it is important to note that the energy differences here are the result in shifts of the excited state frequencies, rather than shifts in the laser frequencies as was the case for optical molasses. The force on an atom is then F ss(z) = [−~kΓρ(z)+,+ + ~kΓρ(z)−,−] ẑ, (2.77) 36 where the explicit dependence on position comes from the detuning energy shifts in Equation 2.76. This equation has exactly the form of Equation 2.68 in position-space rather than velocity space. The graph of this function will be the same as the red curve in the graph of Figure 2.4a with the horizontal axis being position in units of ~∆L/µ ′BgFBz. This graph agrees with the restoring force interpretation presented above. A change of sign for the magnetic field (B~ (z) = +B′zzẑ) requires flipping the polarization of the two beams to create a trap again. Typical MOT parameters, for our single-atom MOT, have B′z ≈ 241 G/cm. Following the gF derivation in [62] to use gF ≈ 1.33, the distance scale is 135µm. Atoms in our MOT have typical displacements from the center of the MOT on the order of 10 µm, again allowing examination of small position displacements (“small” being defined so atomic Zeeman shifts are much smaller than the detuning of the laser). A graph of this is shown Figure 2.4c. Again, this force is very close to linear so that we can expand the equation to get a restoring force F~ = F~0ẑ − κzẑ with the same value for F0 as Equation 2.69 and { 2 2 ′ | | × (|Ω̃ | 2 m +|Ω̃ 2 2 2p| )(|Ω̃m| −[|Ω̃p| ) +4(|Ω̃m|2+|Ω̃p|2)κ = 8kµBgFBz δL ]2 +[1+4δ2 +2|Ω̃ |2+2|Ω̃ |2] 16δ2L m p L+(2+|Ω̃ |2m +|Ω̃ 2p| )} (2.78) 4(|Ω̃ |2m +|Ω̃p|2)(1+[ 4δ2L)+16|Ω̃ 2m| |Ω̃p|2 ] 2 . [1+4δ2 +2|Ω̃ |2+2|Ω̃ |2] 16δ2 +(2+|Ω̃ |2+|Ω̃ |2L m p L m p ) Returning briefly to the purely magnetic trapping of Section 2.3.1, the magnetic trapping force of Equation 2.53 for the V-atom was F~ ss ′Mag. Trap = −µBgFBz [ρ(z)+,+ − ρ(z)−,−] ẑ. (2.79) This force is identical in form to Equation 2.77 but with scaling µ g B′B F z. For the typically MOT parameters above, the magnetic trapping force is of order 10−23. This 37 is orders of magnitude smaller than the MOT force scaling ~kΓ, which is of order 10−20. As seen with the |0〉 energy level, without electric fields to excited the atom, those excited states will not be populated. Without these fields, then, the V-atom will have no magnetic trapping as Fg = 0. For the full 87Rb atom, however, the ground states do have angular momentum. Thus, there can still be magnetic trapping for 87Rb without near-resonant electric fields. In addition, the |Fg = 1,mg〉 states have gF = −1/2, which results in atoms in this ground state being repelled from the minimum of the magnetic field magnitude rather than trapped [63, 69]. For the magnetic confinement equation, in the case where Ω− 6= Ω+, the F0 term is non-zero and shifts the “center” of the MOT – the position where the restoring force is zero. A similar effect occurs with a background magnetic field B~0 = B0ẑ to give a total field of B~ (z) = −B′zzẑ +B0ẑ. (2.80) This shifts the location where B~ = 0, again moving the center of the MOT and modifying the force equation ( ) ~ ~ − B0F = Fz,0 κ z + B′ ẑ. (2.81) z From Equation 2.78, κ can be written as κ = κ̃B′z. Then the force equation becomes F~ = F~z,0 − κzẑ −B0κ̃ẑ. (2.82) 38 Therefore, a non-zero background field of magnitude B0 = Fz,0/κ̃ can cancel the offsetting force due to beam imbalance to return a purely linear restoring force F~ = −κzẑ. This can also be used to cancel gravitational forces on the MOT. Comparing the equation for restoring constant κ to Equation 2.70, the optical molasses damping constant, we have µ ′BgFB κ = β z , (2.83) ~k which is the same as the equation for the extended two-level atom [25]. Therefore, the extended two-level atom solution has the same overly strong assumptions for the restoring force strength as it does for the damping force. Taking the case with balanced electric fields, the restoring force becomes F = −κV-atomz with ∣∣ ∣∣2 16kµ ′BgFBz |δL| ∣Ω̃∣ κV-atom = [ ∣ ∣ ] [ ∣ ∣ ] . (2.84)∣ ∣2 ∣ ∣2∣ ∣ 2 2 ∣ ∣ ∣ ∣ ∣∣4 1 + 4 Ω̃ + 4δL 1 + 4δL + 2 Ω̃ + ∣Ω̃∣ 2.5.2 Finally, a MOT Combing the effects of the atomic motion-based Doppler shift creating a damping force and the magnetic field creating a spatially dependent restoring force, the force on an atom in the MOT is that of a damped, harmonic oscillator F = −βz − κv. (2.85) Here, the F0 forces from MOT beam imbalance have been suppressed. 39 With such a force, an atom in a MOT should behave as in a harmonic potential, U = 1κz2. As above, the damping of the atomic motion by the MOT lasers will not 2 force the atom to rest because of the random emission of photons. Thus the atom should have an average energy due to motion that has the form 〈U〉 = 1κ〈z2〉. With 2 the equipartion theorem [25], in one dimension the atomic position should follow 〈z2〉 kBT= . (2.86) κ With this relation, measurements of the size of a magneto-optical trap can give a measure of the trap’s temperature as done in many of the temperature techniques discussed in Chapter I. Early MOT experiments expected temperatures close to the Doppler temperature, Equation 2.74, but experiments measured clearly lower temperatures [30, 70]. 2.5.3 Sub-Doppler Cooling Atomic temperatures below the Doppler temperature are a result of polarization changes seen by the atom moving in an optical molasses [64, 71]. In both cases, this enhanced cooling only appears for atoms with multiple ground states, such as the one shown in Figure 2.6, so we’re getting a bit ahead of ourselves for the discussion in Chapter V. For optical molasses with counter-propagating linearly polarized electric fields, alluded to in the discussion of the extended two-level atom in Section 2.4.1, the sub-Doppler cooling mechanism is named the Sisyphus effect. In this linear field arrangement, the polarizations for the two counter-propagating fields created by MOT beams are are right angles. Their interference creates two potentials, which underpin the force of Equation 2.29, that the atoms move through. 40 FIGURE 2.6. Atom with Mutliple Ground States One potential oscillates as sin(kx) where k is the wavenumber of the light and corresponds to the σ+ polarized light. The other potential oscillates as sin(kx + π) and corresponds to the σ− polarized light. Atoms primarily interacting with the σ+ light, such as atoms oscillating between the outer two levels on the right side of Figure 2.6, will follow that light’s potential energy curve, increasing and decreasing speed as it moves. This is shown as the leftmost atom (black dot) in Figure 2.7. However, on occasion, when the atom is at the peak of the potential, with the smallest kinetic energy, the atom can absorb a σ− photon (center atom in Figure 2.7). When it does so, it moves onto the σ− curve, which is at its lowest point in the potential energy curve (show as the arrow for the center atom). The kinetic energy of the atom here does not change (except for a small change due to photon emission recoil), but it now is at a potential energy minimum (rightmost atom in the Figure). Repeating this process lowers the overall mechanical energy of the atom and results in temperatures lower than predicted by Doppler cooling. For optical molasses with circular polarizations, such as the MOT described in 2.5, the sub-Doppler cooling mechanism arrises from an additional enhanced 41 ">=ohAd"gLA_t1xathlsAPxYpdw0J96uXnjDc46hiC+BTvNPuvvecU4yNJ0PDhmhJzQC0ab2kl4A0fAGr42dFdHr1ZiZc4KSw3skqv26VU3MFxMgSLDVdci37BAAVHxPpbdMVAY6wm0KJp946HubXHn2jiD6cN4m6Ah0iQCZ+2BiTsv3NDPAuwvjv+evc0UQ44y2NNJD0JP0Dkh0mrhFJ1zcQwCq0VaFbS2dk7lP4VAJ0XfcAiGTru4c2NdDFJd0rk10ZriFZ1cJ4PKhShwz3Csak2qlvA2f6GV4Ud3dMrFZxZM4gSS3LkDvV6dUcMix3g7LBVAcA3PBxApxddVYY0w90uJn9D64uhXCnBjvDPcv4e6UhyiJCP+hBhTzvCNaP2ulvAvfeGc4Ud4dyr4M3AABi7s32iKcVdqVwDcLUS6gvMkx3FSZ4 eJjY7uCfmvqj9QKd8EZJJE9MJRtUKw0IRoQ2mxci0gcl4ZKakw9/7OVv1b7sCZk+ar+ycuqezqcuy/uorQas+okEZ9+Vsk7yCc1JvZbjOf7SVc9CwG/UakKRk94xleZ8gK0dcqcmxvie2EQjmaRnIQo4wgKa0Yt0RTU9MY9cJcJ6JZEjzEVScajnCsk4QgoGUYa0kyo9TYcR9cc6x/ZJjeacjVESeuvfm9qjdQsnCQk4GgoaUYy0kTo9ccRc/9xZ6jeJKZ9JJJEE8ZKQdjq9mfvuzqcZytuQqZzccoyru4r0ai+RkKZU++ss7lCg1cvxb2Om7IVw90wR/MaaKkk+eEv1bMOI7SV09Jwm/waRKJkE4QlRZogK0tcUc9xJiE28VjcasnCQk4GgoaUYy0kTo9RYcc9c/6xZ 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 XqzBXREWfvVz=OWqsE4qc6g2hl+wX3zjg44>CtDiAxBeLtEaBg4AE8gWfrznq0XqzXRWvOWs4cgh+w3j4BL52QEA==RXvx3jt>itEx4WBESR6g4LA 0 have variances a bit larger than the mean, justifying treating the count rates as Gaussian rather than strictly Poissonian. In practice, following the suggestion of Section 4.3, σn is treated as nearly Poissonian in the form σ2n = 〈nr + B〉 (1 + ζ0 + nζn), where ζ0 is a small non-Poissonian contribution to the background and ζn is a increase in variance due to atomic motion. Typically used values for the non-Poissonian contributions are ζ0 = 0.08 and ζn = 0.2. Motion of the atom changes coupling of atomic fluorescence into the APD, noise in laser power, magnetic fields, and polarization all change the photon emission rate. Thus, r is not constant, but is assumed to also have a gaussian probability distribution P (r) [ ] 1 −(r −R)2 Fl1-at(r) = √ exp , (4.4) 2πσ2 2σ2R R where R is the average emission rate. 102 Combining these two probabilities gives the probability distribution for n atoms if we measure a photon rate of x with an average background rate, B, and average single-atom photon rate, R, ∫ −∞ Fln(x) = Fln(x, r)Fl1-at(r)dr −∞ [ ] √ 1 −[x− (nR +B)] 2 = exp . (4.5) 2π (n2σ2 + σ2) 2 (n2σ2 2r n r + σn) A few plots of the fluorescence probability from n atoms in the MOT are plotted in Figure 4.5 with a single FPGA measured data run. Each peak along the y-axis is the distribution, Fln(x), for atom numbers n = 0 to n = 5. The plotted probabilities are the final distributions calculated from the last measurement of the data run. The values for R, B, σR, and σn that make these distributions are calculated following the method in Section 4.2.6 from the data shown in the figure. 4.2.3 Noisy Measurements The measurements take are noisy fluorescence measurements from the experiment. This noise is not fluctuations in the background signal or in the fluorescence rate from the atom, those are built into the fluorescence distributions Fln(x). Instead, this noise is from measurement errors—dark counts from the APD and possible counting mistakes at the FPGA. Making a measurement, y, produces a sampled value x from the distribution P in(x) plus some noise, written as ζ (note that a noiseless measurement would then sample x values directly from Fln(x)). So the measurement outcome has a value y = x+ ζ. (4.6) 103 4500 Probability Distributions data 4000 3500 3000 2500 2000 1500 1000 0 200 400 600 800 1000 FIGURE 4.5. Sample data for bayesian fluorescence number estimation. Vertical axis is in units of photons/100 ms and horizontal axis is in units of recorded FPGA packets. Note the increasing distribution width for larger values of n. This comes from the factor of n2 in the variance of Equation 4.5. This overlapping fluorescence distributions for n and n + 1 atoms shows the limitations of photon counting to estimate atom number at a large number of atoms. 104 To treat this noise, it is assumed that values for the noise follows a Gaussian distribution [ ] 1 −ζ2 p(ζ) = √ exp , (4.7) 2 2πσ2 2σζζ so that the probability to have had a noise value of ζ = y − x is [ ] − √ 1 −(y − x) 2 p(ζ = y x) = exp . (4.8) 2 2πσ2 2σζζ The probability for the noise to take this value is the same as the probability to have measured y when the sampled fluorescence value is x. For the Bayesian evolution, the likelihood to measure y with n atoms in the MOT is needed. This is the the probability for the noisy measurement to have sampled a value x and averaged over the probability for the fluorescence to have value x with n atoms in the MOT as below. ∫ P (yi|n) = p(ζ = yi − x)Fln(x)dx [ ] √ 1 − [yi − (nR +B)] 2 = ( ) exp ( ) (4.9)2 2 2 2 2π n2σ2 + σ2 + σ2 2 n σR + σn + σR n ζζ This is the desired likelihood function for the bayesian estimation of the number of atoms in the MOT. It tells the probability that when a fluorescence rate yi is measured, there are n atoms in the MOT, given that there is 1. an average photon emission rate per atom of R and with standard deviation of σR, 2. average fluorescence background B, 105 3. background (plus atomic motion) fluorescence rates with standard deviation σn, which are slightly super-Poissonian, and 4. systematic noise with variance σζ . 4.2.4 Keeping P (n) > 0 From Bayes’ rule, if P in(x) = 0, for any time ti, then for all later times, the probability to have n atoms in the MOT will always be zero. This can be resolved numerically one of two ways. First, manually setting Pi(n) =  for some fixed small value  at each time step if Pi(n) < . Second, using a loading-rate method. This method includes, in the probability evolution, atomic loading and loss terms for the MOT. With the loading rate method, the probability to have n atoms in the MOT evolves as dP i(n) = −nΓP i(n) + (n+ 1)ΓP i(n+ 1)− LP i(n) + LP i(n− 1), (4.10) dt where Γ is the rate that an atom is lost from MOT, and L is the loading rate of atoms into the MOT. The first term represents any one of the n atoms leaving the MOT. The 2nd term represents any one of the atom leaving a MOT that used to have n+ 1 atoms. The 3rd term represents at atom loading into the MOT from the background gas (to create a MOT with n+1 atoms). The 4th term represents an additional atom loading from a MOT with n − 1 atoms. This loading-rate equation is identical to loading-rate model analysis done for small numbers of atoms in a MOT [20]. Using this method numerically requires a maximum number of atoms, Nmax to be set. Doing this adds a cut-off term Θ(Nmax−n) to the 3rd term so that an addition 106 atom “cannot” load if n = Nmax. This definition also allows n=∑Nmax dP (n) = 0, dt n=0 so that the loading-rate method appropriately conserves probability. Tests with both methods give the same predictions for N(t) under a variety of other parameters. In practice, we use the loading-rate method with typical parameters L = 0.006, Γ = 0.003, and Nmax = 8. These values are measured atomic loading and loss rates from the experiment. 4.2.5 Number Estimation Combining the Bayesian evolution with the loading-rate differential gives an overall evolution for the atom-number probability [ ] dP i+1(n) = −nΓP i(n) + (n+ 1)ΓP i(n+ 1)− LP i(n) + LP i(n− 1) dt+  P (yi|n) ∑ − 1P i(n). (4.11) P (yi|n) n Again, this equation conserves probability when summed over n. It also solves the issues of any P (n) → 0 as the probability for there to be n atoms in the MOT will be increased slightly by the loading rate portion (first term) of the differential. This probability evolution could be interpreted as a deterministic “Hamiltonian” like evolution at all times, punctuated with the noisy measurements at times ti, similar to the stochastic evolution of a system in quantum measurement theory [123–125]. 107 This algorithm is simple to implement in real time while we record fluorescence from the MOT and estimate the number of atoms in it. To determine the number of atoms in, we typically assume the n with the largest probability is the correct number of atoms in the MOT. This generally works well as often the average single- atom fluorescence rate is much larger than the width of its fluorescence distribution, R σR. In some cases, such as poor alignment of the APD lens system with the MOT center or cases where atomic position distributions are large, just taking the largest probability as the number of atoms in the MOT causes problems. The main error seen is constant fluctuations in the estimate of the number of atoms in the MOT as two values for Pi(n) are close to 0.5 (typically for n = 0 and n = 1). In such cases, assume that states with n > 1 remain essentially unpopulated and initially P i(0) = 0.51, so that there are believed to be no atoms in the MOT. Updating the probabilities after measurement could give P i+1(1) = 0.51 so there is now believed to be one atom in the MOT. Another update gives P i+2(0) = 0.51, so the state again returns to there being zero atoms in the MOT. This can repeat often if measurements of the fluorescence tend to stay in the “middle” between the peaks of the likelihood functions for zero atoms or one atom. This most often happens when the initial assumption about the single-atom fluorescence rate, R, is larger than the actual rate in the experiment. Other than using a more realistic single-atom fluorescence rate, this can be solved via more complicated assumptions about when the number of atoms in the MOT changes. One method is to require the largest probability at ti+1 to be above some threshold value (larger than 0.5) before determining the atom number changed from time ti. A second method could require the largest probability be above the 2nd largest probability by some determined factor, limiting jumps 108 between two states with probabilities close to 0.5. A third method would require that the maximum probability remain the maximum over a number of fluorescence measurements, avoiding the possibility of fast fluctuations in number. A final method is to reduce the possibility of such oscillations by updating the background and single- atom fluorescence rate as data were being recorded, described in detail below. 4.2.6 Background and single-atom fluorescence estimation Mean values for single-atom fluorescence rates are often 5 or 6 times larger than the standard deviation of the background signal (this fact is occasionally used after data is recorded to locate times when a given number of atoms are in the MOT, as noted in the atom counting method in Section 4.1.2). Because of this, in many cases P i(n) is very close to unity. For example, for the data shown in Figure 4.1c, at times when n=0 gives the largest probability, the mean value of P (0) is 0.9986± 2.2× 10−4, and at times when n=1 is the largest probability, the mean value of P (1) is 0.9979 ± 1.4 × 10−2. With good alignment of the imaging system with the MOT center, these are not uncommon values. With P (n) ≈ 1 for some n, we can leave n fixed in Equation 4.5 and use the fluorescence measurements to update values for R or B rather than update predictions for n. Taking n = 0, Equation 4.5 gives [ ] √ 1 − (x−B) 2 Fl0(x) = exp . 2πσ2 2σ20 n Because the noise in a measurement is assumed to be Gaussian, updated values for the background mean value and variance after a measurement can be written analytically 109 [122]. Following this, a measurement of yi allows updating of B and σ 2 0 as B σ2i ζ + y 2 iσ0,i Bi+1 = σ20,i + σ 2 ζ (4.12) σ2 2 2 0,i σζ σ0,i+1 = ,σ2 20,i + σζ where, again, σζ is systematic noise in the measurement. These updates are easy to make and include in subsequent measurements and probability calculations. When there is an estimate of n > 0 atoms in the MOT, instead of updating the background and its variance, the atomic fluorescence rate and its variance is updated using the n-atom fluorescence probability Equation 4.5. The background level and its variance complicate the calculation for updating R and σR. This calculation is done in Appendix D and the conclusions are R 2iσζ + (y 2 i −B)nσR,i Ri+1 = n2σ2 2R,i + σζ 2 2 (D.3)σ σ 2 Ri ζσR,i+1 = .n2σ2 2R,i + σζ These two updating methods are effectively just a Bayesian filter [122] for the background (when n = 0) and for the fluorescence rate (when n > 0). 4.2.7 Bayesian Algorithm The full Bayesian algorithm is sketched schematically in Figure 4.6. At each time step, ti, a fluorescence rate, yi, is measured. From this, calculate the likelihood functions for each atom number according to Equation 4.9. These are then used to update the atom-number probability distribution with Equation 4.2. The number of 110 Record Fluorescence Rate Calculate likelihood functions,P (yi|n) Evolve atom number distributions with Bayesian evolution P i+1(n) = P i(n) + dP (n) 6s=>PLKAO7+cBfOEiN4CMLnpJu5UjiI1"=AAAHAVisbLD1SRNAFC2br3pYUpcj3IS"UAmARATiCbUDLSPNBFR2mrRphU ci3eSaUAmdRMTFCnUDLJPOB5R7mR<4clitjxOtAsBac_+azeO4i"K33iPz"c=A4j6ceRsFaDbO_71Na=h>sA CtHicxVeLtsaAlLAAACAHicbVDLSsNAFL2pr1pfURcu3ASLUCmURATdCEU3LiPYB7RpmAAACAHicbVDLSsNAFL2pr1pfURcu3ASLUCmU EtOhAxkTJgOKquzH3HuylRSJT8yYNT2umtZ77rrGkadHd7azC/qnUVpv42RbquWSFDQ0RONCFfMmAq/jkECRuufn+2jlT6ZQDqgHnGGqiSa5YkEItPaRKGOawECUWGNf4+Z7Obv1S7sxZk+aruyC34v2qclZ84x5m5ecGu7mdQrham2x9rzh/ZrEDFuV8yKBhMxMJJDd7myUV5fc7Fu77rrGkadHd7azOHuJT5KAhVgkCh/nUVpv42RbquWSFDQ0RONCFfMmAq/jkECRuufn+2jlT6ZQDqgHnGGqiSa5YkEItPaRKGOawECUWGNf4+Z7Obv1S7sxZk+aruyC34v2qyUmA7tddDTJqJNMnxuMRhUB7Klyg8DV+fkcFuFF4DCEkrmZT/3hkzurg9Zxj2fmCa/hMrNQQdWmR7puqGacder5Zm25yxS4u8qZOlVcKqH2Hvq4Q36Cly2unruaR+EkjZqxmsf7CSO10vDbSOu7bZ2+v4VfnN/GzW7UHCaEGwraZOtGuKTRYa8PJtRIyEHkzYx5EaASJi5qhGTGOcmM84Z5mxec5u7GdQmhar2xmrz9/ZhEDrucFV8fKByMxhJJMd7DyUmAq/jkECRuufn+2jlT6ZQDqgHnGGqiSa5YkEtmu2TNYy8TJSItPaRKGOawECUWGNf4+Z7Obv1S7sxZk+aruyC34v2qclZ84x5m5ecGu7mdQrham2x9rzlHOuTKhg5VJkAOEtxqz3HuylRSJT8yYNT2umtZ77EOtyxHqRzu3lrGkadHd7azC/qnUVpv42RbquWSFDQ0RONCFrZrED/qcfV8yKBhMxMJJDd7myUf 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 w=w=w==titxwla/ Estimate number of atoms in MOT, Ni+1 Ni+1 = 0 Ni+1 > 0 Ni+1 Ntrig Ni+1 < Ntrig Trigger Experiment Update Update Bi+1 & 2 i+1 2B,i+1 R & R,i+1 Evolve atom number distributions with deterministic evolution i+1 P i+1 i+1 dP (n) (n) = P (n) + dt dt teYxnigtX Fs=h=AAAAAA"=>urI3YFXZkrgfeYuewn=Y4/MZrl/"Z=l VWapN+vtAZysX8bOcn/z4gHW/tBl3CNT3ov4ZllowgbqtP3JQeJIwkzktM0TbCGZOAaXcrVi5mLRoiCJhiUCaRV62uR9nmXFYIsANDSLcZbFi303ZWVa3pINq+TvottAlZoylslXP8ebJONc3nv/Zz44wgbHgW3/QtJBwm9Row4bqlP3gQetIwJzkJM0kbCtTZOGaXumXnYIFANsSLDZbci3FCArVc5liRoLCJiiUhaVWapN+vVWRabpsNR+mv3tFAiZVyIsZXF8SbXOCchn6/9zn4AgLHcWC/itDBNlYNuT23AoUvazCk2kRMu0XtYbFCNTSZDObGiaFXCA3rcVZcL5smAiIRnomL9CVJ6iRtAXysZ8bXcnOz4/HWgtB/NTlov3Zl3ow4bqlP3V2m6LCrRRachCUoiii5JVgkGaOCZtTbMQtJkzwJIe doWxw6l5oKicoA8YkEbQgEtaiwa2/k/tdtZZ1bcbX1gsbx1X2bg1cs7xt2d7112ca+cqEJgE+oYE89q2KOocJZxAohE2l4HxoHcaosEXr/qr9cYFYv2Hix46Oa1LAjcCtZq9ZxT/S5AFrcLihPVGca2gkQ49ldxaHHcajXsbro/crQYqFHvEiYxa6o14L9cgQaV6LPGir5SFc5Tlqx/9tcACZbX1gs1xcbt721dZ2+cqa/EEtogEY9dW9kQcg6aVGLP5ircSFl5WkT/xc9tZACAj1L4ao6ixYHEvYFqcQrr/oXbscaHHaxl4o2whxAoZicKO82k9YEgotEEJaq/+2cd1dWj9xk5QZcFg/69aCVSGlLTPq5citrAcAZ2ctbsx1Xg7qk28OKciZoAAJxhw2o4lxaHHb1acsbXo/rrQ1LY4YaEoq6HivxFc hUsafHVCTkCH5WVi3FzBwIVw6ehHDH6EJFDuupL5anx7YmxlrjGX2TKoC8sAGhxS0aSDTgrU+PSEWe9cQ6oVQJjeqQ6V2o4c4QGQ6cqoKVJQ0eT9xPPWRwhSUUz++DsrDzHTsaVS5hf0T3ixUAFGVoCaJbsGorImVj6V6IEFgUSi8TXfl57VpsEHeCBahsnVb3SzAd+QqvlOOlvqQ+AzS3bVnshaBCeHEspV75lfXT8vwQ4c6oHVqQJek90PxWHwPSROQlFqd+3AuSVbanWhCB2eDEGp67KlJXT8DSCgaEV6A6xhxDn63JYD0aCbhJxrSm5xanaYTCuxz5LaruWLDWuu+FFRUPRHSxP0wkHJWqxHP6z4rwLauTSa5Cxh30YxnxGAm0JrwoVsbDaCKTJD6Gj2h66VFIESgU9ikbJsdOlvqQ+dAzS3bVns nl5YEd0zsbKdJnppCdGpFVZHfkYcVBlHwaAIbjHVoEVY5OfOBkscK1ndfInPqqEq1I4EIIVeDknTlZqKCEEswE1+5m4mAGIAG1V6bTDPeIqcdIOoVcHdHQpYzgnl5hBFKvHqFPZZ0JKCsoZ0E+mEJQmKAmfw1nHsoDQ1gIhZvBP5JmoE+KQ4mSnmDFIAB0mJKRSIFz0mRFzGFssAOO7Vm78fNm0bk8I1INqDI01TkkOeEIjEaIBqkPVIddn1bcdklOoOEY+EGV6jIIoaQHgBhcvkPHJVop+dQpmnndDbIzBdmYKlSnFo0VREz5F+sfOG7Bm68sNI0q+keIEIqqPId1ckOOYEVjIaHBckHVpdpndbzdYlnoVE5GBfsI6onKHgQhFlqPvJCZ0+oQKEwnmD1sZBImE54SKFAmJR0zmIGsFOVAfm781bD0NT5A5/HFfWkW9iDGg658ux/4GlUb9nBfPx=lklP9BW/GU/GifFubExxl4g85kD95fHfA5lxnitBP6x9GtU/Gi/FufExbl4x85gD9kfH5A5fxnlk=6PBnWk95GAU FIGURE 4.6. Bayesian algorithm flow chart. Initial background fluorescence measurement and algorithm ending elements are not shown. 111 atoms in the MOT is updated according to any of the methods discussed in Section 4.2.5. If the algorithm is set to trigger an experiment based on atom number, this number is checked against the triggering atom number, Ntrig. If the number of atoms meets or surpasses that value, whatever experiment is to be done is triggered. It is also possible to trigger an experiment manually after a given time. The experiment is a predetermined order of commands and runs “in the background” while the Bayesian algorithm continues. Based on the predicted number of atoms in the MOT and the measured fluorescence, the background fluorescence parameters (with Equations 4.12) or the atomic fluorescence parameters (with Equations D.3) are updated. Because the atomic number estimate is made before updating rates, Equations D.3 should be modified to use the number of atoms assumed to be in the trap when the data was recorded. Thus the values N should become Ni+1 in the equations. This completes the number probability update due to the measurement. During the time before the next measurement, the probability is evolved according to the deterministic evolution of Equation 4.10. The time variable dt is the time between measurements. Alternatively, the probabilities could be checked as to not fall below the pre-determined minimum  as described in Section 4.2.4. This prepares the probabilities for updating on the next measurement. Starting and stopping of the algorithm are done based on the information from the FPGA. The algorithm does not start on the first data from the FPGA. Instead, all experiments are designed with a few second “dead time” where the MOT trapping laser detuning is shifted above resonance. This guarantees no atoms load in the MOT. The data recorded by the FPGA is only background fluorescence. The algorithm 112 knows when the dead time ends and once it receives photon data whose time stamp matches the end of the dead time, it can calculate initial background fluorescence and then begins the Bayesian evolution. The algorithm is ended once the FPGA transmits its final count rate. The final FPGA time-data packet is followed by a series of packets that just read ’‘STOP” rather than photon data for this reason. 113 CHAPTER V ATOMIC FORCES IN A MOT In this chapter the MOT theory in Chapter II is expanded to more closely model the behavior of a real 87Rb atom. The atom is expanded from the two-level or V-atom to the full D2 level structure, in Figure 2.1. The magnetic and optical fields, along with the atom’s position and velocity, are expanded to three dimensions, requiring 6 optical fields. The discussion in Section 5.1 briefly discusses modifications to the Hamiltonians in Chapter II to change to the full atom in 3D. Section 5.3 looks at these changes in the dynamics of the atom, but internally as the evolution of its density matrix and externally as the MOT trapping force on the atom. After laying the groundwork for the simulation, a discovered atomic loss mechanism is discussed along with a resolution to agree with established MOT theory. 5.1 3D and 87Rb Hamiltonians 5.1.1 Atomic Hamiltonian The free atomic Hamiltonian closely matches that of Equation 2.57 with a few small changes. The detuning is defined relative to the ground state energy rather than the excited state energies. This is done as a separate electric field is needed to excite atoms from each of the two ground states. The |Fg = 2〉 ground state is defined with (detuned) energy ∆M and the |Fg = 1〉 ground state is defined with energy ∆R. Here, the M subscript is in reference to the “MOT laser”—the laser field for the MOT trapping transition (see Section 3.2 and Figure 2.1). The R subscript 114 Energy Level With Repump Without Repump |Fe = 3〉 0.35 5.6×10−7 |Fe = 2〉 2.9×10−4 4.6×10−10 |Fe = 1〉 3.6×10−5 5.8×10−11 |F = 0〉 2.7×10−8 2.6×10−14e |Fg = 2〉 0.65 1.0×10−6 |Fg = 1〉 8.3×10−4 1 TABLE 5.1. Repumping field and populations of 87Rb D2 energy levels. Values are steady-state populations (summed over magnetic sublevels) for energy levels. Repump field powers, relative to trapping field power, are 10−1 and 10−10. Values calculated with ∆M = −Γ, ∆R = 0.5 MHz, trapping beam intensity 10Isat, and without Zeeman shifts of the magnetic sublevels. These values were calculated as described in Section 5.3.1. is in reference to the “repumping laser”. With these, the free atomic Hamiltonian is ∑3 ∑Fe HA = ~ ∆Fe |Fe;me〉〈Fe;me|+ Fe=0 me=−Fe ∑2 ~ (∆M + ∆Fe=3) |Fg = 2;mF 〉〈Fg = 2;mF |+ (5.1) mF=−2 ∑1 ~ (∆R + ∆Fe=2) |Fg = 1;mF 〉〈Fg = 2;mF | , mF=−1 where the terms ∆Fe give the (relative) energy differences between the various excited states. The two ground states also have different energies, but because their energy is defined relative to the excited state energy (through the detunings), the energy difference between the two grounds must match the energy shift of the excited state to which they are coupled. These energy differences are associated with the angular momenta of the atom and are defined in [48, Eq. 7.134]. For hyperfine splitting, transitions which do not change the total angular momentum are allowed, which makes excitations |Fg = 2〉 → |Fe = 2〉 possible. The likelihood of this excitation is small as it is detuned from the MOT trapping field 115 by 267 MHz (peak D in Figure 3.9). Even with such a large detuning, excitations that do occur can then decay to the |G;F = 1〉 ground state. Excitations out of this ground state are detuned by 6.8 GHz from the MOT trapping transition [62], effectively making this a dark (inescapable) state—atoms that fall into this state will remain there. The repumping field is thus required to excite atoms out of this state back into the MOT transition by coupling |Fg = 1〉 back to the |Fe = 2〉 (as shown in Figure 2.1). These atoms could then return to the |Fg = 2〉 state and the MOT transition. This result is shown numerically in Table 5.1. The values listed are the steady-state level populations (summed over magnetic mF sublevels). Populations on the left are with a repump field (10% of the power of the trapping field) and values on the right are without the repump field. The decay of atoms into the inescapable |Fg = 1〉 state without a repumping field is clear. 5.1.2 Atom-Magnetic Field Hamiltonian In one dimension, the magnetic field for the MOT was assumed to be linear with gradient −B′z, which agrees with the magnetic field strength for permanent magnets in Figure 3.12 and for anti-Helmholtz electromagnets in Equation 3.6 near the center of the MOT. In a full 3D theory, the magnetic field at any point in space for anti- Helmholtz coils, with their axis defined as a z-axis, will be ∫ [2π ((z+s/2) z+s/2~ µ I cos θx̂+( ) sin θŷ+(1− x cos θ− y sin θ0 R R R R ) )ẑB(x, y, z) = dθ2πR 2 2 2 3/2 0 1+( x ) + y +R (R) ( z+s/2) −2 x cos θ−2 y sin θR R R ] (5.2) ((z−s/2) ( z−s/2− cos θx̂+ ) sin θŷ+(1− x cos θ− y si)n θ)ẑR R R R 3/2 , 1+( x 2 2 ) z−s/2 2 +( y ) +( ) −2 x cos θ−2 y sin θR R R R R where R is the radius of the coils, s is the separation between coils and the origin is along the axis exactly inbetween the two coils. As long as the atom stays close to the 116 origin, the magnetic field is nearly linear along the axis of the coils with a (negative) gradient of ∣ ′ ∣ 3Bz ≡ − ∂Bz ∣ 3µ0IsR∣ = . (5.3)∂z ~r=0 [R2 + (s/2)2]5/2 In order for the divergence of the magnetic field to vanish, the x- and y-direction contributions to the divergence must cancel the contribution from the z-direction. Additionally, the x- and y-direction magnitudes must be equal as the magnetic field should be symmetric around the z-axis. These, together with the linear assumption of the z-direction magnetic field, require that the magnetic field have the form ( ~ ′ x y ) B (~r) = Bz x̂+ ŷ − zẑ . (5.4)2 2 This can be checked simply with the full form of the field in Equation 5.2. Looking at the field along the x-axis is ∫ ~ µ 2π s 0I cos θx̂+ s sin θŷ B(x, y = 0, z = 0) = dθ( R R( ) ( ) ) (5.5)2πR 2 3/20 2 1 + x + s/2 − 2 x cos θ R R R The y-component clearly integrates to zero so that the field is just along the x- direction with magnitude  [ ] [ ]  2 µ Is (R + x 2 + (s/2)2)E −4xR K −4xR 0 (x−R)2+(s/2)2 (x−R)2+(s/2)2 Bx(x) = √ − √  , πx (x−R)2 + (s/2)2 ((x+R)2 + (s/2)2) (x−R)2 + (s/2)2 (5.6) 117 where K[k] and E[k] are complete elliptical integrals. Near x = 0, the field is nearly linear with a slope of ′ dBx 3µ0IsR 3 Bx = lim = , (5.7) x→0 dx 2 [R2 + (s/2)2]5/2 which is half the z-direction gradient with the opposite sign. Thus the form of equation 5.4 is valid near the origin. Do note, in our simulations, the z-direction gradient value, B′z that is used is the measured value from the constructed water-cooled anti-Helmholtz coils. As noted in Section 3.3.1, this values was also numerically calculated, but must summed over many coil pairs with different radii Ri and separations si to account for many layered loops of wire in our coils. In writing the form of the atom-magnetic field coupling Hamiltonian, Equation 2.49, it was assumed that the atomic dipole moment µ~ stayed aligned with the magnetic field. This allowed the Hamiltonian to be written just as the Zeeman shifts of the energy levels. This assumption is kept here, giving the Hamiltonian √ [ B′ ∑ ∑Ĥ = µ z x2 + y2z B + 4z2 g2 Fe me Feme |Fe;me〉〈Fe;me|+ ∑ ∑ ] (5.8) gFg F m mg |Fg;mg〉〈Fg;m | ,g g g where the gF values are as described in Section 2.3. When considering the motion of the atom, this assumption requires that the direction of the atomic dipole changes along with the magnetic field. This can be particularly burdensome when passing through the origin as the magnetic field direction changes abruptly (from +ẑ to −ẑ when traveling along the z-axis, for example). Allowing for the atom to do so also assumes that the motional times scale of the atom is much slower than the time 118 scale for the atomic dipole moment to precess and align with the magnetic field, the adiabatic limit [63]. 5.1.3 Atom-Electric Field Hamiltonian With the full level structure of rubidium, there are many transitions between the various magnetic sublevels. Rather than writing individual transitions independently, and with thoughts towards the polarization of the experimental light fields, they can be grouped by linear and circular transitions that change mF by ±1 or 0. In this way, the lowering operators are written as in ∑ Σ̂q = s (Fe, Fg,me) |Jg,me + q〉〈Je,me| , (5.9) Fg ,Fg ,me with coefficients ([48, Eq . 7.407]) √ s(F , F ,m ) = (−1)Fe+Jg+1+Ie g e (2Fe + 1)(2Jg + 1)×     (5.10) J J 1  〈 e gFg,me + q|Fe,me; 1, q〉  , F F g e I where each term in the sum lowers the atom from the |Je,me〉 state to the |Jg,mg = me+q〉 state. Do note that under this definition, transitions with q = +1 increase the mF sub-levels of the atom, which correspond to σ− transitions as used in Chapter II. With these lowering operators, the atom-field interaction Hamiltonian can be defined in the same form as Equation 2.56, with appropriate care due to the σ±,0 and Σq=∓1,0 119 Propagation direction Polarization P[ olarization V] ec√tor +z σ− [ 1, −i, 0 /] √2 −z σ+ [ −1, −i, ] 0√ / 2 +x σ+ [ 0, i, 1 /] √2 −x σ− [ 0, −i, 1 ] /√2 +y σ+ [ −i, 0, 1] √/ 2 −y σ− i, 0, 1 / 2 TABLE 5.2. MOT Beam Circular Polarizations. These are defined so atoms along an axis are pushed towards the origin as shown in Figure 1.1. Note the change flipping of polarizations between the x- and y-directions compared to the z-direction, which arrises from the change in the sign of the magnetic field gradient in the x- and y- directions. The polarization vectors are in cartesian coordinates. Labels of σ± are in reference to the polarization seen by the atom. In the frame propagating with each beam, the polarizations for opposing beams is identical (see Section 2.4). relationship. This Hamiltonian is ~∑[ ] ĤAF = Ω ∗Σ̂ + Ω Σ̂†q q q q . (5.11)2 q The optical field for the MOT is made of six lasers as described in Chapter I. A single beam has the form of Equation 2.14 with the E~+ component (in the rotating atom frame) as E0̂ E~+ (~r) = e−iφ i ~ e k·~r, (5.12) 2 where ~k is the beam’s propagation direction, ̂ is the beam polarization, and φ is the beam phase. In the MOT, the beam polarizations are circular and their direction is closely linked the magnetic field along an axis, as discussed in Section 2.5.1. The required polarizations are given in Table 5.2 in terms of their circular polarizations. These polarizations are defined so that atoms located along the beam axis feels a net force pushing them toward the origin. 120 The net electric field E~T is the sum of the individual fields ∑ E~+T (~r) = E ~+ i (~r) , (5.13) i where the sum is over each MOT beam present in the trap. To sum the fields, the polarizations need to be written in a common (lab-based) Cartesian basis, which are the polarization vectors listed in the right column of Table 5.2. This gives the field in the cartesian basis, which then needs to changed to the linear and circular basis as to write the field Hamiltonian as in Equation 5.11. In this basis, the Rabi frequencies are      Ωq=−1   −√ 1 √i 0    2 2  ~   〈Jg = 1/2|d|J3 = 3/2〉Ω    ~ z =  Ωq=0  = ~  0 0 1 ET , (5.14)    Ω √1 iq=+1 √ 02 2 where 〈Jg = 1/2|d|J3 = 3/2〉 is the D2 dipole transition matrix element for 87Rb [62]. This basis has the angular momentum quantization axis in the z-direction. To change to a basis where the quantization axis is along the magnetic field direction, the angular momentum vector must be rotated. From Rose [126], rotating an operator that changes total angular momentum by 1 to a basis with a different quantization 121 axis is done with a rotation operator4 whose matrix elements are [ √ x (1) −imγ −im′ ∑ α (−1) (1 +m)!(1−m)!(1 +m′)!(1−m′)!Dm′,m(αβγ) = e e − ′ − − ′ − ×(1 m x)!(1 +m x)!(x+m m)!x! x ( )2+m−m′−2x( ) ′ ]m −m+2x β β cos − sin , (5.15) 2 2 where the angles α, β, and γ correspond to the standard Euler angles. Because only the direction of the z-axis is important to the rotation (the orientation of the x- and y- axis does not matter), we have γ = 0. With possible values m = −1, 0, 1, the (correctly indexed) rotation matrix is the 3× 3 matrix :   1eiα (1 + cos β) −√1 sin β 1e−iα (1− cos β)  2 2 2       R(α, β) =  √1 eiα sin β cos β −√ 1 e−iα sin β  . (5.16) 2 2     1eiα (1− cos β) √1 sin β 1e−iα (1 + cos β) 2 2 2 This rotation matrix is identical to an operation that converts the circular basis polarizations to the Cartesian basis, rotates the Cartesian vectors with classical rotations around the z-axis by (polar) angle α and then around the y-axis by (azimuthal) angle β, then converts back to the circular basis. For each position in the MOT, the magnetic field has spherical coordinate angles φB and θB (corresponding to azimuthal and polar angles respectively). With the 4This operator, based on historic derivation by Wigner, has indices ordered different from common ∑matrix defintions. So the matrix elements are for the rotation operator, R, given by R|1,m〉 = (1) Dm′,m|1,m′〉. m′ 122 atom aligned to the magnetic field, it “sees” an electric field with polarization vector Ω~ B = R (φB, θ ) Ω~B z (5.17) The circular and linear components of this vector are the Rabi frequencies Ωq used for the atom-field coupling Hamiltonian 5.11. 5.2 Matching Simulation to Experiments Because of a lingering mismatch between experimental measurements and numeric results, the simulated MOT has had a large number of additional features included to better approximate the experiment. Some of these are discussed in brief here. 5.2.1 MOT Beam Power The most obvious issue present in the MOT beams is imbalance in beam power, as noted in Section 3.2.2. This is corrected easily by providing each MOT beam its own field strength E0 in Equation 5.12. In the simulation, a global laser beam power P (in milliwatts) is defined and each beam has a power ratio factor ri relative to this power. With these, the field strength for each MOT beam is √ Ei 2 (Pri) 20 0 = × , (5.18)πw2 0c where the first term is the central intensity for a Gaussian beam and the second term relates beam intensity to field strength [48]. In this equation, w is the MOT beam 123 waist (measured in cm) and the second term has an extra factor of 10 to convert from lab-measurement-units for beam intensity in mW/cm2 to MKS units. In our experiment, the MOT beams have a gaussian beam (intensity) profile rather than being pure plane waves as written above. The beam waist (measured as the 1/e2 power radius) is 0.33 mm. For a single atom that remains within tens of microns of the MOT beams, the field intensity seen by the atom should be fairly uniform and close to the peak intensity. However, there is a small change in intensity which can be taken into account. For an atom located at ~r and a MOT beam propagating in the direction ~k and originating from position ~b0, the (square of the) distance from the axis of the beam and the atomic position is 2 ∣∣∣ ~ − ∣ ∣2 [( ) ]2 d = λb0 ~r∣ − λ~b0 − ~r · ~k , (5.19) where λ is the laser wavelength. This has been calculated by looking at two points along the MOT beam separated by one wavelength and basic point-line distance formulae. With this, the atom seems a field strength of i → i −d2/w2E0 E0e . (5.20) Note that this is the correct formula as the beam waist is the 1/e2 radius (rather than its variance) and the field strength is proportional to the square root of the intensity. MOT beam power measurements were done outside of vacuum (obviously), but the experimental cell was not anti-reflection coated. As such, the horizontal and vertical MOT beams inside the cell will have different powers. As seen in the vacuum system in Figure 3.1, the vertical beams enter the cell at nearly normal incidence, so that the beam is almost entirely polarized in the x-y plane, which is S-polarized 124 to the surface. However, if this beam is not exactly vertical, light polarized along the z-direction is P-polarized. For the horizontal beams (which enter the cell at nearly 45◦), the light polarization component in the x-y plane is P-polarized and light polarizaed along the z-direction is S-polarized. For the appropriate wavelength and material, the field strength for each (Cartesian) polarization direction is reduced √ √ by a factor of either 1− 2RS or 1−RP , where RS and RP are the S-polarization and P-polarization reflection coefficients, respectively. The factor of 2 accounts for reflections on the outer and inner surface of the experiment cell walls. In the case where RS =6 RP , this will shift the beams out of purely circular polarization inside the cell. 5.2.2 MOT Beam Direction For an ideal MOT, the laser pairs are exactly counter-propagating and are normal to beams in other directions. This is, of course, not the case in physical MOTs. Instead, it is common while building a MOT to adjust the beam directions very slightly until the MOT “looks” good—that is to say it appears approximately round when imaged and there is a high atom number in the MOT. As discussed in great detail in Section 5.4 below, this adjusts the interference pattern of the lasers to minimize pathways for atoms to escape from the MOT. This is done numerically by rotating each beam’s propagation vector slightly with a classical rotation matrix. This rotation must also be applied to a MOT beam’s Cartesian polarization vector, although done with an angular momentum rotation matrix as above. Based on our experiment’s MOT beam alignment system, the misalignment of our MOT beams is no more than about half a degree, putting an upper limit on the angular displacement of a beam in the simulation. 125 Similarly, it is possible that an entire MOT beam is displaced from its ideal launching position. For example, a horizontal MOT beam could be a little too low, but angling its beam upward with a slight tilt would still have the beam strike the center of the MOT. This could still load a MOT without much difficulty. In the simulation then, each beam has a three dimensional positional offset vector to deal with this. It is important to note that this only comes into account when using a beam’s gaussian profile, where the offset is the vector~b0 discussed above. If the beams are plane waves there is no intensity dependence on the transverse dimension. Only a beam’s propagation direction matters, not their displacement from ideal launching position. 5.2.3 MOT Beam Polarization As discussed in Section 3.2.4, the experimental MOT beam polarizations are elliptical rather than circular. How much the beams are elliptical can be found by measuring the power (of the elliptical beam) through a polarizer. With perfectly circular light, the power through the polarizer will be constant for all angles through the polarizer. For elliptical light, the power will maximize at some angle and minimize at 90◦ from that angle. For Cartesian axes a and b, normal to the beam propagation direction, the difference between the two powers in each polarization can be quantified as (√ ) Pa γ −1a = cos , (5.21) Pa + Pb where the angle is defined relative to the a-axis and the values Pa and Pb are the (measured) powers of the polarization in the two directions. Note that if the beam 126 is polarized completely along the a-direction, Pb = 0 and thus γ = 0. Similarly when the beam is linearly along the b-direction if Pa = 0 and γ = π . If the beams are balanced then γ = π/2. Incorporating this into the polarization vectors in Table 5.2, √ the 1/ 2 is replaced by cos γa for the polarization component the a-direction and sin γa for the polarization the b-direction. Each of the six beams will have its own value for γa and the direction a can be defined as either of the two directions normal the propagation direction. 5.2.4 Magnetic Fields It is also possible to have a more complex formula for the magnetic field. Rather than the linearized form of Equation 5.4, a full form of the field at all positions from anti-Helmholtz coils could be used. Additionally, a background magnetic field B~back could be present either from the Earth, the Helmholtz coils discussed in Section 3.3.6, miscellaneous equipment in the lab, or the lab next door. In this case, the magnetic field is (x y ) B~ (~r) = B′z x̂+ ŷ − zẑ +B~0, (5.22)2 2 where, as above, B′z is the field gradient along the axis of the MOT magnetic field coils (or permanent magnets). Dealing with this magnetic field is straight forward as its magnitude is easily calculated for use in the atom-magnetic field Hamiltonian and the angles to use for the polarization rotation matrix in Equation 5.16 are calculated from this field. The experiment sits on an optical table, whose top is a large conducting slab. As such, currents in the anti-Helmholtz coils will produce mirror images in the conductor. While this effect is small, it could play a larger role when magnetic field is modulated. 127 Including these effects just calls for including another term in the total magnetic field for another pair of anti-Helmholtz coils (with current in the opposite direction) whose placement is below the surface of the table a distance equal to the height of the real anti-Helmholtz coils above the table. 5.3 3D and 87Rb Calculations 5.3.1 Atomic Equation of Motion Spontaneous emission is handled identically to the |Fg = 0〉 → |Fe = 0〉 case with the appropriate forms of Σq as the lowering operators. That is, the differential equation governing the evolution of the atomic density matrix is d − i [ ] [ ] [ ] [ ] ρ = ĤA + ĤAF + Ĥz, ρ +ΓL Σ̂q=−1 ρ+ΓL Σ̂q=0 ρ+ΓL Σ̂q=+1 ρ (5.23) dt ~ where the Hamiltonians are defined in Equations 5.1, 5.8, and 5.11, and the Lindblad superoperator was defined in Equation 2.6. With the 24 magnetic sublevels of the D2 transition for 87Rb, the density matrix is 24× 24, probably not analytically solvable in 3D, but it can be done numerically. We implement this by reforming the density matrix into a 24 × 1 vector ρv and the Hamiltonian and Lindblad superoperators into an appropriate 24 × 24 matrix M. With these, the equation of motion simply becomes d ρ =Mρv. (5.24) dt Steady state solutions for this equation are found via LU decomponsition to invert the matrix M [127]. This is implemented through the LAPACK library [128]. 128 The populations for the MOT trapping transition excited states are plotted in Figure 5.1a as a function of position from the origin in a 1D MOT. The preference for populating the outer most Zeeman sublevel is a result of these levels having only a single ground state to which they can decay. The growth (or much smaller decay) of the states with me > 0 states result from these states having their energy levels red shifted (recall B(z > 0) < 0 as shown in Figure 2.5). The rapid growth (or decay for me < 0) of populations in the outer most magnetic sublevels result from their having the largest Zeeman shifts, which most quickly moves these levels into (or out of) resonance with the electric fields. Additionally, as seen in other mutli-level atoms with ground state Zeeman shifts, at small magnetic detunings a two-photon process couples neighboring ground states [64]. As a test for success of the numeric simulation, the atom can be returned to the |Fg = 0〉 → |Fe = 1〉 model and the steady state populations were identical to the analytic results in Appendix A. For comparison between these two atomic models, the (MOT trapping) excited energy level steady state populations in a 1D MOT for the V-atom and the full rubidium atom are plotted in Figure 5.1b. For the rubidium atoms, the populations are summed over Zeeman sublevels for clarity in the graph. 5.3.2 Force Formalism The force on the atom is calculated exactly as in Equation 2.29. The effect on the magnetic field Hamiltonian is exactly ∇~ ĤzĤz = [xx̂+ yŷ + 4zẑ] . (5.25) x2 + y2 + 4z2 129 0.1 0.12 0.09 (a) (b) 0.1 0.08 0.07 |me = +3 i 0.08 0.06 |me = +2 i |me = +1 i me > 0 V-atom 0.05 | i 0.06m = 0 me = 0 87e Rb 0.04 |me = 1 i me < 0 0.04 0.03 |me = 2 i |me = 3 i 0.02 0.02 0.01 0 0 0 10 20 30 40 50 0 10 20 30 40 50 Position, μm Position, μm FIGURE 5.1. Excited state populations for 87Rb and V-atom. Calculation done for a 1D MOT along the +z-axis. (a) Populations for 87Rb excited states. (b) Populations for both 87Rb and V-atom model. The populations for 87Rb are summed over magnetic sub-levels for clarity. Noting that along the z-axis, the strength of Hz is twice as large as along the x- or y-axis, the force that results from this should also be twice as strong along the z-axis. This is expected as the magnetic field gradient is twice as large along the z-axis. If the system is treated as having no background magnetic field, this equation for the gradient of the atom-magnetic field Hamiltonian does not change. The effect on the atom-field Hamiltonian follows the derivation of the optical molasses force. The gradient of Hamiltonian becomes just a gradient of the electric field propagation term e−i ~k·~r, adding a factor −i~k to each term in the sum of the total electric field in Equation 5.13. Because the electric field is already a vector, the gradient here is a tensor which is most clear when written in terms of individual directions. For the (pre-rotated) Rabi frequency vector of Equation 5.14, the gradient 130 Excited State Populations Excited State Populations in the ` direction is   1  −√ √ i 0 2 2 ∂ 〈Jg = 1/2|d|  J3 = 3/2〉  ( ) Ω~ =  ∂` z ~  0 0 1  −ikE~+` + ikE~−` . (5.26)  √1 √i 0 2 2 This vector must then be rotated to the atomic dipole reference frame. In this view, the polarization rotation matrix is assumed to not change with position to first order. However, because the magnetic field changes with position, its direction will also change with position. This would change the rotation matrix as well, but this effect is ignored to first order, when the change in field orientation is small with respect to the change in the field polarization. 5.3.3 Steady State Force To find the force in the s-direction for the atom located at position ~r, calculate the steady state of the atomic density matrix, ρss(~r), as above. The gradient of the Hamiltonian in the `-direction is then calculated and is given by ∂ ∂ ∂ Ĥ = Ĥz +R (φB, θ ~B) Ω`. (5.27) ∂` ∂` ∂` Calculating the expectation value of the force is done by tracing this operator over the steady state density matrix: [( ) ] 〈 ∂F ss`〉 = Tr Ĥ ρ . (5.28) ∂` 131 The form of Equation 5.27 does include both the purely magnetic trapping of Section 2.3.1 and the MOT magnetic confinement of Section 2.5.1. The magnetic trapping derived from the gradient of the magnetic Hamiltonian and the magnetic confinement derived from the gradient of the atom-field Hamiltonian. Both derive their the spatial dependence from the spacial dependence of ρss. For comparison, a one-dimensional force in steady state for the full 87Rb atom and the V-atom is plotted in Figure 5.2 both with a trapping field gradient of 242G/cm (a current of 9A in our magnetic field coils). A few features are clear. First, near the center of the trap, the force on the full 87Rb atom is larger. A linear fit of the force full rubidium force data near z = 0 gives a force of F~ (z) = − [3.68× 10−16Rb N/m] zẑ, an order of magnitude larger than the V-atom force F~V-at(z) = − [4.3× 10−17N/m] zẑ with the balanced restoring constant from Equation 2.84. This enhancement to the trapping force is a result of having multiple ground state magnetic sublevels for the full atom [48]. For ground state sublevels where the energy is redshifted due to the magnetic field, a two-photon absorption process can couple neighboring ground state levels with slightly different detunings from resonance [64]. The effect is to enhance coupling to the MOT beam that the atom is closer to, leading to a larger force from this beam5. This effect “turns off” at larger Zeeman shifts when the ground state magnetic sublevels shift further apart in energy. This gives rise to a second feature, the change in slope of the force for the full 87Rb atom around 20 µm for these parameters. In general, the width of the two-photon feature is AsΓ2~ δz = ′ | | (5.29)4µBgFgBz ∆ 5For slowly moving atoms, a similar effect occurs in momentum-space (with Doppler shifts playing a similar role to the detuning) and is the polarization gradient cooling discussed in Section 2.5.3 [72]. 132 0.25 4 040 0.2 3 030 2 020 0.15 1 010 0 0 0.1 -1-010 -2-020 0.05 -3-030 -4-040 -4-400 --2200 0 20 40Positi0on, um 20 40 0 Position, μm -0.05 -0.1 -0.15 -0.2 -0.25 -3 -2 -1 0 " 1 # 2 3 PositPioosnit,io un,n sictasl eodf ~ |L|z µ g @BB F @z FIGURE 5.2. 1D forces on 87Rb and V-atom. Red curves show V-atom results, blue curves show Full 87Rb results. Inset shows forces in scaled to region near where the atom in the MOT is. Axes for main graph and inset are scaled identically to figures 2.4(a) and (c) respectively. 133 ForcFeor,c eu,n sictasle odFf [~k] Force, 10-22 N Force, 10^-22 where s is the saturation parameter defined in Equation 2.38 and A is a numerical factor that depends on the structure of the atom [64]. The final feature is a larger trapping region for the complete 87Rb atom. This results from the smaller Zeeman shifts of the inner magnetic sublevels. Once magnetic field grows to a large enough magnitude, the |me = ±3〉 energy levels are blue shifted out of resonance with the MOT lasers. However, the |me < 3〉 energy levels are still red-detuned of the electric field, letting them continue to interact with the beams. This extends the region where trapping can occur. 5.3.4 Velocity The atomic velocity is not taken into account in our calculation, meaning that there is no doppler shifting of the MOT beams. This could be done, and requires restructuring the atomic hamiltonian of Equation 5.1, as each MOT beam will have a different detuning due to the atomic motion. recalculation as the detunings in must written separately for each beam. Without including velocity in the calculation, there are also no sub-doppler effects that may arise. The sub-Doppler effects on the atom can be taken into account simply by limiting the maximum energy (temperature) of the atom when calculating the probability distribution from the potential energy of the atom. 5.4 Escape Channels Even the results graphed in Figure 5.1 is relatively simple as it represents a just a 1D system. Expanding to three dimensions as laid out above and can lead to an 134 60 (a) 40 20 0 -20 -40 -60 0 2 4 6 8 10 12 14 400 (b) r, µm 200 0 -200 -400 0 2 4 6 8 10 12 14 r, µm FIGURE 5.3. 3D MOT forces for an assortment of beam phases. Calculation done for a 3D MOT along the +z-axis. Figure (a) shows the forces and figure (b) shows the resulting potential energy from these forces. The potential is found by numerically integrating the force along the positive z-axis. unanticipated result. Figure 5.3 shows (a) the force and (b) the potential energy along the positive z-axis for a random selection of phases for the six 3-D MOT beams. From the figure, it is clear that different arrangements of the MOT beam phases can greatly change the force on an atom. Mostly clearly, unlike in the 1D case, the force oscillates. This is a result of interference of the MOT beams creating an optical lattice inside the MOT (see the brief discussion in Section 1.3). Additionally, from the potential energy graphs, there are some phase arrangements where the force on an atom pushes it from the center of the trap. These are cases where the potential energies decrease as the atom travels outward from the center of the trap (the green 135 Potential Energy, µK Force, 10-22 N and red graphs). While not shown, in some of these cases the potential energy does eventually “turn around” becoming positive again. In such cases, this effectively just moves the “center” of the MOT away from the location where the magnetic field vanishes. This is similar to the beam imbalance or background magnetic field as discussed in Chapter II, both which also displace the center of the MOT. In other cases, the potential does not “turn around” but instead continues to decrease. In these cases, atoms which found themselves along these paths could potentially escape from the trap. While shown below for just along the positive z- axis, these paths can arise in many directions and typically form narrow channels along which the force is not restorative. These channels are narrow, on the order of the light wavelength, and they grow wider further from the MOT center. Such non- restorative forces in MOTs [92, 129] have been observed before. In optical lattices, similar non-cooling forces can appear in momentum space [130] The origin of these channels can be seen by more carefully consider the electric fields in the MOT. First, consider only the two z-beams. These beams have (positive- rotating) fields, as defined in Equation 5.12,      −1   1   ~ (+) E√0  −  −iφz− +ikz ~ (+)Ez− =  i  e e and E E√0 2 2 z+ = 2 2  −  e −iφ i z+e −ikz. (5.30)     0 0 136 Note that the field sub-scripts refer to the direction the beams come from, no the direction the beams propagate in. The total field is then   −e−iφ +ikz −iφ z−e + e z+e −ikz  E0  ~ (+)E = √  −ie−iφz−e+ikz − ie−iφz+e−ikz T ,2 2   0 and, with Equation 5.14, the Rabi frequency is   e−iφz−e+ikz 〈  ~ d〉E0 Ω =  0   z . 2~   e−iφz+e−ikz Taking a position on the negative z-axis where B~ ||ẑ (for simplicity), the rotation matrix of Equation 5.16 is the identity, R (φBθB) = 1, so that Ω~ = Ω~ z. The Rabi frequency vector has magnitude   ∣ ∣ 〈 〉  1 ∣ ∣ d E0    ∣Ω~ ∣ =  0  , (5.31)2~   1 as is expected. The electric field (made by the two beams) is equal parts σ− and σ+ light. This is the normal assumption made when working with a 1-D MOT. When looking at forces, we need the negative gradient of each component of Ω:  ( )  −ik e−iφ z−e +ikz  −∇~ ~ 〈d〉E0   Ω =  0  ẑ.2~  ( )  +ik e−iφz+e−ikz 137 Since Ωq=−1 is associated with σ+ polarization, this equation suggests σ+ light is responsible for forces in the negative-z direction. This is exactly is what was expected as defined in Table 5.2. Similarly, σ− light (associated with Ωq=+1) is responsible for forces in the positive-z direction. Optical molasses, discussed in Section 2.4.1, takes advantage of these different force directions by increasing coupling to the beam that is counter-propagating to atomic motion in order to cool the atom. MOTs take advantage of this by using Zeeman shifts to increase coupling to the beam which will push the atom towards the center of the trap. The gradient equation above is exactly the situation plotted in Figure 5.2. The phase dependence can be seen by adding just a single additional MOT beam. Say a beam propagating in the −x direction (from the +x direction). This beam has field    0  ~ (+) E0Ex+ = √  2 2    e−iφx+e−ikxi .  −1 With the two z-beams in Equation 5.30, the total field is   −e−iφz−e+ikz + e −iφz+e−ikz E0   ~ (+)E = √  −ie−iφz−e+ikz − ie−iφ −ikx z+e−ikz + ie−iφx+eT 2 2    .  −e−iφx+e−ikx 138 The Rabi frequency vector (with the magnetic field along z still) is   e−iφz−e+ikz − 1e−iφx+e−ikx 2 〈 〉  d E0 Ω~ =   ~ −√ 1 e−iφx+e+ikx  2  2  . e−iφz+e−ikz − 1e−iφx+e−ikx 2 Looking at this vector in terms of forces, we have gradient  ( )   ( )  +ik e−iφx+e−ikx   −ik e −iφz−e+ikz 〈d〉E0  √ ( )  〈d〉  E  −∇~ Ω~ =  −i 2k e−iφ −ikx  0 x+e  x̂+  0  ẑ (5.32)4~  ( )  2~  ( )  +ik e−iφx+e−ikx +ik e−iφz+e−ikz The terms associated with forces in the z-direction do not change, which is good. The z-forces should still be controlled by the two z-beams. For the x-direction, forces arise from all three polarizations. This is to be expected as the field in the x-direction beam has has a component of each polarization (in the circular basis). Note that the circular polarization components result in forces in the positive x-direction—opposite the propagation direction of the beam. From the gradient, it appears thethat force in the z-direction should be the same as when there is no off-axis MOT beams. However, because Ω~ is different, the steady-state atomic density matrix will be different. Thus the average force, as calculated from Equation 5.28, will be different. From Equations 2.27, A.1 and A.2, the excited state populations are generally proportional to the square of the exciting Rabi frequency component (just |Ω|2 for the two-level atom and |Ω |2and |Ω |2− + for the two excited states in the V-atom). Assuming this generally holds for the more 139 complex multi-level atom,6 we can look at the square of the Rabi vector:   5 ∣ ∣  + cos [kz − kx− (φz− + φ4 x+)]  ∣∣~ ∣∣ 2 〈d〉2E2   Ω = 0  ~2  1/2  . (5.33) 4   5 + cos [kz + kx+ (φ 4 z+ + φx+)] The linear component has a fixed value, but the two circular polarizations depend on position as well as the phase relationships between the various beams. This leads to a complicated force in the x- and z-direction as a function of position. Of specific interest, are locations where the oscillating terms have larger negative values. In these cases, the σ+ or σ− components can be less than their associated values in the two-beam case above, Equation 5.31. This would tend to decrease the excited state populations, and with the same form for the z-component of −∇~ Ω, and would reduce the overall force in the z-direction. Under some phase arrangements, this could change the sign of the force, resulting in a force that pushes the atom away from the center of the MOT, as seen in Figure 5.3. As an illustrative example, the specific case for φz− = φz+ = φx+ = 0 is shown in Figure 5.4. In this figure, the graphs for both the 1D two-vertical beam MOT and the 3-beam field arrangement just discussed is shown. Graphs (a) and (b), showing the (normalized) polarization components and the Rabi frequency magnitudes, agree with the calculations in Equation 5.33, where in all four sets of graphs the σ− and σ+ components are equal. In graph (d), which shows the components of the force on the atom, the total force in the x-direction is always negative as is expected because the beam is propagating in this direction. 6Looking at Equation 5.11, the average force is a sum over 〈Σ̂q〉ss, weighted by the gradient of the appropriate component of Ω~ . For the analytically solved atoms, 〈Σ̂ 〉 ssq ss = ρqq. 140 From graph (c), which shows the expectation values of the atomic lowering operators, the Σq=0 lowering operator always has a larger expectation value than either the Σq=−1 and Σq=−1 lowering operators. Compared to the forces in graph (d), 〈Σq=0〉 is minimized at positions where Fx = 0. These are also locations where the light is split evenly between circular and linear polarizations, as shown in graphs (a). This balance and cancel the two directions for the x-direction force as noted in Equation 5.32. Because the two vertical beams essentially form a one-dimensional MOT, the force is (nearly) linear as expected and results in a quadratic potential. However, the addition of the third beam produces strong oscillations in z-direction force.The z-direction force does go to zero where 〈Σq=−1〉 = 〈Σq=+1〉 (graph (c)), suggesting the atom experiences forces from the counter-propagating σ+ and σ− beams equally. While not clear from the z-direction force itself, the potential curve (calculated just from integrating the force) does show MOT trapping as from the 1D case, overlaid with deeper potential energy wells associated with the spatial polarization oscillations as shown in graphs (a). Detangling these two potentials is necessary for estimating the MOT temperature. Adding additional beams further complicates the equation for Ω. The complex position relationship is not of specific interest as calculations for the temperature will integrate over position. However, the phase relationship between the six beams as evident above will greatly change the Rabi frequencies, the steady state populations and the force on the atom. This does includes situations where the overall force is away from the center of the MOT as shown in Figure 5.3. In these cases, the polarization arrangement creates a state where one of the 〈Σ〉q values moves the atom away from the center of the MOT. For example, for the beam arrangement 141 Two vertical beams Two vertical beams & negative-X beam (a) 1q0UlbBB96Lq/h1rq3Webe8==b"1>qAeAsA B<66Hlilc6bbVhBiNaSr8qNLAUEqJ93"P4je9an_/aisjtDxEtzlpUK1O/Fhc9YBvBm0IEJlaBsceN6A4U=/"9lB9EeWI8m"vAYAc6FiObKBpSzNEED3jriq/qn69LjbPUDlE8S8SNBVNcib6BHAAA"=>bW8DPe9njij/EzDKOpcYFmIv9le=4"es6b_aah1 tsxeialtNqAEJ3Ur1q/qh6DE9LBbBU0lAAAB6HicbAVNtixetal/<==QtM2peA8eXc+lnPw9Th585JF0ql+cT7Xf+Fc6RFhdJE7UAM00owozzBWy>Cixttae/27tTi0x0eft6aWlU/6AAAB73ic2V220O7NLuJBx8hNBAEJ3Ur1q/qh69LBbBiyUJBttb0YvHCvYD2lAlEbiyUJBttb0YvHCvYD2lA22027dLOJuxuhL96hq/q1rU3JEAN8SNBVdci37BAAA>"=8P9u74lENtF80fZfMT73lT3enaR"=46esab_1ahs tixetalAAAB73icdVBNSuhx"u6JZOaLld37h2A092F27ARlb2iD7YAv8C7HNv0YM0lbnt=tsB1J UeyB=LP9u64hEqt/8qf1frTU33TJeEaA"N48eSaN_BaVsdtcxit3l8NAEJ3Ur1q/qh69LBbBiyUJBttb0YvHCvYD2lA22027dLOJuxuh< 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 EFYKSS19NGCQok4wX+6QfISILFDKk0VPLTT+KTDtCxdjEkd1ZWj59Ed5ujbai43G9F5ivdVQSCHzNyTyGhI+T+bVSLI1qdmVR8C=+DPEkSXb/OkF7iZ9bidSUdxkW4Ny+ke/VWxemOU1e8RWxbG/O+4aBFh4OVwdY3ETLWajMFAC1hNRK7EkF7BUD+gm8xwBBYMM8KWD3Bf3W/oS/1VwiLvgS/5/beC>1YYC+6cDwTgd/jYbL51H9jA1g5v5+ajG/igQ=zZDdFxENYeSx1UNRCGo44hXw6EfaSALNDEkBVgLwTMKWDfCodVEvd5ZCjY9cdguYb1iA3v9j5gvyDhFTESYqSR1+NkC/o74bXU6Wf+Vxm LvgS=AXP30kMO2ABVxKjVAASrf>0BpcmJTBjx4TAn0YnkCbDYy"zAsAb7mi+dqDJSKNjEo1rbYkAOAnYTkVjKoq+JbsmyDzn0C=4"8de6ab_1ahs tixetalrbYkAOAnYTkVjKoq+JbsmyDzn0C=4"es6b_aah1 tsxeialtlIARAeAIBP713d8< 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>tLilx1eWt7aIlb/Z<+6I/PI8djPBc3xA+QOsPUAM8ZHWMVjj5gy31C1Eac28My/oqTOy/L02OEc+LIOU8IP+jQAV86QrAIlBOiDoCI1eNFKJAyqrLMPpEjAoOMjUQTIYOd52wSRMePrJWa9xivW327dHWmz6e2dEJ6aF8k00/V/naLByEKbV0MyFIZ1E62MUIm6H+IU37v7E2axJLs+MPSQyTd2YoLAUTMZy8jopc3WrMyBECFJe3bjoIilgjIBrb8BaE/0/aJ8ezdd2Wi9WreWw5RIQOOAjPLEAKq1CNOlDQ8AjPAOL8O0cOq/M2/11a5jyH8MPOAxc+dIP6Zo8j3pcMWrEyBJCFbe3IjogilBjI8r>tixetal/<6VV36QKsIAy7V+LLWZnEQ2VLPy0T+okyZ8FcIv6x1aEJUP2MMS62IdmYITHU+M7oUj3p/TocxPOP+8HAj5M11y2MaqO/0O/LOcPj88QAlOAC1DKANLPqAOEQIj5wOerR9iW2dWzeWJt>i+x5e8taaPlM/1<+6O/AIHdjPyc1xka2a80//aBEbr8AQWAelOOiDJCw1WN2KeAIq5LRPrE9AWOdjzQd860/I/baEBd1k+goiI3jbFeJBCEryZMFIBlEU2M6IAUmYdLxdHT32jQvS7yVMI+7PELUJ+sIV0PKI6VLynQWVW Ev7EB6NrqALA7tdBi3BctVSNE8kA3JqUm1q/LhE9ZBJBWE1S"mYv4mnqf4"c9=e_aZEtatSSVkYDqPlS7Sm"A>U=MmrJwA7qlLVbYnVSQtRcS4fYD=k"e46bsaa_1 hsxtiaetals1b_sahi ttxe"=MUrJBAZEEmJwWv1eRL89KEgbaZPtbVft0m0E7v6nURR 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 iYsyhQhR5xe09SKaUibg0gxnKhQ6H1XPx0fPc20mrX>byd1epxW3/ZUtGKZqxX0GiZixlZJRZweIqzViIGQUM/iAaG=hzAvhCpVwLD9Ud1EoOAoQMEAOzDZ/XAZEdz291agdS9kWzix9vW1KC/RLV1ppLyKVBwzUKkN8QmloCy6NfzCpURJswF0A5+ENwCZdL1zR9il7vKexetal/<=VRzQCaDR1vh/vsntxxztzlk9xaZzGAXDqEKAt1wDhp3AxDeUdibIXGmx2GPq0tP31e6bhmnPgPg6inagSa00ywyYzsycNQnwk8CmhQoyCND6zpwRzUsFEAz0+NLCywd1lI19v79K/9eqTA>KnzFKKBMsQ/rvZRqaAQ79K19INwCF0ApURy6N8QmcwQwsYxZGXqKtZ3xedbXm2P0P16hnggiaS0xwyYisycyQUwN8kmlQCyoNf6CpzRJUwFsA50EN+CZq99beAzKTqnKsBF0M/RvQVraZQ33DL1dIR9l7vKe99qeAbKTznKqBFsM/0vQRraVQGHzHIUiGh3DpAhUwDA1oOQEAD/9Ez9ad9Wi/WKpL1zyVCRzhD1xvnkzz=AMlwGHIxzyiiGyUy3UDNhkAlhCpowfDCUz1JowAsQ5EEO+DZ/LAdERzl9vaed99eWbiT9nFqWt p 1/2 ztD"=46esab_1ahs tixetalsRy9Fj04rGC/cA2lM120ytdcbjOLuRCih/8KLB0W8+m+8+W/ls"=y 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 b2Rpm0CHtiFmHxhKyRJ20T+qOm5Y7KmRySEbYhSfa5KMFCfFZHz7KFRKmnkZKoT78WzhKIDCT+RSnR5D9bv2iMbSGat8ZbHtSyhORy2aBZ8mo8MTa9sGRSMBma4m7bqtb/GxSbS7tJNRWFCJ/5UEpKazxkJzhRxvbtHhb88s742GINFUJJVH22bVRRmCCFthFJH+h5ymJE0S+KOf5z7RmkyTEzYDSRa5KvFbftZHzhK2R8mMksKMT48qzGKSDNTCRUna5J9xvHi8b2GFtVZbHmSthHRy20BO87oyMYaasFRZMKmm4K78qKbTGnS9SitGNZWSCR/BUopaaRxmJ7hbxSbtHWb/8p7x2hIbFbJ7VI2J DusysGFOGNqSiDcFbfQi4bOJi/e8ZJGOOflEVMbHGPIPQpwinrxn63KjHdc6qdFcs5wZTc4OG8RuUy2wf8dQ1BzDsBNEmKf1FtDESMNqOLVNKHw8ZNmsVK61zdxQi4bOJS/e8ZJKOOflEVMbHOPIPQpwinrxn63KjHdc6NdVcN5wZTc4OG8RuUy2wf8dQ1BzDsBNEmKf1FtiEqMGqFLG4Fbfl2nUeRwGO4QQirb5Jd/883JdOcfVEnMjH6PcPZpOiKTwBQD6BnEHKx1KteElM4qVLZNOIUw2OfTdQ14zOsGNwmRfIcDFisSGqONVGKMnrNLqMEt1KEBDiQbEJO/J8fBHPjPQp6i38ddc5OZc8yuw jjg==tig==it==> 1/2 0USBbbBLrUi967hHq6/hqa1FrgUi3LJpE=AsN18 SeN4"e=ag_AaWsvtYxotcl2KY3laBsceN6A4U=/"9rBlKpWHgH"pALAQ6SiebiB5S7NgEg3xrUqOqF6YLcboUY22rlBbUL9Bhq6q1/U3rEAJ8SNBVNcib6BnAAAAB6nicbVBNS8NAEJUr1q/qh69LBbBU02KA>"=gAWvYocYFOUxgg7K05ieSQLpHHp 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 kt6Ux>tgG8HkcRTuAUzsGDK6jDHTSgCQyG8Ayv8OZI58V5dz4WrQUnnzmGP3A+fwBYC40qtixetal/kcHTuRUzAGDs6jKHTDgCSyGQAy88OvI5ZV58z4drQWnnUmGz3APfw+YCB0q4/ll/aq0ql1grVUg3EJkERA4Na8aStNtBCVbbScEirnq6LB2A2AUAb>L"6=qQqlrn3hECNgSGBTbpiV6kAeA4"gQAnsCNGEpwkA4CAkNcwjCycRy55"==64s6be1sha bi_elaAAAQ"=hlnGCgVTp4kesgAwNEkACycj"R56=4aes1b_sahi ttxe"nQlghC1ahnsc VtNi8xAeJtUa1l/itg5V85IYO8vyA8GyQCgSHkcRTuAUzsGDK6jDHTSgCQyG8Ayv8OYI58V5dz4WrQUnnzmGP3A+fwBbE40sTHDjHkec3RQT4uzAzUwz/s>GWDnKG6+jbDsHaTiSdC4QryUGn8mAPyAvf8BOEY0I<5l8tVx5t6P3>tKinxTeWtzaHlQ/z8AcAqAJaJiehobE6s6sbNKuUMNioUP6"eAAAAB6HicbVBNS8NAEJ3Ur1q/c=JkhJoeNE9sMu0iBHK0bvBpTw=Z0"64beEa1_ aMsitaxitcixetal/<==AtMmReA8nPw9Th585JF0ql+cXeFf+Rc6dF>tixFeitaxat/lAReA8nPw9Th585JF0ql+cXeFf+Rc6d> 0 2 4 6 8 10 12 14 0 2 4 6 8 10 12 14 r, µm r, µm (b) 1.0 0.8 0.6 0.4 0.2 |⌦+| |⌦0| |⌦| ="aL6Wt9kIt94Sve"7s7xvlCbfEKtA=2>/hf 8ipeeasy=QtGE5wxsT8V/rANfevG71SzIJWE"B4NegaS_LaDsVtdxctil8FvohdvRjCJkF3C7KQ3GZx9MNMrRC7KQ3GZx9MNMrhEABBNngcSVLLDg"=k6DbpqSm9OKtKY70EiMbY4o478N"=46esab_1ahs tixetalAhAY/nBcdiDLVgNSEJB1GzNreTxVGQ5CpyAb7yO4oB3vKdgVhsssO+yNhlw+sOdVcpiKn9/EBSAKA7AA>q"m=Okt6YD0bik4mo44a7e8_N6"s=b1ahtsx ietalJfaS7MMnqfdGVS3BH6ZbIjt5kV8dld6aJNs/pcRTOBnkP2sQT+WR9jOk2acdeESZLVlSoC08/uXSiw/RZjTVA7jRLD0VdmHgRgEXmaidb5AVmj87ZQJRpj0DFzRVmLQmJdQgF7FgAHOXMELaAMSdK3MOTcUSVlI0/XL/yTd5SVkdvdmaAN3/OcATlB8k/2lQV+4Rajx/tpDijyR8Qz7qjbVy5djvbZRk9Yws6VS1HTuvBO8p1ACZSDSHoaV5GZZ6/QEIfmLi>ttex/tkl=tixetal/<=4flo8LA+1LY/MSWeWDhrf2yW/MW9cWdTmsHPJnCOFRapksjJR6+lQ82kktBITZcH/ANvaVdkddVq5yT//7XI0UlJSTcKOm3SdLMVaMEAXOHFgQ7FgJdmmiLRV0zwDpjZRiQ87AjzVb5+jybvRv9ZwY6hSsH1uYBT8O1OCpSZS1oDVaGuZ5/6EjfQLLmLmQQfjj6E56u/a5DZ1uZGpaOVODToY11SsZhSYpZCvOv1yO+8bTzBAY8ui1ZHpswS0hR6iYmwJZF9QvFROvAbMyVjL+ScmLKST0JoUXIT7f/5ydq=ddkNV/v/ATHaZBI2tekQ8Rlj6kJts 1A=g"FYyzaiVWodXlMEKaEOc0c37vzJEBNgSLDVdcin/BARv"=46esab_1ahs tixetalP4zKfcYz20QLmBQyfWnaYROaEVZnNAg=hYKovddEBco33"aeC_CDWdHi2/iABAb"AgDF8i6apVKlrXMMeOGE1czvJ7EvB=N6gsSbihasa AttAmKYBA/A>A=g"FYyzaiMerQKgpd6O8oDKAZbnB3iB2vHhWNCECYafV2QY fez1PiQhWboxEiOe8aT/GyQZ+tNgGzYISL+G7OQTwf5y98s+xdv8bkyXzI+M5OT6VjTjlT8x+TkqlttTIEaZmJgiqhfCJJlzY0nV/B/qX80yfTOGLIzgtZy/aei>txtl1OBpibj1gDVaWujo/yxWrwMEPInMBOR6FjkjGTyxlTw+lktzIRacm7gFqyffJllcY1n8/S/5XI0W5JS7CaOe0SdLt06iEAXtANgU7FwJdGjiqBV0zJChiijEJZqTtTCD5VTy+zDbCWwr/xyAa+8Rdz6k8+TTkxkTXjPjI6nXMkB8OdR+68FyjfkTjOGGTLyIxzlgTtwZ+yl/kateziI>Rtaxctml75"1=Xg+OhJMPdB0emsiF0zbPSwL8L5WwnsAtNhcVCE>rOOe/PtwVVk/IIv4"a=a4t6te4sXaObi_k1darhhsz CtOigxMeYt7aVlFI"j=igsOiJePCBieRsLFTz6P0wl8h5UwbsKt4hBVhEbrlOa/4taVbkjIvv/u9NU4FkN7SaD>d=iOBPAeAi"pg6JYBOstzVwr5/s6heEsOatbk_v1=a6hssb 1thi xieela3tAicxXeOtHa0lN/GiLXcfBp06/Ubmtsa1RXzdL5HNlQHHvBANggGjTBDWiONsBgDDyAj/BJ3pfGBJ/N0x6OmDU3bOaTtisDXy1BRuLvzvcONB5WHBHslgQjBAH/vsgBAPNojxgmGNDfBsfaBtDtxitlsxn>+eu/maYeTiw>=t=xAAc"=AXIP/YLLz4iSWTAVvS7SJRGD"lQ6=4aes1b_sahi ttxeiFtc rsdhCaL1n_ObSa3sAe36Z4s=I"VQel4RFDiGE7"JASAVASivdWDASTN4FSYi9LbzLLpPNYN/cAWIWX9cC=Ut8nlAAACA3icdZDLSsNAFIYn9VbrLepON4NFcFWSWi9dCEU3LivaCzShTKaTdOhkEmcmQgkFN76KGxeKuPUl3Pk2TtNSVPSHgcP4="QlRDG7JSVSvWAT4SiLzLPY/AIXc=">AAACA3icdZDLSsNAFIYn9VbrLepON4NFcFWSWL9dCEU3LivaCzShTKaTdOhkEmcmQgkFN76KGxeKuPUl3Pk2TtNSVPSHgcP3n8OZ83Z833 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 tixWRetNaAlxteAtAaAlu/><>=3k0Aumn1HdufxEWl16l83y1WHWIuYxAR16AMaWIObtAaPi2EH31pgfXDyljVydE2fsj9e+eNixlta=/jiy3gxfj38NRA0nuAitAxEA3DuExxayuHAsmjAInMfaAE21su9u+ANu3oKNn+A9/sj2HdyV+l1DwfspuHH2APAAabuIAaRA11wA+YHI/HnAoeAA=j/Xa1eiOtAjtexxtAl8nA4Rw/c4+H6RHxBX/sIkKVoJL9Or9cGo8O5lOeEP6=zgd8R7Lm0dYHxlfztLRChJQ1rGcKEaSxO_m"MIm3ZNYAnZ0IkdJEKVnUBYUcvYiOKK7w480dBVVIhgbZXHAAt=JNLkE3CPiIER8"Nel_Rtdtbarab4VK9OzGYXI1FoxJM>gCScLLDxZzdrcliiHJBbC0AiAnA0>m"S=Q0fJLa6N8oLo/m+kncE1NyR3bLbX9vYPFRMGSPDIdpiOBYARAq"K0Ta"o=m4c6yeLsvaRbP_p1YaqhTs= 6tsibx"=0JaNoomkc1y3LXvPRGPIpOYRqKT"=46esab_1ahs tixetal8aIgiiyolnU8d3v3VAi8IVRQX3UxWCn80rlLWQoXssyK3nB3ayAIrOHBnwLEAoA4QQ0M=AAFX48Ergsfn3qqKXXzGenvsWxzygX+j3X4y1Yq=NI=/itXxdCv83rVLRQ>tixetal/<==gNYqX1B4j3y+xgszG>sWevWqXzqX0rn3gfWA8EF430QAQMAAu4EonHAwOBAI8I3yioynsKlUnXLQR3VrC8vXdxUXnslKony3iyII8OABHwAEnoA4uQAM0AQF34AE8gWfr3nq0XqzXeWvGWszsgx+y3j4B1XqYNg== 0 we have 〈Σ̂−〉ss > 〈Σ̂+〉ss, then the resulting force along the z-direction would push the atom towards in the positive-z direction—away from the center of the MOT. These phase arrangements are the ones which lead to the anti-trapping forces in Figure 5.3. 5.5 Recovering Potential In addition to anti-trapping, the force curves shown in Figure 5.3 have an additional problem. The forces, in general, are non-conservative. Integrating the force along another path (besides along the z-axis as calculated in Figure 5.3b) result in different values for the potential energy. In such a circumstance, it is impossible to define a potential energy of the form F~ = −∇~ U. (5.34) Rectifying this can be done one of three ways. First, the Helmholtz theorem states any vector field, ~v (~r), can be written as ~v (~r) = ~v|| (~r) + ~v⊥ (~r) , (5.35) where ∇~ × ~v|| = 0 and ∇~ · ~v⊥ (~r) = 0. This first irrotational (curl-less) term is conservative, so that applying the Helmholtz theorem to the force we can define a potential as ∫ U = − F~ || (~r) · d~r. (5.36) 143 The α component of irrotational force can be computed directly frosm the full force [48] as ∫ [ ∫ ] || 3 ′ 1 3 kαkβ ~ ~′Fα (~r) = d r d k e ik·(~r−r ) F ~′β(r ). (2π)3 k2 In a more numerically accessible form, this is [ ] || −1 kαFα (~r) = F B , (5.37)k2 where ( ) B = ~k · F̃ ~k , and where ( ) F̃ ~k = F [Fx (~r)] x̂+ F [Fy (~r)] ŷ + F [Fz (~r)] ẑ is the Fourier transform of the force field. These two transforms, to find the fourier transforme of the force and the inverse transform to get the irrotational component, are three-dimensional. This requires numeric solutions for the force over a large grid that encompasses the region in which the atom exists. The region should be at least large enough such that the probability for an atom to be outside of the grid is small. Similarly, the position-space spacing of the grid should be smaller than the wavelength of the lasers to allow for oscillations in the results. For a fair sized MOT (approximate 50 µm in diameter) and square grid spacing λ/10, this is about 263 million grid points, each which has 3 components to the force. While the calculation for the irrotational force is not challenging, calculating the atomic density matrix in steady state at each point is numerically intensive (but definitely possible). 144 The second approach, much simpler numerically, is to calculate the force in a smaller region, or even in one dimension, for many phases of the six MOT beams. The forces are averaged together before calculating the potential energy. As our experimental lasers are not phase controlled, drift in phase, and potentially even change beam direction, an atom will see a variety of phase relationships between the beams. Averaging data over many data runs should approximate the numeric phase averaging. The third approach, which we have not investigated numerically but mimics the experimental process to make “good” MOTs discussed above, is to adjust beam direction and phases until the escape channels vanish in the simulated results. 145 CHAPTER VI POSITION AND TEMPERATURE MEASUREMENTS Soon after loading a single atom MOT with our water-cooled electromagnetic anti-Helmholtz coils, pure curiosity lead us to check on the power spectrum of photons collected from our single atom. A very clear peak at around 21 kHz was visible and, soon after, an oscillation of the same frequency was seen in the anti- Helmholtz coil current. While it isn’t revolutionary that an oscillation in one property (magnetic field strength) results in oscillations in properties that depend on it (atomic fluorescence rates through the Zeeman-shifted detuning), because the magnetic field strength varies with position, the strength of the field oscillation also has a positional dependence. This should encode some atomic position information about into the fluorescence oscillation strength, allowing us to learn about motion of the atom in the MOT, as well as it is temperature, from the light that it emits. The theory and data analysis technique are laid out at the beginning of the chapter. Following that, measurements of the atomic fluorescence oscillation in a variety of experimental contexts are shown and discussed. To close the chapter, two interesting additional effects are examined. 6.1 Theory Following Equation 2.28, the rate that a multilevel atom scatters photons is ∑ R = Γ ρei,e , (6.1)i i where the sum is over all of the possible excited states of the atom. These excited state populations depend strongly on the detuning of the electric field from atomic 146 resonance. This is directly evident in the steady state equations for the two-level atom 2.27 or the V-atom excited states A.1 and A.2, which all depend on the detuning from resonance. For demonstration, the two-level population and the single-excitation field excited-state population for the V-atom (Equation 2.61), can be written in a generalized form as ∣∣ ∣∣2∣Ω̃∣ ρe,e = ∣∣ ∣∣2 (6.2) 1 + 4δ2 + 2 ∣Ω̃∣ where δi and Ω̃i are the detuning from resonance and Rabi frequency for light coupled to the excited state, both scaled by the spontaneous decay rate Γ. Obviously, a smaller detuning from resonance results in a higher population. As was done in Chapter II, the detuning can be written in terms of the laser detuning plus a Zeeman shift (see Equations 2.76) as δi = δL,i +miδB, where mi is the magnetic quantum number for the excited state, and again the detunings are scaled by Γ, e.g. δ ≡ ∆/Γ. When the Zeeman detuning δB is much smaller than the laser detuning δL,i, which is true for an atoms in a MOT as discussed in Section 2.5, the population can be expanded in terms of small Zeeman detuning. In one dimension, with Equation 2.75 defining the Zeeman shift’s frequency, thus becomes ∣∣ ∣∣2 ∣∣ ∣ | | ∣∣ ∣ ∣2 Ω̃i 8µBgFmi δL,i Ω̃i∣ ρei,e =i ∣∣ ∣∣2 + ( ∣ ∣ ) B ′ zz. (6.3) 2 ∣ ∣ ∣ ∣2 2 1 + 4δL,i + 2 Ω̃i Γ~ 1 + 4δ2L,i + 2 ∣Ω̃i∣ Interestingly, this form shows a clear breakdown of treating the atom in a MOT as either a two level atom or as the extended two-level atom described in Equation 2.4.1. For the two level atom’s single excited state, m = 0 and there is no Zeeman 147 shift, and hence no position/magnetic field dependence in the fluorescence. For the extended two level atom, m± = ±1 and when summing over both states to get a total scatter rate, the position and magnetic field dependence vanishes due to the linear dependence on mi. Of course, an expansion to order δ 2 B would return the position dependence as terms of m2i would appear. Checking for linear or quadratic dependence of the fluorescence rate on the position/magnetic field would inform us as to if the extended-two level atom is a fair model for out experiment (hint: it is not, see below Section 6.3.1). 6.1.1 Fluorescence Oscillations In any case, the magnetic field dependence of the excited state population is clear from Equation 6.3. As noted above, in a MOT, this dependence is weak and unlikely to be detectable in our MOT without much improved efficiency in photon collection. However, introducing oscillations, as done by accident above, should clearly reveal the dependence on the magnetic field. In one dimension again, an oscillation in the magnetic field of the MOT can be written as B~ (z, t) = −B′z(1 +  cos(2πft))z (6.4) where  is typically a small value,  . 0.15. To examine this analytically, we will turn to the V-atom model in Section 2.4 which had steady state populations7 given by Equations A.1, A.2 , and A.3. With the magnetic field as defined above, the 7While it is true that the magnetic field is changing in time, steady state populations can still be used as the timescale for the internal evolution of the atom to the steady state is very different than the magnetic field oscillation period. The internal atomic dynamics timescale on the order of 1/Γ = 27 ns [48] while the timescale for the field oscillations is on the order of 100 µs or longer (frequencies for driving oscillations in the experiment are typically on the order of 100 Hz to 1 kHz). With the much longer time scale for the magnetic field, the atom basically sees a constant field 148 magnetic field detuning frequency (scaled by atomic decay rate) is µB δB(t) = B ′ zz [1 +  cos(2πft)] (6.5)~Γ (see the derivation of Equation 2.59). Here, we have imposed |m±| = 1 and positive or negative shifts in frequency for the different excited state levels are written as δ±(t) = δL ± δB(t). With this, the fluorescence is (by expansion around small ) F ≈ 〈Fl〉+ dF cos(2πft), (6.6) with ( ∣ ∣ ) 〈Fl〉 = Γ ρss ∣ + ρss ∣+,+ −,− and (6.7) ( ∣=0 =0 ∂ρss ss ∣ ) ′ +,+ ∣ ∂ρ−,− ∣ µBB z dF = ∣ z ∂δ ∣ + ∣∣ . (6.8) B =0 ∂δB =0 ~ Thus, the fluorescence oscillates with the same frequency and (unwritten) phase as the magnetic field, around an average value that matches the non-oscillating field value, and with an amplitude that is proportional to both position/magnetic field and the driving frequency. While the values for m± have been suppressed here, it is clear from the form of δB that the linearity of dF with respect to  requires linearity with respect to m±. Then, when experimentally measuring values for dF , linear scaling with respect to  will rule out the extended two-level atom model as noted above. while evolving to steady state and we can assume the atom is always in its steady state value for the magnetic field at time t. 149 It is not particularly enlightening, but the derivatives of the excited state populations for the V-atom are given by [ ] ∂ρss− − 1 ∂N, = ρ′ − ρss ∂δB N [ −− −−∂δB ] (6.9) ∂ρss 1 ∂N+,+ = ρ′ − ρss , ∂δB N ++ ++∂δB with ′ s(1− w) { [ ] }ρ−− = 16s (2δB,0 + δ 2L) + 64δB,0 1 + 4δ2 B,0 + 32δL [1 + 4δB,0 (3δB,0 + δL)] ′ s(1 + w) { [ ] }ρ++ = 16s (2δ 2B,0 − δL) + 64δB,0 1 + 4δB,0 − 32δL [1 + 4δB,0 (3δB,0 − δL)]2 ∂N ( ) [ ] [ ] = 8δB,0 2 + s+ 8δ 2 2 2 ∂δ B,0 4 + 5s+ 16δB,0 − 24swδL 2 + s+ 24δB,0 + B [ ] 128δ2LδB,0 s− 16δ2B,0 − 64δ3L [sw − 8δLδB,0] where δB,0 is the scaled magnetic field detuning (Equation 6.5) with  = 0. In these equations, the two Rabi frequencies have been written in terms of beam imbalance ratio w (as defined in Equation 3.1), saturation parameter s as defined in Equation 2.38, and here s is calculated with the average power of the two beams). In these forms, the (decay rate scaled) Rabi frequencies are ∣∣ ∣2∣ ∣Ω̃±∣ s = (1± w) , (6.10) 2 which simplify the form of the steady state equations greatly. There is another piece to note briefly here. The magnetic field underlies the mechanism that traps atoms in the MOT through a harmonic restoring force described in detail in Section 2.5.1. The restoring force strength κ, as shown in Equation 2.84, is proportional to the magnetic field gradient. Modulating the field gradient through 150 the current in the anti-Helmholtz coils then also modulates the value for κ. In doing so, it is possible to excite an additional resonance in the atomic motion, called a parametric resonance [131]. With a well damped atom, (i.e. large damping constant β), a small oscillation parameter , and driving the oscillations far from resonance (or rather far from twice the resonant frequency, see Appendix F), the influence of the field oscillations on the atomic motion is negligible and our theoretical framework is still valid. These resonances are discussed in much more detail in Section 6.5. 6.1.2 Position Averaging The form of the equation for the fluorescence oscillation amplitude dF shows a dependence on position that is difficult to get at experimentally. Instead, it may be beneficial to look at spatially averaged values for this amplitude based on the temperature and potential energy seen by the atom. Following the damped harmonic oscillator formalism of Section 2.5.2, this potential will be 1 U(z) = κz2, (6.11) 2 where κ is the restoring force, given for the V-atom by Equation 2.84. From this equation, it is clear that modulating the trapping strength will oscillate the potential energy of the atom. In doing so, it is possible to heat the atom and thus expand grow the position distribution for the atom [39]. At high frequencies (relative to a characteristic frequency fc = κ/2πβ, where β is the damping coefficient for that atomic motion), the oscillations are too fast for the atom to respond and it experiences the average non-modulated trap. At low frequencies, the particle’s position variance −1/2 grows by a factor of (1− 2) above its non-modulated value. The characteristic 151 frequency, with the form of the trapping strength of Equation 2.83 is µ g B′B F f = zc , 2π~k where the only experimentally controlled value is the MOT magnetic field gradient B′z. For our large field MOTs this frequency is on the order of tens of kHz. Experiments are typically modulated at frequencies of hundreds of Hz to a few kHz, putting the experiment deep into the low frequency range. Here, the modulation is slow enough to impact the motion of the atom and increase its positional variance. However, with small oscillation amplitudes, the growth is only on the order of a few percent of the non-oscillatory variance, which should not affect our calculations seriously. In this “small” frequency limit, and defining an effective temperature value T ′ as just the variance of the position distribution as via the equipartition theorem, the average value for the fluorescence oscillation amplitude for the atom is √ ∫ [ ] 〈 〉 κ κz 2 dF = dF (z) exp − ′ dz. (6.12)2πkBT 2kBT Compared to the “true” temperature of the atom, the effective temperature is √ T ′ = T/ 1 + 2, just a few percent higher than the “true” temperature. This small difference will be ignored in the remainder of the calculations in this chapter. The form of dF (z) in Equation 6.8 appears to be an odd function with respect to z (so that the average is just zero), but the derivatives of the excited state populations are also spatially dependent through the oscillation-free Zeeman detuning, δB,0. 152 Looking ahead to extracting this oscillation from experimental data, the RMS oscillation amplitude is [√ ∫ [ ] ]1/2 κ 2 − κz 2 dFRMS = 2πk T ′ dF (z) exp dz] , (6.13) B 2kBT or, more generally, ( ∫ [ ] )1/2 U(z) dF 2RMS = A dF (z) exp − ′ dz , (6.14)kBT where A is some constant to normalize the position distribution. With the form of dF (z) for the V-atom, both dF ′RMS and 〈dF 〉 should be proportional to Bz, the amplitude of the field oscillation. Again, looking ahead to the measurement, rewriting the fluorescence of the atom as F = 〈Fl〉 [1 +m cos(2πft)] , (6.15) then the dimensionless amplitude for the fluorescence oscillation can be calculated as either 〈 〉 ∫ [ ]〈 〉 dF A U(z)m = 〈 〉 = 〈 〉 dF (z) exp − ′ dz or (6.16)Fl F l k [ [B T ] ] 1/2 ∫ 1/2dFRMS A 2 −U(z)mRMS = 〈 〉 = 〈 dF (z) exp dz] . (6.17)Fl F l〉 k T ′B The dimensionless form for the fluorescence amplitude has at least two benefits. First, it is easily comparable to the similar dimensionless amplitude for the driving oscillations in the anti-Helmholtz coils, . Second, when comparing to experimental results, photon collection efficiency factors vanish. Rather than fitting data to 153 an exact number of photons per second for the oscillation amplitude, which is complicated by the unknown factor limiting collection efficiency (see Section 3.4.1), the fit is relative to the average count rate, which has the same unknown efficiency factor. The two collection factors multiply both photon rates (average and oscillation amplitude) and thus cancel in the final calculation. 6.1.3 Numeric Calculations The RMS form of the dimensionless fluorescence amplitude, m, is written in a generic form so that any potential energy can be used, including for a three dimensional system with z → ~r and dz → dV . In the case of the one-dimensional V-atom, there is analytic form for the potential and for dF (z) as described above, although it is (likely) there is no analytic form for the integral. For the full rubidium atom calculation found discuss in Chapter V, there is no analytic form for the potential, but it is possible to numerically find a solution for a given temperature. The potential energy is found as done in Section 5.5—integrating the irrotational component of force (in one dimension, the entire force is irrotational) and the force is found as described in Section 5.3.3. For dF (z) (or dF (~r) if in three dimensions), there is also no closed form solution as there is for the the V-atom; rather, this can be done by assuming the fluorescence oscillates at a constant frequency and finding the amplitude from the difference in its extreme values, i.e. taking half the difference in the average fluorescence from when the magnetic field has its largest value (B(z) = (1 + )B′zz) and its smallest value (B(z) = (1− )B′zz) : Fl+(z)− Fl−(z) dFRb(z) = . (6.18) 2 154 This is exactly equivalent to taking a numeric derivative of the fluorescence rate with respect to . In deriving Equation 6.6, the general form for the small  expansion is ∣∣ F (z) ≈ dF (z)F (z)| ∣=0 +  .d ∣=0 Ignoring the cosine term (remember that the atom is assumed to be in steady state at all times, so the cosine is effectively a constant and only its extremes are of interest for finding the amplitude), comparing to Equation 6.6 dF must be ∣ ( ) dF (z) ∣ dF =  ∣ Fl+(z)− Fl−(z)=  , d ∣=0 2 where the derivative is written numerically to first order in . This simplifies to exactly the form of dFRb in Equation 6.18. The average fluorescence rate is found numerically exactly as in Equation 6.1— summing the steady state excited state populations and weighting by the atomic decay rate Γ. These three numeric calculations for the potential (Ui), fluorescence rate (〈F 〉i) and fluorescence oscillation amplitude (dFi) can be done for all one dimensional positions zi or for a multidimensional grid. Then, the two averages for the fluorescence 155 oscillation amplitude of Equations 6.16 and 6.17 are calculated as 1 ∑ exp [−Ui/kBT ] dFi∆z A 〈m〉(T ) = i (6.19) 1 ∑ exp [−Ui/kBT ] 〈Fl〉i∆z A i √ 1 ∑ exp [−Ui/kBT ] dF 2i ∆zA i mRMS(T ) = ∑ (6.20)1 exp [−Ui/kBT ] 〈Fl〉i∆z A i where ∆z is the spacing of points in the numeric calculation and A is a factor to ∑ normalize the potential, A = i exp [−Ui/kBT ] ∆z. These equations are written fully, without simplification, because while it seems trivial to simplify factors of A and ∆z, it is not always so clear while coding (lol!). In multiple dimensions, a similar form can be used most easily where the label i refers to points in a multi-dimensional grid, ∆z factors are dropped, and A is recalculated appropriately (also without the spacing factor ∆z). Experimental measurements of either average, mRMS or 〈m〉, can then be used to fit a temperature for an atom to these expressions, with an appropriate model to calculate U , dF 2i i , and 〈Fl〉i. 6.2 Analysis of Photon Arrivals Extracting oscillation information can be done particularly well by looking at the spectrum of photon arrivals. The spectrum can be numerically calculated directly from the record of photon arrivals by an autocorrelation measurement of the photon arrivals as per the Wiener-Khinchin theorem [42]. It is numerically simpler to bin photon arrival data (creating a list of photon counts per timer) and calculate a power 156 spectrum directly with a fourier transform as is ∣∣∣ ∣ ∣2 f̃(f)∣ S(f) = , (6.21) Tmax where f̃(f) is the Fourifer transform of the binned photon time data [48]. This is implemented, for this dissertation, via Octave to do the Fourier transforming. The Octave implementation returns both positive and negative frequency components of the spectrum, but in a non-intuitive order common in numeric Fourier transform algorithms. The code snippet below takes in photon count times in the array data and produces an appropriate power spectrum with the array indices ordered from most negative frequency to largest frequency for natural graphing of the power spectrum. Spec=fftshift(fft(ifftshift(data))) * dt; PowSpec=real(spec.*conj(Spec))./tMax; In this code snippet, dt is the for one photon bin and tMax is the maximum time for the data. This clearly indicates two times must be defined. The first is a time to bin photon counts, ∆t (dt in the code snippet). This sets a time step and gives a maximum frequency that will calculated in the spectrum as fmax = 1/2∆t. Selecting a shorter bin time allows for higher frequencies to be revealed in the spectrum, but produces larger data sets to analyze impacting calculation time. Second is the maximum time to collect data, Tmax (tMax in the code snippet). This gives a frequency spacing for the spectrum as ∆f = 1/2Tmax. Selecting a longer maximum time gives a higher resolution spectrum, but limits the number of spectra that can be calculated and averaged together to reduce noise. In Figure 6.1a, one data run with 100 ms binned photon data (the binning time used to trigger the bayesian algorithm while collecting data) is shown. In Figure 157 6.1b, spectra for a few different maximum times are shown. In Figure 6.1c, spectra with a few different maximum frequencies (and thus different bin times) are shown for this same data. Figure 6.1d, the photon oscillation amplitudes for the spectra in (b) and (c) are shown, calculated using the method described below (Equation 6.28). In graphs (b-d) here, the spectra and oscillation amplitudes were calculated from the photon arrival time collected from the data run in (a). In practice, the bin time is chosen to give maximum frequencies of 25 kHz to sample multiple harmonics (see Section 6.6) of the typical few kHz driving frequencies. With useable data collection times of around 90 s per experiment, maximum times for a spectra are typically around a 5 or 6 seconds to allow for many spectra to be averaged for a single photon’s experimental lifetime. 6.2.1 Oscillation Calculation From the spectra, information about the average fluorescence, amplitude of the oscillation and phase can be extracted. The measured fluorescence is assumed to be Poissonian distributed8. Defining the background average photon rate to be B and the atomic fluorescence rate to oscillate with frequency fα α(t) = α0 [1 +mα cos(2πfαt)] , (6.22) so that the Poissonian sampled fluorescence signal is Flm(t) = B +α(t). As noted in Appendix E, the average rate of photon collections, 〈Flm〉 = B+α0, (Equation E.2), 8As revealed in Figure 4.3, the fluorescence measured from the experiment is super-poissonian. The calculation in Appendix E there assumes a gaussian distribution a variance larger than the mean. The calculation there shows that even with the variance many times larger, there is little change in the power spectrum. 158 35 (a) 32.5 30 27.5 25 0 10 20 30 40 50 60 70 80 90 100 Time, sec 350 (b) 300 Tmax=5.0s Tmax=9.5s 250 Tmax=6.5s 200 150 Tmax=3.5s 100 Tmax=1.5s 50 0 3001 3001 3001 3001 3001 Frequency, Hz 350 (c) 300 250 fmax=10kHz fmax=25kHz fmax=42kHz fmax=60kHz 200 fmax=5kHz 150 100 50 0 3001 3001 3001 3001 3001 Frequency, Hz fFmmaax,x ,k kHz 0 10 20 30 40 50 60 70 0.2 (d) 0.15 0.1 0.05 0 0 1 2 3 4 5 6 7 8 9 10 TTmax,,max sseecc FIGURE 6.1. Sampled oscillation spectrum. (a) Fluorescence rate as a function of time. An atom appears in the MOT around 10seconds. (b) Averaged single atom spectra for a variety of maximum times. (c) Averaged single atom spectra for a variety of one photon bin times (or maximum spectrum frequencies). In both b and c, only ±2 Hz is shown around the driving frequency, 3001 Hz. (d) Average values for unit-less oscillation amplitude mα calculated from the spectra in (b) and (c). For this data run, the anti-Helmholtz current modulation amplitude was 0.058. 159 FFluluooreresscceennccee A Ammppliltituuddee,, mmαα PPoowweerr SSppeeccttrruumm,, kkpphhoottoonnss//sseecc PPoowweerr SSppeeccttrruumm,, kkpphhoottoonnss//sseecc FFlulouroersecsecnecnec eR raatete, ,k kpphhoototonnss/s/seecc can be found from the tails of the power spectrum as S(f →∞) = 〈Flm〉. (6.23) This appears clearly in the spectra shown in Figure 6.1(b) and (c), which all have similar backgrounds to the peak near 3001 Hz. When this is done numerically, an average of a power spectrum’s tail is used rather than just the spectrum value for fmax. The average rate can also be gotten directly from the photon arrival data without needing to take a spectrum. With ni photons in each time bin, the average rate is 1 ∑〈Flm〉 = ni. (6.24) Tmax i This, of course, is the total background rather than just the single atom fluorescence rate. To get the single atom fluorescence rate, the background fluorescence rate just needs to be known. As was noted in the single-atom detection method Section 4.1, our experimental runs always begin with a few seconds with blue detuned laser frequencies so that no atoms can load into the MOT. Thus, we have a built in measurement of the no-atom fluorescence, B. Then the single atom fluorescence rate is α0 = 〈Flm〉 − B. This rate can be measured entirely without a spectrum and is necessary for the Bayesian algorithm in Chapter IV. For the oscillation amplitude, as calculated in E and in [132], there should be a peak in the power spectrum, calculated over a maximum time Tmax, whose amplitude is α2T α2 2max S (f ) = 〈Fl 〉+ 0 m2 + η 0mαα m 4 α 4〈 , (E.12)Fl〉 where η is defined in Equation E.3, a parameter that describes how much larger the standard deviation of the fluorescence is compared to the average fluorescence (i.e. 160 how close to Poissonian the signal is). The dependence of the peak height on Tmax is revealed in the spectra in Figure 6.1b, with a few qualifications. First, the spacing of the spectral points changes with Tmax. With wider spacing, the peak gets spread out which over emphasizes the peak height visually. Second, with Tmax being integer seconds, there are spectral points that align exactly with integer frequencies, as such as the peak at Tmax = 5s in the figure. This peak appears higher than the peak at Tmax = 6.5 s because the peak is shifted over a bit and widened to align with the spectral points. This lowers the peak heights when the maximum time is not an integer. The lowering is compensated for by the integration method described below for calculating the peak height (see Equation 6.26) and does not have an impact the calculation of the oscillation amplitude as shown in Figure 6.1d. Assuming a perfectly Poissonian distribution or (η = 0), as noted in Appendix E, when there is a large background fluorescence (〈Flm〉  η), the amplitude of the oscillation is then √ 2 S (fα)− 〈Flm〉 mα = 〈Flm〉 − . (6.25) B Tmax In this form, all values are directly measurable. The peak height, or as we’ll see the area under the peak height, is S(fα). The background fluorescence rate, B, is measured from the data run before any atoms were loaded into the trap. The measured, experimental fluorescence rate, 〈Flm〉, can either come from the tails of the spectrum (as noted in Appendix E) or just from the raw photon arrival data (Equation 6.24). The power spectral height of Equation E.12 assumed that we had f = fα, the externally driven magnetic field frequency. This is an implicit delta function in the general spectrum, however the experimental data are certainly not a delta function, see Figures 6.1b and c. Instead, due to the numeric calculation, the delta 161 function is spread over many frequencies. To get the original peak height back, we can numerically integrate over the peak. This produces ∫ k∑=N S(fα) = S(f)δ(f − fα)df ≈ S(fα + k∆f)∆f, (6.26) k=−N where the value N is a defined window size to integrate over, typically done so that the window around the peak is ±1 Hz, i.e. N = 1/∆f . In this numeric language then, directly from the power spectrum it is possible to calculate the square of the fluorescence modulation amplitude as k∑=N dF 2 = α2m2 4 m 0 α = [S(f + k∆f)− S(f =∞)] . (6.27)T dt2 αmax k=−N Note that a factor of ∆f has been lost here in order to return the integrated area of Equation 6.26 to a delta-function peak height. There is an additional factor of 1/dt2 present in this equation. The spectrum is calculated from data are in raw photon count numbers, so that the amplitude measured is in terms of just photon number2. The factor is to return the amplitude to units of photons/s. This is done to compare directly to the atomic fluorescence rate defined in Equations 6.1, which is given in terms of the atomic decay rate Γ. Additionally, rescaling amplitudes to photons/s gives a systematic unit for comparison between analysis with different values of dt (the blue data in Figure 6.1d). Because the value for dF 2m can be pulled directly out of the experimental data, this implies that the RMS form of the analytic fluorescence amplitude (mRMS of Equation 6.20) should be used to for calculating the temperature of the atom. For this comparison, the measured dimensionaless oscillation amplitude is then √ √ dF 2m dF 2 mα = = m α0 〈 (6.28) Flm〉 −B 162 where again α0 is the average fluorescence rate from the atom (which can/must be found separately as noted above). The phase for the fluorescence oscillation comes directly from the spectra as well. Take, for example, the cosine function with frequency f0 and phase φ. The Fourier transform is 1 [ ] f̃(f) = e−iφδ(f + f0) + e +iφδ(f − f0) (6.29) 2 Taking the tangent of the real and imaginary parts of this wave gives ( ) Im[f̃ ] sinφ [−δ(f + f0) + δ(f − f0)] tan = − = tanφ (6.30)Re[f̃ ] cosφ [δ(f + f0) + δ(f f0)] when evaluated at f = f0. Then, if the complex Fourier transform of an oscillating time signal is written as ∣∣∣ ∣ ∣ f̃(f) = f̃(f)∣ eiφ, (6.31) the phase φ corresponds to the phase of the underlying oscillation and can be found just by taking the inverse tangent of the real and imaginary parts of the transformed function. Doing this numerically with the binned photon data returns the appropriate phase, offset by factors of π that are related to the total time and spacing between bins. As noted above, choosing shorter segments of time to create power spectra from limits the resolution of the spectra, but it allows for more averaging of data for a clearer signal. This averaging is particularly important in light of Equations 6.25 and 6.27 as the background fluorescence is subtracted. For individual data runs, particularly those with small MOT coil current amplitudes, , small peaks in a spectrum or a noisy background and produce negative values for dF 2 and thus imaginary amplitudes for the fluorescence oscillation, mα. This is resolved by taking 163 just the real part of the calculated dF 2 and averaging it over many spectra before calculating the fluorescence amplitude. The averages are calculated as below. 1. Photon arrival times are recorded with the FGPA system described in Section 3.4.2. 2. Using the bayesian algorithm to predict times when a single atom was in the MOT (see Section 4.2), photons collected during these times are counted in bins of time length ∆t, as above. The size of the bin is determined by the maximum desired frequency of the spectrum. 3. For a defined maximum collection time, Tmax, the appropriate number of bins (Nbin = Tmax/∆t) are spliced out of the data. 4. The spliced data are used to calculate a spectrum. Total background fluorescence 〈Flm〉 (Equations 6.24 or 6.23), the square of the fluorescence oscillation amplitude dF 2m (Equation 6.27) and oscillation phase φ (Equation 6.30) are calculated from the spectrum, or photon counts, for a number of harmonics of the known driving frequency of the MOT coils. 5. Splicing and calculations are done with consecutive numbers of bins until remaining number of bins is smaller than Nbin. 6. Average spectra, fluorescence background and square of the fluorescence oscillation amplitudes together. Note that averaging the square of the fluorescence oscillation amplitude and taking a square root calculates exactly the RMS of the fluorescence oscillation amplitude. Together with a measurement of the non-atom background, B, from each data run, the three measured averaged values for dF 2m, 〈Flm〉 and B calculate the 164 dimensionless oscillation amplitude of Equation 6.28. When these three measured value have error σdF , σFl, and σB, respectively, the error in the calculated oscillation amplitude σm is 2 2 2 2 σdF dFm (σFl + σ 2 ) σm = ( )2 + ( )B4 . (6.32) 4 〈Fl 2m〉 −B dFm 〈Flm〉 −B This can be done for individual data runs and thus specific atoms, as was done for the data in Figure 6.1. With a particular MOT model, these data can be fit to Equation 6.17 or Equation 6.20 to find a temperature of the atom while it is still in the MOT, as is done in other atomic temperature measurements [133]. The individual atom, then, can be used for other experiments. It is also possible to average the spectra and amplitudes for many data runs together with specific MOT system parameters. This does produce an approximate atomic temperature for an atom in the MOT with those system parameters, rather than an in-situ measure of the temperature of a specific atom. However, as is done with the other temperature measurements discussed in Chapter I, it is assumed that atomic temperatures are primarily a function of the MOT system parameters. 6.3 Measurements Fluorescence measurements were previewed briefly in Figure 6.1, which showed the fluorescence peak in a spectrum measured from a single atom. Some complete spectra are shown in Figure 6.2 for a variety of driving frequencies, but all with a driving anti-Helmholtz current modulation amplitude of  = 0.025. Some clear features can be noted about these spectra that are universal to single-atom fluorescence spectra measured in our experiments. The tail of the low-frequency peak 165 1e+06 101Hz 467Hz 953Hz 1499Hz 2711Hz 100000 10000 1000 100 10 0.1 1 10 100 1000 5000 Frequnecy, Hz FIGURE 6.2. Assorted measured spectra. Single-atom fluorescence spectra with a variety of driving frequencies. All spectra done with a driving current amplitude of  = 0.025. scales as 1/f 2, which is not surprising. There is a wide peak in the spectra around 27 Hz. This peak originates from an unknown magnetic field oscillation that appears ubiquitous in the building around the lab. It can be seen even on simple loops of wire connected to a spectrum analyzer. The frequency is not fixed, but drifts slightly causing the widened peak in the spectrum. There are peaks at 60 Hz because of course there are. As calculated in Appendix E, the spectral tails equal the average measured fluorescence rate. Finally, although the current modulation amplitude is the same for all the peaks in Figure 6.2, the peak heights, and thus the fluorescence oscillation magnitudes, change with frequency. This fact is analyzed in more detail in Section 6.3.2 below. To check the developed temperature measurement technique, comparisons to known methods should be done. Figure 6.3 shows temperature measurements for a single atom in our high gradient MOT with the release-recapture method described in Section 1.4 [4, 16]. In the figure, two sets of data and three simulations with different atomic temperatures is shown. The red data are done with the magnetic field turned 166 Power Spectrum, kphotons/s off with the MOT lasers, while the blue data leaves the magnetic field on during the test. At short times (below about 6 ms) the recapture rate is quite similar in both cases. This is due to slow ramp down of the anti-Helmholtz coil current while turning off the field.9 The field persists during this time and magnetically traps the atom (see Section 2.3.1). At long times, the two cases diverge as the field is no long trapping atoms. With the influence of the magnetic field taken into account, the long-time tail for the recapture rate should be used to estimate the temperature. Thus, the atom is likely around 160 µK, surprisingly above the Doppler temperature. From the data and simulation, it is expected that the atom has temperatures around 160 µK, which should be the “target” for verifying our temperature measure technique. The measurements and simulations in Figure 6.3 may be discussed in more detail in the dissertation of Erickson [134], but serve as a fine comparison for this work. 6.3.1 The Linearity of Fluorescence Amplitude The most pressing measurement is verifying the scaling of the fluorescence oscillation amplitude to the driving current modulation amplitude. As noted in Section 6.1, linear scaling of the fluorescence amplitude would rule out the extended two-level model to describe the atom in the MOT, as this model would predict fluorescence modulation only at higher orders of the current modulation amplitude. The measured fluorescence amplitude as a function of current oscillation magnitude for two different driving frequencies are shown in Figure 6.4. Red data 9This ramp-down is not caused by the RC circuit created by the anti-Helmholtz coil and the filtering capacitor discussed in Section 3.3.5. These have an RC time constant of τ=4.6 µs, much faster than the ramp-down time. This is also much slower than the ramp-down time due to internal capacitors in the current supply, which have capacitance 1.5 µF as programed for our power supply [100]. This time delay, instead, may be a result of self-inductance in our MOT coils or eddy currents in the various conducting elements around the experiment. 167 1 B-Field OffB-field On Simulated T=30K Simulated T=90K Simulated T=160K 0.8 0.6 0.4 0.2 0 0 2 4 6 8 10 12 Release time, ms FIGURE 6.3. Single-atom temperatures with release-recapture method. Measured data in red done while ramping down MOT magnetic field on release. Data in blue have the MOT magnetic field left on during the measurements. f =1154Hz f =1154Hz 0.8 (af0=)3001Hz (s 0caled 4x) 0.8 (bf 0 0=)3001Hz (scaled 4x) 0.6 0.6 0.4 0.4 0.2 0.2 f0=1154Hz 0 0.8 f0=3001Hz (scaled 4x) 0 0 0.02 0.04 0.06 0.08 0.1 0 0.02 0.04 0.06 0.08 0.1 Driving Current Amplitude, ε Driving Current Amplitude, ε 0.6 FIGURE 6.4. Fluorescence amplitude scaling with current amplitude. The same data appear in both graphs, but is fit (the curves in each graph) with a power law in (a) 0.4 and linearly in (b). 0.2 168 0 0 0.02 0.04 0.06 0.08 0.1 Driving Current Amplitude, ε Fluorescence Amplitude, mα Recapture Probability Fluorescence Amplitude, mα Fluorescence Amplitude, mα are recorded at a frequency of 1154 Hz and the blue data are recorded at a frequency of 3001 Hz. It is important to note that the blue data have been (vertically) scaled by a factor of 4 so that it is behavior is clear on the graph. In Figure 6.4a, the lines are power law fits. The low frequency data scales as m ∝ 0.82±0.05α and the high frequency data scales as m ∝ 0.81±0.03α . While not quite linear, the scaling is still far from 2. Thus we can rule out the extended two-level atom as a model for our experiment. Linear fits to the data are shown in Figure 6.4. Here, the lower frequency data have a slope of 5.2± 0.55 and the higher frequency data have an unscaled slope of 1.0± 0.03. The large difference in slope (and larger fluorescence amplitude in general) for the two measurements in Figure 6.4 corresponds to the different frequencies of the two data sets. This is analyzed in Section 6.3.2 in more detail, but we’ll note here that the higher frequency measurement is a better model to calculate temperature. The lower frequency data can excite mechanical resonances in the atom causing additional motion in the atom, as noted in Section 6.1.1. The higher frequency oscillations avoid these resonances and should reveal the behavior of the atom without influence of the magnetic field oscillations. Figure 6.5 shows simulated results for the slope of the fluorescence amplitude compared to the magnetic field modulation amplitude for a variety of models. Here, the slope is plotted as a function of atomic temperature following Equation 6.20. In Figure 6.5a, the three lines are all calculations for a 1D MOT for the V-atom, full 87Rb atom, and a Jg = 1 → Je = 2 atom (referred to as J1→2 through the rest of this work). The J1→2 atom was used in some of our simulations for comparison to an analytic derivation of the spring constant for this atom in the MOT [92]. The MOT 169 0.016 0.016 (a) (b) 0.014 0.014 0.012 0.012 0.01 0.01 0.008 0.008 0.006 0.006 0.004 0.004 0.002 0.002 V-atom MOT beam phase integration 87Rb atom Additional MOT oscillation Jg=1 -> Je=2 atom (20x scaled) Effective 3D distriubtion 0 0 0 100 200 300 400 500 0 100 200 300 400 500 Temperature, µK Temperature, µK FIGURE 6.5. Simulated 1D modulation slopes. Values are slopes of fluorescence amplitudes to driving current modulation amplitudes (i.e., the slopes of the linear fits in Figure 6.4). The two graphs have the same vertical scale. spring constant in our simulation does match this analytic form and thus this atom model has been used to examine some results of our simulation. With the larger total angular momentum of the excited state for the J1→2 atom larger than the V-atom’s, the outermost excited state energy level for the J1→2 has a larger Zeeman splitting than that of the V-atom. This would imply the J1→2 atom should have a larger amplitude in the fluorescence modulation, as it is driven by oscillations in the Zeeman shift of the atoms energy levels. However, in calculating the fluorescence amplitude for the V-atom, the simulation added an additional factor of 3 to its excited state Zeeman splitting by changing gF to 3gF . This was to better mimic the behavior of the outmost energy levels of the full 87Rb atom and leads the larger fluorescence amplitude for the V-atom compared to the J1→2 atom in Figure 6.5a. All three of these 1D simulations clearly show slopes that are significantly less than the slope of order 1 from the data in Figure 6.4. Thus in Figure 6.5b, simulations are of the 3D MOT with the full 87Rb atom as described in Chapter V. For all of 170 Fluorescence Amplitude Slope Fluorescence Amplitude Slope these, as noted in Section 5.5, the calculation is averaged over many phases of the MOT beams. This direct averaging is the red curve in figure (b). While its results are larger than the 1D results, and particularly true for low (reasonable experimental) temperatures, they are still much less than the measured slopes . Recalling the effective temperature discussed in Section 6.1.2 and with prior knowledge of additional atomic motion as seen in Section 6.5, we can add an additional “forced” modulation of the atom to our calculation. This is done by assuming that the atom’s entire position distribution oscillates around the peak position with amplitude, A. This amplitude is small enough to not change the coupling of light from the atom into the APD detection system (unlike the oscillations observed in Section 6.4). With this “forced” oscillation model, it is assumed that the modulation frequency of the magnetic field is much faster than this oscillation frequency of the atom in the trap. This is generally true for most of our measurements and certainly for the data in Figure 6.4, where the relevant position oscillations are on the order of 100 Hz (see section 6.3.2). With this slow position oscillation, the atom would then see the full modulation of the magnetic field at each position, so that the average fluorescence amplitude (and thus its slope) would just be its value at each position averaged over one period of the atom’s motion. For a fluorescence amplitude that depends on position, mα(a), then the average rate is ∫ A 〈mα〉 = mα(a)P (a)da (6.33) −A where P (a) is the probability to be at position a for an object oscillating with amplitude A: √ 1 1 P (a) = − . (6.34)π A2 a2 171 The results of these slow atomic oscillations are shown as the green data of Figure 6.5b. While the calculations above for the “3D” MOT are only looking at the forces (and thus the potential energy) along one axis, the larger 3D environment can be modeled more directly with an effective 3D distribution. A majority of our calculations, and indeed all of those seen above in this section, have been done along one MOT beam axis. This gives a position distribution along one axis rather than a true 3D distribution. If we assume that the atom’s position distribution is small in the other directions (a fair assumption with a strong confining force due to the high-gradient fields), we can map an effective 3D solution onto our 1D calculation. With a well confined atom (and exactly true for a 1D MOT), the field strength largely dictates the probability to be at each position along an axis. With the 3D mapping, we re-weight the 1D position probabilities by the number of points in the 3D with the same magnetic field magnitude. As points in the 3D move away from the axis, the field magnitude grows and thus the higher magnetic-field tails of the distribution are move heavily weighted. This effectively turns the 1D position distribution into a (local) magnetic field magnitude distribution for the atom. Renormalizing this distribution with its heavier tails should result in a higher fluorescence oscillation as the field oscillation is also higher in the tails. While this seems artificial, it does mimic the behavior of the 3D atom. The atom explores regions off-axis that have a stronger magnetic field and stronger oscillation in the field than points on the axis. This numeric mapping method takes these into account without calculating the (potentially non-conservative) force off-axis. The blue data in Figure 6.5b shows this effective 3D distribution with a off-axis grid size of 15 µm. 172 It is still evident that simulations are not matching the measured experimental results. Thus, calculations for approximate MOT temperatures will be dropped from the remaining calculations and functional behavior will be the focus. 6.3.2 Frequency Dependence of Fluorescence Amplitude As revealed in Figures 6.2 and 6.4, the amplitude of fluorescence oscillations changes with frequency. In Figure 6.6, the amplitude as a function of driving frequency is plotted. Here all the driving currents have modulation amplitude10  = 0.025. In this graph, also shown are fits to the low frequency response (f < 500) and the high frequency response (f > 700). The low frequency signal scales as a power law f 0.55±0.02. The high frequency tails scale with a power law as f−1.51±0.11 (or with exponential decay that has decay constant (8.30± 0.36)× 10−4 Hz−1). These two power law scalings closely match the scaling of a Lorentzian distribution weighted by f 1/2. For a particle with a small modulation to its trapping frequency ω0, its position variance should have a Lorentzian shape as a function of the driving frequency [39, 135]. Since the fluorescence amplitude depends closely on this position variance (see Section 6.1.2), it is expected that the fluorescence amplitude follows this same shape at high frequencies. As discussed above, for a fast modulation of the fluorescence rate, at all positions the atom sees the whole range of fluorescence rate values so that its fluorescence rate can just be averaged 10As a practical note, the modulation in anti-Helmholtz coil current was introduced by adding a signal from a function generator to the control signal that went to the coil power supply. Due to both the MOT coil low-pass filtering capacitor and the power supply’s own frequency response, a number of amplitudes for the function generator signal have to be tested and the current response in the coils measured with a Hall probe for each frequency. The amplitude of the coil current, A, was calculated by comparisons to the measured current variance σ2, which are related by A2 = 2σ2 for a sinusoidal oscillation (Table 6.2 gives the relationship between amplitude and variance for different waveforms). The amplitudes for the function generator signal were adjusted until the measured current amplitude reached the desired value. 173 0.5 Low Frequency Power Law High Frequency Power Law 0.4 High Frequency Exponential DecayMeasured Spectrum 0.3 0.2 0.1 0 0 500 1000 1500 2000 2500 3000 3500 4000 Driving Frequency, Hz FIGURE 6.6. Fluorescence amplitude spectrum. Along with the data (blue) low- frequency and high-frequency fits to the data are shown. over positions—producing the same frequency distribution for the fluorescence. At lower frequencies, the spectrum is weighted by f 1/2. Therefore, the spectrum can then be written as √ a f/πΓ S(f) = (f − , (6.35)2f 2 2z) + Γ /4 where Γ is the (Lorentzian) FWHM, a is a scaling factor, and 2fz is the (Lorentzian) peak frequency, which is twice the frequency of the (undamped) atomic motion in the √ MOT [39, 136]. The atomic motion has frequency 2πfz = ω0,MOT = κ/m where κ is the MOT spring constant. There is an additional trapping constant associated with magnetic trapping in the MOT. As discussed in Section 2.5.1, the magnetic trapping force is of the form F~mag = −κmagzẑ, (6.36) with µBg ′ fB κ zmag = κ~ MOT . (6.37) kΓ 174 Fluoresence Amplitude, mα With this, the magnetic trapping frequency is √ µBgFB′ ω z0,mag = ω0,MOT . (6.38)~k The data shown in the spectrum of Figure 6.6 were recorded with MOT laster detuning ∆ = −4.8 MHz, magnetic field gradient B′L z = 215 G/cm, and laser Rabi frequency |Ω| = 0.575Γ, where Γ is the atomic decay rate. With the (laser power balanced) V-atom solutions for κ and β in Equations 2.84 and 2.71, the MOT has a trapping frequency of f0,V-at, MOT = 1056 Hz and the magnetic trap has frequency f0,V-at, Mag = 77 Hz. It is also good to look at the frequency for the full 87Rb atom calculation, as the spring constant for the full atom had a much stronger trapping force. Using the scaling between the full atom and the V-atom confinement forces from fitting the simulated data in Figure 5.2, the MOT has a trapping frequency of f0,87Rb-MOT = 3094 Hz and the magnetic-trapping frequency is F0,87Rb-Mag = 228 Hz. All of these values, of course, are for the vertical axis of the MOT. In the horizontal direction, the magnetic field and thus the trapping strength, is reduced by half. These, √ then, give an additional frequencies scaled by 2 compared to the z-axis frequencies. In principle, then, the spectrum of Figure 6.6 could be made of four Lorentzian √ peaks, all scaled by f . However, due to the high MOT frequencies, it is more likely that the magnetic trapping results in the peaks shown in Figure 6.6. This gives a spectrum of the form [ ] √ Γz/2π Γxy/2π S(f) = a f − + √ . (6.39)(f 2f )2 + Γ2z /4 (f − 2 2f )2 2z z + Γxy/4 The two spectrum formulae of Equations 6.35 and 6.39 are plotted along with the fluorescence data in Figure 6.7. The green curve in the figure is a fit to the the single 175 0.5 Single Lorentzian Fit Double Lorentzian Fit 0.4 Bad Double Lorentzian FitMeasured Spectrum 0.3 0.2 0.1 0 0 500 1000 1500 2000 2500 3000 3500 4000 Driving Frequency, Hz FIGURE 6.7. Lorentzian fluorescence amplitude spectrum. The green curve is a fit to the single Lorentzian spectrum of Equation 6.35 and the red curve is a fit to the double Lorentzian spectrum of Equation 6.39. The black curve is a fit to the double Lorentzian spectrum without the group of 5 “low” values on the interior of the curve, hence its “Bad” labeling. Lorentzian spectrum of Equation 6.35 with fit values a = 42 ± 0.3, Γ = 1442 ± 11, and fz = 100 ± 3 Hz. The red curve is a fit to the double Lorentzian spectrum of Equation 6.39 with fit values a = 20 ± 0.1, Γz = 2608 ± 68, Γxy = 834 ± 14, and fz = 94 ± 1 Hz. The black curve, which should be treated as suspect, is a fit of the double Lorentzian spectrum without the 5 “low” data points in the interior of the spectrum. Without these, the fit parameters are a = 21 ± 0.2, Γz = 1153 ± 88, Γxy = 1500 ± 105, and fz = 143 ± 5 Hz. While this is not an appropriate way to analyze data, it does show that the spectrum equation is pretty spot-on to (most of) the data. All of these frequencies are the appropriate order for the magnetic trapping frequency. 176 Fluoresence Amplitude, mα 0.09 700 4A 5A 6A 7A 8A 9A (a) fα=2711Hz, ε=0.03fα=6007Hz, ε=0.07 400 (b) 0.08 0.07 100 0.06 20 2711 2711 2711 2711 2711 2711 0.05 Frequency, Hz 700 (c) 4A 5A 6A 7A 8A 9A 0.04 400 0.03 100 0.02 0.01 100 120 140 160 180 200 220 240 260 20 6007 6007 6007 6007 6007 6007 Magnetic Field Gradient, G/cm Frequency, Hz FIGURE 6.8. Fluorescence amplitude scaling with field gradient. (a) Calculated fluorescence amplitudes. (b) Low-frequency (red data) fluorescence spectra. (c) High frequency (blue data) fluorescence spectra. In both spectra (b) and (c), graphs are labeled as their desired anti-Helmholtz currents while horizontal positions in (a) are in terms of measured field gradients. 6.3.3 MOT Size: Trapping Strength The “size” of the MOT can be scaled by changing the trapping strength, κ for the MOT. This is done easily by changing the anti-Helmholtz coil gradient. This is done for two data runs in Figure 6.8. In Figure 6.8a, two very different relationships are seen. The blue data shows clear independence of the fluorescence amplitude at different DC field gradients. This could be expected as the current amplitude  is constant for all field gradients. With the weaker trapping of the field, the atom may explore further distances from the center of the trap, but the (relative to DC) size of the oscillation is the same everywhere. The red data in Figure 6.8a show a linear relationship between the gradient and the fluorescence. This could be expected as at every position the total oscillation of the magnetic field magnitude (in Gauss) is larger for the higher field gradient. So 177 Fluorescence Amplitude, mα Power Spectrum, kphotons/sec Power Spectrum, kphotons/sec 0.0003 40 µK 120 µK 160 µK 300 µK 0.00025 0.0002 0.00015 0.0001 5e-05 0 0 50 100 150 200 250 Magnetic Field Gradient, G/cm FIGURE 6.9. Simulated fluorescence amplitude scaling with field gradient. They are linear. even while the atom explores shorter distances from the center of the trap with the higher gradient, the change in magnetic field strength is larger in these regions. These two different scalings is also reflected directly in the spectra of Figures 6.8b and 6.8c. In Figure 6.8b, the peak heights clearly grow with increasing DC current, while in 6.8c, they are relatively constant. So which is correct? Apparently the linear results. In Figure 6.9, a simulation for the 1D 87Rb atom shows the fluorescence amplitude as a function of the DC anti-Helmholtz current at a variety of temperatures in µK. Clearly the results are linear. The two experiments shown in Figure 6.8 have similar MOT parameters—their only clear difference is the frequency of the oscillations. It is unclear why the two give very different results. 178 Fluorescence Amplitude, mα 6.3.4 MOT Size: Detuning It is also possible to scale the size of the MOT through the detuning of the lasers. As can be seen from the “traditional” view of the atomic Zeeman shifts in the MOT of Figure 2.5, there is some distance from the center of the MOT where the Zeeman shifts cause the lasers become blue detuned. From Equation 2.75, this happens at a z-direction “radius” ~∆L zrad = (6.40) µBgFm ′FBz and twice this distance from the MOT center for the x- and y- directions (due to the halved magnetic field gradient in these directions). This blue detuning as a function of position is also responsible for the “turn off” of the enhanced trapping of the atoms with multiple ground states discussed in Section 5.3.3. For our high magnetic field gradient MOT, B′z = 215 G/cm (an anti-Helmholtz current of 8 A to match the data in Figure 6.10) and lasers detuned by ∆L = −Γ and the outer most excite state for the 87Rb atom (mF = ±3), this radius is 48 µm, which is significantly larger than an atom’s position distribution at MOT temperatures, as seen both in simulations and images of the atom (see the single-atom fluorescence profile in Figure 6.15a, which has standard deviation of around 15 µm). This is examined in two ways in Figure 6.10. The red data shows the fluorescence oscillation strength as a function of laser detuning with a fixed magnetic field gradient of 215 G/cm. The blue data show the fluorescence amplitude as a function of detuning with a constant MOT radius as defined in Equation 6.40. Figure 6.10a shows measured experimental data and Figure 6.10b shows simulated results. A few things are noticeable. First, the vertical scale is very different between the two, but that is understandable based on the discussion in Section 6.3.1. Second, while the shape 179 0.4 20⇥105tJ+9alxEb_21t"qxteO=T+7hEiCs8 naZeatJsHaZNnSMxEIZn61etX1WPXoJF8GLZFUWPRSxTA2ZHXL>OBncZ"+AzAC9oi=b8wYK9ePagclL6ZmaDt7JaNKLfoWvH/Qxhv/KZvxxr/dgTa+7HhEZ44ZvaEsEUHh/XXyKZtr6xv6zc5W9L7usnHXoG/iR6NeD5GYdDMghOuhkA0eStFMN6dhaHkIGNjO61<7E6jYBEjL9+zCo=">AAAB9XicbZBNSwMxEIZn61etX1WPXoJF8GLZFUWPRS8eK9gPaLclm6ZtaDa7JLNKWfo/vHhQxKv/xZv/T=rCx+/nvLZ2xx/Ev7K9xJQZh8Hqv"/4oefaW_KaNsLtJx7tal2=4nZOzC+="oAA>B9AicXZBbSwNxEMZnI1e61WXXoPF84LZGUWFRSPeK8gP9Lcam6ltaZa7DLNJWfK/vohQHKvxxZ//xvTdr7agEH+4ZhaZ4EEvhUsX/HZyX6tK6xr5zvLWcs79Xnu/oH6iGDNRY5eMdGOgDkuheAhFS06MtadNIHhjGk1ON"tqC8nZaJHLxrTdg7a+EHh4Z4aZvEEshUxilewait41Hro2I1AOOY+6HdKBEXzCK1uq0UYTcEM/tpGeXkt6>xile5a (a) Fixed CurrentFixed Radius (b) 18 0.35 16 14 0.3 12 0.25 10 8 0.2 6 4 0.15 2 Fixed Current Fixed Radius 0.1 0⇥105AAAB9XicbZBNSwMxEIZn61etX1WPXoJF8GLZFUWPRS8eK9gPaLclm6ZtaDa7JLNKWfo/vHhQxKv/xZv/xrTdg7a+EHh4Z4aZvEEshUHX/XZyK6tr6xv5zcLW9s7uXnH/oG6iRDNeY5GMdDOghkuheA0FSt6MNadhIHkjGN1O6T=rCx+/nvLZ2xx/Ev7K9xJQZh8Hqv"/4oefaW_KaNsLtJx7tal7laB9AicXZBbSwNxEMZnI1e6X1tPXWJFoGL8FUZPRW8eS9gKaLPlmcZt6DaaJL7KWNo/fHhvxKQ/xvv/ZrTxg7d+Eah4H4aZvEZshEHXUXZ/K6yr6tv5xcLz9sWuX7H/nG6oRDieYNGM5DOdhkgheu0FAt6SNaMhIdkjHN1GEEjxT2HLOnZ+zCo=">AAAB9XicbZBNSwMxEIZn61etX1WPXoJF8GLZFUWPRS8eK9gPaLclm6ZtaDaOO61XNQG+jrkZH/IKhydsa4NdM/6/t7S6FX0UAEe4hHu7krhZgKOHDfdLM6Gt5KYZe/NHDhREiv6aGZoh/EHangXTux7vsx9vWxLhcvzo5WvNxJ6<1r8k12I8U1lDkO1oPYTH96mXEddF4Bms0XC6dCw+W1e75qp>UUXFYTr3OEWRx/9NKphqwekGSklof6xIO5mygbbq6meU/gTNqeCYBm6jOI2+1tHtxklo0AbzPIHznLFA65wM/ODUUBJ++gMKkHd3+lYKF3KMwERNSBczqGqZnKoIlgJuyqBq/0UNXYaYsXvl+c6KXwuMIFWMBtrSacqGSsLB0Xcqxnc2N6QZBxBlGJgJ++agxdlBx/Vc77FqSy7X/ajBxh2dUUoPAbMFGlI7J1uhmT0/tfq1RquEn1SydP5RyWeZFEJFdshxYhPrtCw9So26UYKfGSuuYk9hhKFhWYy4qj/H1+FoUUBIymcAdPJOx02+X8GHtCMKcTYE0duzKlzKEwKuH9+0OOAYIeocHK4G6odUS2CrTwPtxsCYdhEZ8JeFRP0yd51EPSun1fmRtqThU0um7l+JGIbPjMoAdhYUx2aXNj7/q7US7F/BqVlxg+yx+aJlIgBGZ6oBNQn>GcBqx0cqcsLqSNMSaBrRwFWuI3lKX+6FYXv1s3K+UZjdkHHm4gBM4uQ+JoAI3UMUawND/DKjw5Lw5786F8zTvzznZ8CHPkf15Akbi=b0leipe/aE/T=Ubqi11C5XfB8dC6ZYTOv1z27r8/5tjewaK/w=NbaiI135Afu8QC4ZmT4vHzZ7jLUw1js5vw+/6KXDuwIDWNBMraaUqISUL30JcAxocuN+QQBBB4GMgm+ga4xklHxHVZ7dFjSK7U/+j1832dU1oPAbMrGlI7J2uhmT01tfq1ROuEn1SYdP5Ry6eZFEJddshxYBPrtCwXSo26UCKfGSu1Yk9hhqFKWxyUqW/I1TFmUgBEymc5d6J>xt2iXxGettMaclY/0vAEBBX+c"ZbHsIx/HKKTdXY242lptjZKPhQaHMWwLliRoHCJqCeiE8LHh6rWva9nYIEAN8SNBv4jb2EWA=i>AAiB+XbcSZANN8YAAI9nra"Wh6EHE8jbWv4HIxsHK/TdKY2X2l4tjpKPZQahMWHLlwRoiCJHCeq2sIH/HxKTKXYd422ptlZKjhQPHMawLWiRlHCoqCJsHIixv/vHNK"KhTYdBXAY42L4r29lEpStbj+ZAKEPbh8QHa6HWMaWnwILAl8iNRZocHXCBJAq>C=eWi2EjiLH86rhvaWnY9EAI8SNBZNcib+BXAAA"=>eEWEibj24v Cs7wrsw+5cYaP625jpqX2j82w8WZ9Z70bd/aAc1+psSwY7tsjCEjkOiyLbBz6TbzORyzzT/dv+8ApPHSSStYpCmjkE9UJiBLO0LYx0xY00zRb76Tbz529yjOWwrC7sPw8w+s2qYc6a0b35Xpb3RZlbZ6n785t922xW0r0wYPL8BwO2LYUqi0J3kb9bE1Ccjtmntlp3YvS8P+SHpSHScpjm1CX9bJpUbO5L20jxcnt8ZZlv3dA/8H+PSpSpSYjtm9CEikJOULYBL306SRbz6T7z5b9y2OWjrCwsP78wws2+Ycqa0635bpbX1+PAl8t/nvcdZ38Z2j x7LSq9/x+SbhFCoXun7yv1Pv5PPgbkGrWbz9INC+EbmWqPeuUbwSz7ESFcWiPq5Um+eVthb4SzqhUQ7qbwvLW/DbCowhGCFXtn+yq1DvxPGgxkMrMbM9GNX+dbMWijwGXiUCTy9GWSdcjiiqGUF+iVwhC4UzyhIQGqTwcqUXFw5yn1qtbPvgGD+rkbMxGN9+XMzWbjidFiGCUwIGyS9TEiWqSUP+mVuhe4bzxhSQUq7wQFSxhCxbz/bG4GethquoVMmM++PDUqS5qwWLibEhc+9NS9TbGrIkygUPCvw1iyFnGXiCj7dvWPzWbCXq7CohbPWvL/w7ZaW5CdMoKyx3RdczFcX7iJ6YleRkMoT0pJ191jlPJTcnP3dO5j5qIErdC1jXIlhySJNiYcbMAcRT0oVpCFH1xCKIpjqkeMum5ERqRjlOK36mMCV0RAbYNSIhCrjC5I5ndPjTP09Jeok7YJdczM3xZ5ixWapKHkjIueqi5V5dPnTPj9J0ocC1xiTFuJepKy5oJlcTZXIcI1PMjdo5JEdR5qaRpjelrO5Kd3n6Pm9M0Ck5Y07RzA3bMYiNWSxhHIjjkCqkeYJ7czd3xM5iZWaxKHpjIkeqc1rCIIhjNYSARbVC0m6MKO3jRlREqdM5cX1loTpJyFiuB0LUfjMMHPK1MxwWj667sd/F03EkkTKKSSzZkRBqPWdkE771FfX/i/6iKJMOOMzxpRx4ZxjMEX4ZfkS0avRaaK/FLnYbEwiKVuJokkaDHkl5QhKjx4aaR4SfZEjzxpBFdxdQ6KxhPkM5UlBaqHnDboYkKkuVVoklDkaJ5uhwQK4iaYaEfbE/ZnpLOqK/iBE0RUSfTM3MdP111xyWz6XiXEkRZSKTk3Bd01s1vyPzaX6701s1vyPzaXj4d6Wx1PMMfU0B/qLn/bEYiKwuJVkkoDHal5khKQx4jaRaSf4EjZxpzOMK6iXEkRZSKTk3Bd01s1vyPzaX XkH5zin7LuF=9FaCL4V3mazgM8joUbQD07woA+>BeM/g3U2Bf4DMVVbg945U7ABGHMDBH+8ok7f0Xt5xAt2lBVHmGzUM4jBUkozFWInVLoFb97a9LbV5mCz7MDjBUWQI037Aou+wB4MVg4UGHgVwH+uGo4AV743w0uIAQ3WIUWBBjDD7MC75zbC9m75bVobVLF9Da87f95b2Fgo3L=V/naFeziD>HQ 6tc AsWh3am1=_AbHaZs8eI6a46=C"CjiQEiwh"QANAn9aiTbcBRSCNSEoY19Uvqrr1Ulatexit sha1_base64="jQihQNnaTcRCSo1UqKC3CKimEcw=">AAAB9HicbZBNS8NAEIYn9avWr6AAAB9HicbZBNS8NAEIYn9aN"h=Qw"c4Eemai_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 1bS0JEDrYRT5Ly8zx2rak+vWQZ6ykMlDy6DKoT1+n5mdNLIDOiOrNFphJySGZP9zt7J6j2Qtjvda2eZGd+SnAmJWiF6gMZHLkYDPRdrwSF5oRS6bgQMk6l8yKD+o51dnLmDNgIiOFOqNhpGJiSPZ79St6JtjfQvjed22GZndlSmAFJXig6LMYHPkVDfRwQrKosl3W6v1hyJjZbU1jZe7pIrAaCoVIOPEphZFknRGQ6kMlDy6DKoT1+n5mdNLIDOiOrNFphJyZPSzt9J672Qjjvta2dZGe+SdAmnWiJ6gFZHMkYLPRDrwdF5SRJoY8g8RLxgnqEkRSif50rl2X1vVwkZFOEdhkptIyPiOjCJVnobrMaAA27jIZpNjNeoZkbD1HU6JJZSjZ1dyQhJW9vS6psO3Ilmr1oDKlfQQPwYVLyZ1gXFvWlmrn2+5Gfe0aSvRti2E6q7kzgPxGnyRh8FLr8iYDgLJdR5o+5TFKS6wDrMd6RSyphfKQrlos63Whv1jyJUZbZ1jpe7AIroaCOVIpPEFhIZPOCVoraA7IpjeZb1UJZj1yhWv6s3lroKfQwVy1Xvlr25f0SRiEqbZkFEkkgxnR8L8YgJRo5SFSwrd <=/ZlQrzZUKcx=igt+>eBdAHxi4NS0Koe8++E3PIJEZduWhCB1HQ6AgX/SUqRTzHkvZiO/XS9u1zl6d86IxpO/OOCKodQWxZY2udDUJ8qEieiYNnpYENIi8ieqQJ2DUuwYdxZQQobr6Wv0pIDpgtVm+b6jp8XdHoRl/9z0FX3ZBNckBRTizU6gcHg6QB3d>hetqxat=lFetqppYpNlEjiZEmid80qIdtJbe8Do29uXW6YpUQxDdVQAC+opZXrXbRWS1/0FzN3UkicTRB/HgHn6863gIBIv+dOQJhewUZZQKeO1xO6K+O4Quz+xP/zBcHZgStA=/=axit out HtblaBsceN6A4n=W"HUbfLdzUQ+"IAQA19ei4b5BVSXN/EpYE9HvMrppZ6rWva9nYIEAN8SNBciH9BAAA>"=QYz8LLbTpMHp6rWva9nYIEAN8SNBZbciH9BAAA>"=QYz8LLbTpMHEp/XV54e1QI+UdfU"=46esab_1ahs tixetalU"f="Q4Yeza8_LaLslixaetp 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 OOO/g+8K+qZs3H3UQIBeRnoJ1IOzeQ8ZnwQhHQcd/vXBVgd6Bl7/xH/gOz6UOB+iITIRwcvk63zNTq3ZFFSXX/p0AST9URSlMXGoC6nuXdcpz8B+xjQA4bKVxm/TOsKoZWUUeEJde4ZShrdsBX6MHUgBUqiGRfk1NEZCXT079DlnoEd>8tjibxmestWaEl4/r=SXRldoX+p8bjATVmWsodUEr4SMsXqUB1GfTECn7DaEPeattxi6H1OxE>nBDN7ZTUCeEX1RfHGhqeBOUZMkXisgr6Sd4ZdJEUUKW/oFsqT3mcVTbBAzji+68gpvdQXwoQXIlnRI93S+08/OH6tK/1u4HOPQBzzxcayegS=g=2"x=vYxSt=x"h>vApAUA7B7+030ifcbbEZPB9NeSS8PNUAlEAIFYo3P9Ya8vKWgra1KibPfEWvr93aIEYN8ANBSbcZ3+iAAB>"AYS=0 rXSIIQzWCIclcpkHuBekkiCEcOdaeRrvzeKuRTc0D+YCjVekiB4pa0pzw+zwgPsDEcCdaORdrezevuRKc0T+YDjPs++sPDwzgwz+0ap4Bpd+scPWDhwCzcgTwIzX+k0OaapF4IBCpQdI9e9ksdqrruKW1TvHbpGl4VVctz0Ssr7kuB3uvi7EBC/aZR9zeTvjuYR0K+cDkCcsq9K1rbGvVt4s703vuB/7huZWTaF9Tsq9K1rbGvVt4s70wzDwzg0a+4Bppd99sqrK1vbG4Vt0s7u3v7zBu/ZhuaWTTFHWpIlIVCcCzQSXrIkcBeukikEcCdaORrezevuRKc0T+YDj3v7B/Zhuak+uceBWrIXTSQVpCIIlFTWH qOFS8ZE7uSYiQnnzsvCVusVYLIcArpcTDUKzROtYricUQQwK6meuTrAkmEXgBzRAVqmnRmgYricUQQwK6meuTrAkmEmzkKX/X8jLtJcFBqwcvV3mLuVD9uCnsFQlQPYSZ7zDISuy+rANwopTcvxCDlUAkX8/LtjcFJqwBvVlnmqAVxACRvzwBo+NmuDyzS3ZP7QSDQnlCFu9uuLD93cmwVFBgXABlRqVOmQRYtUrQcKQPwm6rezTkAzmImkKLXD/jtucyJ+rANwopTcvxCDlUAtXgBzRAVqmnROtYricUQQwK6meuTrAkmEmzKkX8/LtjcFJqwBvVcmL3VDuuC9sFnlQQYSP7zZISDlA+UrxpDACwTovNcuy HWuWCggzXZxyKf0TZ/iGI4/eemZg4aE0qnWaZQ6HwTNECUa+URA/tjEV7ZqgacCZKC0sAtUNaIC6NwWtt/6mZWgtGu/UXw/nHAgWH+tT2IRn7yNNzYaUMsg/CZXiut//WCGteWgcmIasZZ0iacfaQZT0x0EO+RKORAjN0UVAr+iR325RpA5T5nTAJNqzXN2yfYDxqMSEYTa+0ROQUjRVAfANNAVa0RjOR0K0+ZEZxcTiQafsaY0MZSaYmqgaeDGyWf/zu2wqtX/N6qtnZJZ7/TXTC5gRT5yEzpW2r5i+335tpi5A5rTHJWq4Xz2gfyDnqTSHYUHTnygz4WHrAit3+52pE5R5TT7JnqNXq2zfyDaqYSMY erR4yJx9jKS+kkeqiLJ65VIKiRj98bskZE9KrO8RyL7pGRmjHSqj67Cq9SqdL9n//ibJEAY7iwS745qoJ5KaKmg7ADdR6rcx7eKIOsB8wmvC+nIY7qPgRczC5qVnLboYFSLq9K5gWd9cLKaBtvpIzPmz3jKLJF79JWRLytDzI3RJjJjyk1rIejRkyexyjjSkkie5iiJ85ZIriyjG8Hs6Z99Lr/8Eyi74GJmKHAj6j7RIk1DyRJ7JK3mzptaL9W59LFoLVj5zRP7I+vwBOKq6k1jZs9r8y7GKOBwv+I7PRz5jVLoFL95W9Latpzm3KJ7JRyD1k7c/6Hd9AEgLK6KRJbqn4qSCiqYmIeijjkr8j5iRIJekSjxy ixet=lT5A==ST5A==ST5A==x>ttS tAAAB7XicbZBlatexit sha1_base64="7gxsz35rV55iow5b8W6EqAAAB7XicbZBaeBFdw3i/vB/9fYpWoeYvVLSdsvkRtZNR2pkPtNpOMYXvq/2kDYYBvJCeHMvYX0EjGBeddR42XR7F683FOJjoGXiPpkSPNXpN2tCe01R68nXZXI6EE6SxM/wiSdNvvpKUvNY3BNl8/KZsT3dr6Xsujch1ab15+3svCGMWgGsjYZddLg9sZassTMNlvpPv/NPRF3wRBKfBorV0sctv25tCMGqYDjv2HgYXjsdp2vFkFdoNPp1ktk1SnpIYxYwWNCG+v7315hbX1jc6u4XdrZ3ds/KB8eNY3KNGUNp0SwMxEIZn61etX1WPXoJF8FU2NNGNKdY8esBj/3dXZYdu416HjbX3hv1v7W+DCG6ZMqsd3sYMLaOT9tNlvpP2/vPiFtw/BYfsoSVNsVtk2XtoM2qCDfvvHBYBjRd82wFJFWoFPX16tP1ZnMI/xSwPNlvTNs9aOsLgYd3ZsjMG6WCG+v7315hbX1jc6u4XdrZ3ds/KB8eNY3KNGUNW1X 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 7+8nIiMafpIh84tiln7jV+iRCadcw8c5x9+nzYG7<8Va/lxteXh4zM5BxhcR8iCt9sffXlijPhzV3RGM3CFdAh1cA4qMAMIMXh4pBARAqcad1FACiR33GVljziPfsvX9filaCc889wxz5n7+GVY8lYqalvislCAa19AiFf3sGX3vziPfiPXlfz9jC38VcGxi53zRGFYCVA/aa1etix>e/al8c8G tBZZrbyc_iAXb7"Bxs"e=sklsc0By>QsttqWEw65W38gb45awaotit5lVbri573AzAi"Bk70"Q=q4668e5soa5bV_51zaxh7s= 6tsibx1eht aile"=ks0yQtqE6W8b5woi55Vr53zsxg7"=46esab_1ahs tZgeBxbxsaiXcBA7A>A=k"0ystqQ6WEb58oiw5V553rsxz7"g46=sae_1bhsht>iixeea/V<8lYI71GR+Fzqn45x98xAwAcA8X8BltCta<9RiCfasdXcvpi4fMPhlMzhj>3iVeGai/3VY7GR+lzvnV5s9fxjwicC8X8ilPCza39Gi3fFAa1hdhAMcMqIpXA44BR> E~ = EÊ D̂mid ~ D̂out ~g6tjtcutFf1aPxb>IuF4xlMPgLSBL4DMZZbscaitH69FBIAgAZAH>A"==PcI4n3fPcNYf"aeI_esMaY"=c43PNfaIeMYnvuVfTZtcS tZiP<6lAaItIe=xHigtF Msfh4a>1B_cbLaxsPet6Y4e=a"NH3scg"YAcAj9SicbtDZSTMfFVbu1vuntjYgcH"s46=sae_1bhsati etxlUtRiKxLemt4aplE/xtti4xleOtEaSlk/Y<4=P=cgnYHSHmOkTRueEAcAP1LDitCCBLtcxuDEEO1xpBAr4FA8mbeF4fe19yZ4eF4SgslU>Z/ukxMmHUnBihak=VS0/bfmMhvCwmnLXMtVx4tJlWaBsceN6A4n=W"CjCQiiEhwQ"NAnAa9TicbRBCSSNoE1YU9qvKr"U1qKC3CKimE>=A6rWva9nYIEcN8SNBZbciH9BAAA>"=wcEmiKC3CK6rWva9nYIEAN8SNBZbciH9BAAA>"=wcEmiKC3CKqU1oSCRcTanNQhiQj"=46esab_1ahs tixetal<xbtYpENji8iDq8eu2xWdUQdrCoZWbI1l6ppmQ9DbV8A0+dp9XXR0SZ/ZFNqR3NciTgBkzHHB6RgdvZQiweQeIUnUIK3g+O8xOHOK6Q16Oxzz/BHcu+Pd+gSeIJneIU3Z+K8OO/Ox6K14OQzxdlXCQhYNpiEtiT8uqOdIJletDz2buoWiYZU6xed4QgC8otZWrxbJWN100E6zIBpepZQxtxDZmitpi9xXeNtcaBlH/6HigQzIUnBIi3T+R8cOkO36N1qOZzF/XH/u0PS+9SRAl=X/oaXe E Haetxxi tb p"2=24X2TbKidEHqeCjjCv2b=4oE>"BAiAX+ZiLbNBASWNIEaYH9vvHrQh8LCEPeJqRCZoli9LtWvahWl6LHe84iqCoJYHiRWldwHMPaKhZKlj4pH2x2HXxT/Ks/sIIHEWWHsIx/HKKTdXY242lptjZKPhQaHMWwLliRoHCJqCeiE8LHh6rWva9nYIEAN8SNBZbciX+BAAA>"=EMHXaIQKh2PTK/ZHjYtdpKlH2x4sW2bjv4vjb24WE=">AAAB+XicbZBNS8NAEIYn bXq1pj8c322tw8bn0ZYlwZP3rdyvp/s8iA0+7PPH8p0SLSEWYbP5d286z65cat+aYEjBkcL0YBp+YiLtksjSY/pwAZdTZ72b5Yxs0BLiOCUkJj9jCtmSp2SpSAH9+/8Zv53ZlSnjtXc71Tb6bz3b0RqRYb2zw6pTm7Cz95JbU9OyL20OxWxj0rLCOwUsJE9zCzmTpzSbSyHO+j8Cvw3Pl8n6tzcbRXw5b9y2OWjrCwsP78wws2+Ycqa0635bpbX1jc2t8nZlZ3dv/8A+PHpSSSYptmjCE9kJiULOBLY00xRP78bw6w7s529+2YWcrqjas07613bO w7CqcUWQGSGhixSzybi4jebhEu9VTmI+UPwUFSiqdWzx+kGgxbM9GNM+XbzWdjiGFiwCUyIGTS9cEiWqSUP+mVuhe4bzxhSQUq7wDPbvq1ty5nFXwCqhCoWbP/vLbL7v+/CPwbcWzoGCGhqqhCQwjXiFynS5iyUtV14qhvqbdPiDFgwGUkI+Tr9xEbWMS9PGmNuMe+bXxbSzUW7rM9GNM+XbzWdjiGFiwCUyIGTS9cEiWqSUP+mVX+WMFNtGq9vM5bwxCrP+7kyGngXDCPhbovbq/1Lue4bzxhSQUq7wr+kGgDPbvq1t7y5nFXwCqhCoWbP/L7vh 5jdiPInxTqPMjW9HJk0Ioxk5eZYaJK7pcICeruzjRhRSNYbAR0VCMmcciJylX1dEqjO3mC0AYCCIi1cCd3dPnTPj9J0okeYJ7czd3xM5iZWaxKHpjIkequ555IrSjphyNoblRTVXMc61KMldR5RE5RMqcRTjolpOFK11CiFJpyolTXc1Md5ERqRjlOK36mMCV0RAbYNShIjCrI55dPnTPj9J0okeYJ7czd3xM5iZWaxKHpjIkequ36FJqjP5dE1cMTlXypoFiJcC19J0okeYJ7cjlOK36mMCV0RAbYNShIuqxeZkCI5jdpWHiKMx3azjI5rdP5Tn sKkkukVMYEE4fSaRaj4xQKhk5laHDojxkZzp6OJKXiZEwRKSBTi30dv1a1/zbXqnL0/BMUf1MP6xWF7dPy6sHsa1vyPzkEXi6KMOzpxZjE4fSaRaj4xQKhk5laHDokkVJuwKiYEb/nLq/B0aXzLPHyxvw1Msk1k0Ed0BW3KklToKJSiZ/R/kfE1X4iQ6hK5MaODzkpVxuZKjYEb4nfqSBaURMaPjxU7dFMP1xW67dFMRZSKTk3Bd01v1PzyX1adB0kT3SZKkERi6XMOKpxzjEZfS4Raa4xjKhQk5lfDokkVJuwKiYEaXb/nLq/BdF7M6PWUxM1f0 WH>7=ebegi44Vt4bwWux>iI4utVtG5WU3F28foDgV7bC9M5D73BBIuAww+VBG5LoVCmDztMBjAU3Qx0A7wo4+>BtMiglUaH=/ mid out +nBvF"1= 4P69e+sUaLbl_S1saphssz >tcixxtettFRbAAxtp+B8e=L1+eO3h=MAFHpZvgQItF36iAAA"A4Bv9hHeiAczbvZ6DbLhSig8Mtx4FsI"bAPA19FiubtDtS6MpFLb<1luattpeLx6ittF PsIhxag1L_ZbcaHsBeA6>4==U"3vQnp+M2OzL/B8xRRA/x2pnB"e4Le+aO_haMsFt/xt2pvQa4p6Ltut/A8BRHAcxZpLBl1IFbMgxPFS 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 AC3pyKnyg4Boa3obT1ywzKFuE+KqllO7YiouwvIQtvS+TSk588LCmHAAOBlx6A+ShC80gPATp8AsvzcR5dd6Vj0SryjlmPuG8nMYfPZGS=A=l2>xeialt<=/AZ=GYSf8PnPMujGlSmrVyjd0d>6ixttae/tixetaglKlSgpACK0mLaoLvE68TKRkkFITOBSSlzotCy8I8yzwmg8oTTtYw3YO2q52vBiC+8154Sh+SvgQAvPuAiv7sl5la+EuKKFwB1zbyoyog4Ty3nqABC8pc3+AQOuH7llxuBwAb+llbABASANBEEYM9HXp/4V5Qe1UI+Udf4"=s6e_abh1ats eixlta9AABApi9HZcbSBNA8NYEIan9rvWH6pHrW6a9vYInANESN8ZbBiHcBA9avsWxrt6tpaHe"=QYz8LLbTpMHEp/XV54e1QI+UdfU"=46esab_1aA>h 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 AZITdsZODZOVjS8O7UclMk+4vT0U3LYZFWUd9egmnojEzvtI3MfKEjyYmjJItOJD5rGOoker2eMbifBunepo54dIrsMj/LKI/8W0SgcfzJg1Onk/izZii4jLFjLDYiOTjRU3KxscMKTiIZSxv746OzAKDHWCroTpOld1ktlYrJeueJVvbncVfBzqECueq/81e1LABOzDKgHiC4oWpDlR1LtuYGJuunJOvcnzVQBEqqCDe8/L1V1kmOQ6DEVv677SxTZxiMKKcZxU3ORiTYijDKFZUZAcdODcjx7+MYv33UFj9Rntzy3TEJmGti5eoMeD2Bi58Fprd/MUKSWOcAgikOZdBgDDiDWERGVnu7u8O+cM3v0UYFn9gtjzE3fJymt/J5Goe2MiBnp5drM/K/WSczgOkiZiBODgi4WDRLuGunO1LmkQoDOnuGuLRDW4igDOBiZikOgzcSW/K/Mrd5pnBiM2eoG5JtJmyEf3tzjng9UFY30v+zAKDHWCroTpOld1ktlYrJeueJVvbncVfBzqECueqMc78jODZdIAZUjFLDYiOTjRUAzKMDoH8WvCurLoETIpKOqledk1Ok4tSlTYsrCJee/u1e1JmVQvDbVn6c7VxfZBizKqcEx3 3dXB/OXga=RD4nSHq//QrfFwW+qOby3ed3c69gT=Z/BaReziH>HnV7uCU1PGsByMGsSS7dIUOooT1VPAl+gp>X9XUqEX0Sf1UGXBXM4sZSEdsUNomTjVkA8+opRXlX0RiSX/NFUqk3icgTUBHz66gBvdQhwZQeIJneIU3Z+K8OO/Ox6K14OQzx/Bp+z8AjTbcmosdWaES4MreXBU1qgfCEnT=DEETn<7ECH1lg6Hv6tBQdQhxZIeIJte3U8ZEKOO6/Dx1Kz4TQ/xuBEzPcSafegg/=qtF0XXqZcNrkTRzi4UHggHE6vBwdWhQZnesJIe+UmZ8KOOb/6xOKj4zQHx8BuzycdaSe=go=/tixetal/<==ggSeyaPcuzHB/xzQO41K6xO/OO8K+Z3UIenJIeQZwhQgvBH66zHgiUBcTRNk3FqZ0X/RS9olXpXdj8+VAbsmTUoWRE9rSd (b) ↵ + n(t) absHeN6_4n=t"cgA+WIaRtNHeZv8PIca16g83sNxkl07ZiKb/BQS3NrE3YJ98vOrUp1L=E"h> AiAeAaB"=IUO8J3r3Q/KZ0kN3g1cPveNRI+g"=4E86LNHrpN6OrKWcvgaI9JnQY0IgEvAIN=8sSUN8B3Z3b/cZikH371BPAeARA+>""4=e6b_e13aRh3s" PtZiax4e+tea1lkYAcAZANB87AHIna9v8War<6LpEHltxeist ha1_ba 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 v187fiX53FFFa=oVN5woV88Y+aeRRlV0a8F73TXDff88vq1/ler>X+FNTixNFcGLG6E8zFiLFncz/HDkhPCAhPpjwQB<0l0tIxgtIVOeA8rwXoTFxGxFzGEZFct/DhChpwB00IgA5AZ3iFN0c5L86787FoLTnFzDH8kfP>tixetal/<=QqjYP8A5Pfk8HDzFnTLoF78768L5c0NFi3ZAOIgI00BwphChD/cFizEGGFxTFXrl1v8fX3FaVRe+8VwNo3i8PYjqQ=> Sl6tox5thsRae_iabeW4O"blEArAqARBv+i3JigcGb8ZaBWN1SP8oNFAAElIUYP3S9eKgPvahEaKtY2bvLxfxtx0gsX0HmQ7KE/7JE/Umvbp7KvUbLft0s0m7E7EUvpXvHhQxKt/xJv/xm2bg7E7E7m0s0tfLbYKEaPg9JFqAglEUGPRZbci3+BAAA>"=YSoXPi1rWva93SYP=e"l>EAFAPAvBs+o3AiUcSbXZEBmNYSX8JNqAgEEIGYR389vapvUW7r710i0YLb9YSKeEgaKP8axKEAI8SNBZNcib+B3AAA"=>SaYgb7mx2Xpv/H/hJQtvx zaBuZa19WHTtuvpVlHSVcurzrSCKekjBDY+T0cKRuvezerROadCcEkikueD+jcY0TucIRevRzXraOEdQcikBkCekcWIzXCQcClCIIpITWIFWTZaWhB/u7F3s740TVbGrvaKsqp9hd4B+p/0wzDg7wsPv+3u7s0tV4Gbv1Krqs99dpB4pa0+zwgzwDPs+uhZ/B7v3u7ss0tV4Gbv1Krqs99dpB4pa0+zwgzwDP+CcdaORrezevuRKTTOFkHRWcpzITlkIdVrCvccCYziQESCXarRIekecuBKe0u+cIrkcBeukikE0T+YDj+sPDwzgwz+0ap4BpdjD99sqrK1vbG4Vt0s7u3v7B/ZhuaWTTFHWpIlIVCcCzQSX kmEQXRgcBTzVRCAPVKqQmBnuRuOKt+YIr/imc6UrQVQqw3KF6ZmweAu/TwrYAzkDmAEXmmzAKekwXc8t/mLRtXjBcmFDJ9qlwsBnvQVvcxmmLX3tVJDAuSuSCF9DFVryTuem6KwQQUcirYtORnmqVARzBgXropwcFvTCDxUAlJtqwBvVcmL3VDuuC9csFnjlLQ8QkYzSEPk7rzuZmIKSQDUyi+YuOAnNqrAozpgwJcwvvTcCLDVxuUCAslnSQFY7PSzlIyDN+QAorzpAcqTNDYUolQzwku8vLEjCF8qxNUuZDSI7zQP7YlQ9nZsuC3uSVmLBcgvuX+BARnVrmORitprUcKQcwm6reTTkAzmDmkKLXj/ltAcy TTMYSNYZqOatDcyAf6zW2iqaXRNZqtn/Jw7ITsTC5ZR05UEApN2u5/+W3GteigAmraHZW04azfgQyTnxTEH+gKURCj/0XVJT5qNDXSqq2YzafYyMTR5Ep25+3tiArHW4zgynTHgUC/XZuZ/tW6G/etgwmWaIZN0saifCQcTZxaE0+OKURRjA0AVnVANAj0RURO+KEa0ZTxQCciaf0NsIaZmwWtegG6/t/WuZZ/CXgHUnyTz4gHYWSYMaDqfzyqX2qnN75J5ER25p3t+AriW4HgyYMrSpY5q3aTD5y5fiz72TqRXEN2q+ntJAznTHgUC/XZuZ/tW6G/etgwmWaIZN0saifCQc7TxaZ0+EKUORjR0AANNV OKbzIPKcedgxKSqdScIYbJnSCRqYmq7g8c96F7jLL9W3tzyJJj1IDkyRJ7JK3mzptaL9W59LFoLVj5zRP7I+vwBOKsjvBw7+IzPRV5jFLo5L9LW9pat3zm7KJyJRkD1jIRrjkyeRSxjikeIJ58ij9sZyr8m7G6HqqC9/LnYbE4iSKqJAKgnd/9LqvBK7s6ZA9KrJ84yi7Eq/CLm9H6qH6GCy9rqZL8ni/5biEkYjiyce6EqiJ4KJKKgAA6q7HmG7y8r9Zs8jiI5JiekSjxyRerkjjRIk1DyRJ7JK3mzptaL9W59LFoLVj5zRP7I+vwBOKd6S741krIjRRjkjDkRr7eKRmypxaj9S5kLeoiVJ55RI7i+jw8OG T5aAi=e=STS5/Aa=>=AlTa=taeixSi/tS>5<==<=lAt5xTt le6Wr6pHL8sab_1ahs tixetal"=IUO8va9nYIEAN8SNBZbciH7BAAA>"=IUO8J3r3Q/KZJ0kN3g1cPveNRI+g"=46eE86LeHapN6hrsWcvgab9inxY_IgEvAIN=8sS1N BeZ""4=eIaU_Oa8sJt3xrt3lQe/iK Zh01kbNs36sb_1ahs tixetal<="0ga+QINRKNre3vkPZc/13g3aIRtEUOIJ383QrKZ/kN0g14E=+"g3veNPc 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 oNFcN0F0c5L86787FoLTnFzDH8kfP5A8PYjqQ=>tixetaoNwV8+eRiVaF3XL588687>877cF/oDLeTxnVFezqDfHT8wkFfrPE5CAN8iPaY=jYq5Q8=FFo3NfwvVl8X+TeFRGVz BZibcBX7AAA=>"0kstyQ6qEbW8o5wii53zsxg7"=46esab_1ahs tixetal<AAAB7XicbZBV55rBcib7BXAAA"=>s0kQBZVbbc7iiXz76BhAaA5Ax>="s=1k se0tilela8/7V+YcGnzw5Mx8ca8iCs9vffXMilPVzh3iGC3RFaAd1pABq4AhIMXh4hBIRXMp4AAqcad1FACiR33GVljziPfsvX9filaCc889wxz5n7+GVY8l3>iixttae/l/aV8<7GYzn+9x5hRBvMV4lhlXiMwIj4fAspaq8cGA3dz1PaiAXCfF9RCc8 bBZXciA7B>AAk"=ys0qQtWE658biwoV553r5xzs"g76=4aes1b_sahi ttxe"=ks0yQtqE6W8b5woi55Vr53zsxg7"=46esab_1ahZBbVcBi0Xx78BAAsAlAo>q""=ikast0ty5Q5t5q6EQeAsBaXbc_Zs tixetal< vPpYP/vtPYlvqTfNNF6s9davOtsVLwgPYOds3NZFs2jjMHGD6MW2NsSo+BvF7/3v195Lh3bMXC1UjFc86RuR4BX0dvrCZY32dXsp/kKNBk8SeYNpY/3iKdwvMpxvElITZsna6s1gedtZXj1GWWPGXGdNwiB/fSNoJwMxEIZn61etX1WPXoJF8FR2RdBj0YvHCvYD2qXMptk2NtksSVYopCGGSwNxEMZnI1e6X1tPXWJFoFR8Rd2j0BvHYvYC2qDMpXk2ttkNSVsopY/BfNU7GjN4Kh3dY6NXe18+BXKu/cs1db35Z3rvwCW26MGnMHjPsNZD3jdFYtgxLtsqOvaY9dsFNoT1v1lIPwvk/ppXP2vYFCdvw0iBBR/Rf8pJoXYWVXSes6kZtENM2SFPp/vPlvTNs9aOsLgYd3ZsjMG6WCG+v7315hbX1jc6u4XdrZ3ds/KB8eNY3KNGU+vG7B3d1Y5Zh/beXK1Nj3cs6Ku84NX3dNrUvidNHC1qioZTtSr7CQrgBNQE1zlxyhlDRcNultrY1O01dUpMh5BSqRcYYKo1Uohd1Doutr+tK5V6vY5NClYwilZDL3La7cnsFlUDOilI3OlE5xDpK25iOdc1zSQYoUrOSE+sZSL+EZQLwEyQXwly2XflU2yfFUfH2Bl4XIylwsQqEBLVZX+BSXsPEQOxeKdx2yx0w5Cr7vh1LylQLKprZ+Z1itdlYr1UCwx1KixpQEPIXDBsXaVDBwqNs6ltIu4dB1HYeSiMI1OYEuxDpx2Eigd71T4qdIilZsZqpBLVlXLBhX7PCQnxBKFxyyU0f52rlvX1yVwQQKErL+Z1+tSlsrEUOewBnTqoQS7EgNhzxuDcOtYM1UR5S1YKDodturY56wNl3lDsacilDEIO2xp1ideswOlUtE1r+TrSK7QQVg1NvErz5x0hyDxcKuxtQYPOX1BUXMV5BSqRsYlKI14oBdHDyuFrBtn5C67YhNLllwLlpDZ3ZaicdsYl1DCV 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 8ldaVR7l9lsvsilipa>>hwlatixetal/p49hMPfcqpA4IMXh4MBhRtixetal/daiCl88cwx95nz+G7Y8V 0. The limiting case where γ = 0 (i.e. when the heating just balances the cooling) gives a steady state temperature of Tw when the current is √ ( [ ]) ṁCw Ilim = 1− −hcA exp . (C.11) R0α ṁCw Currents above this will heat the coils indefinitely (γ > 0) according to the model. This indefinitely heating results from ignoring radiative and thermal conduction to the air. Currents below Ilim will result in a steady temperature of Tf = β/ |γ|. As noted in the text, currents used in the experiment are far below Ilim. With the solution for T (t), the resistance of the wires as a function of time becomes ′ [ ( −| | )]R(t) = R0 1 +Rm 1− e γ t , (C.12) where R′0 is the chilled coil resistance R′0 = R0 [1− α(T0 − Tw)] (C.13) 217 Parameter Value Aluminum support volume 3.52× 10−5 m3 Copper wire volume 1.73× 10−5 m3 Water flow rate, ṁw [141] 13 3.43 litre / min = 0.217 kg / s Water chiller set temperature, Tw 15 ◦C Cooling channel Length, L 9.7× 10−2 m Cooling channel surface area, A 3.29× 10−2 m2 TABLE C.1. Water-cooled MOT coil parameters. Variable names reference equations C.2, C.3, and C.5. T0 is a reference temperature used in defining the “normal” resistance of the wires, R0, and Rm is a maximum change of resistance I2R0 ∣ ∣ ( [− ])∣∣ ṁCw hcA ∣ −1 R = ∣m | | = ∣α− 1− exp ∣ . (C.14)γ (mC) 2eff I R ṁCw As noted in the text, measuring the resistance of the anti-Helmholtz coils as a function of time is straightforward and provides a clear method to find a value for γ. C.2 Water Cooling Rate Derivation A closed channel through which water flows can be used to regulate the temperature of the bulk medium. The rate that heat flows from the bulk into the water and leaves the medium is calculated here, following the general formalism of [142]. Divide the length of the channel into a small segment dx, as shown in figure C.1. The mass of water, dm, that flows through this channel in time dt will absorb heat, 13We cannot locate a manual for our chiller, RTE100. The RTE101 has the same specifications as our chiller and the plumbing and circuit diagrams match the innards of our chiller, leading us to trust the RTE101 manual. 218 L W H dx FIGURE C.1. Water cooling channel dimensions dq, from the walls of the channel according to dq = q̇dt, (C.15) where q̇ is the rate of heat transfer. This rate is from conduction from the walls into the fluid, which follows q̇ = hc(Ts − Tf )dA, (C.16) where dA is the surface area of the fluid that is in contact with walls of the channel, hc is the conduction coefficient (discussed below in Section C.3), and Ts is the (fixed) temperature of the walls of the channel and Tf is the temperature of the fluid. If the perimeter of the volume of water has length P, then the surface area is just dA = Pdx. Assume the heat is absorbed uniformly throughout the fluid (there is no temperature gradient from the surface into the bulk of the liquid), so that the absorbed heat will warm the water by an amount dT as dq = dmCdTf , (C.17) 219 where C is the specific heat of the water. Combining equations C.15 through C.17 gives hc(Ts − dm Tf )Pdx = CdTf . (C.18) dt In this equation dm/dt is just the mass flow rate through the fluid, ṁ. The total change in the temperature from the input, Tin, to the output, Tout, is found by integrating along the total length of the channel, L. ∫ ∫ h P L Toutputc dTf dx = ṁC 0 T T − T[ ] input s f−hcPL Ts − Tout exp = − (C.19)ṁC Ts Tin Now, consider the total heat absorbed by the water in the channel, Q = mC(Tout − Tin). (C.20) This occurs at a rate [ − ]Ts Tout Q̇ = ṁC(Ts − Tin) 1− − . (C.21)Ts Tin Using the result of equation C.19 gives an equation for the rate that heat is absorbed into the water from the bulk of [ [− ]]hcA Q̇ = ṁC(Ts − Tin) 1− exp , (C.22) ṁC where A = PL is the total surface area of the water-flow channel and Ts is the temperature of the walls of the water-flow channel. 220 C.3 Conduction Coefficient The conduction coefficient, hc can be related to the Nusselt Number, Nu a ratio between the thermal conduction of heat from surface into fluid and the thermal convection of heat into into the fluid. This is, hcLc Nu = (C.23) k where Lc is a characteristic length of a flow and k is the thermal conductivity of the fluid. For a long rectangular channel with a width that is 4-times the height and a uniform temperature of the channel, the Nusselt number is 4.439 [142]. The characteristic length for a flow through a long tube is cross-sectional area 2HW Lc = 4 = (C.24) perimeter of cross-sectional area H +W where H and W are as shown in figure C.1. This gives a conduction coefficient of kNu(H +W ) hc = (C.25) HW Relating this to our experiment, for water, k = 0.6098 W / mK [143], and our channel has H = 1/8′′ and W = 1/2′′. These give the conduction coefficient hc = 5.3 × 102 W / K m2. 221 APPENDIX D BAYESIAN EVOLUTION DERIVATION Equation 4.4 gives the probability distribution for fluorescence rate (variable x) from a single atom, Fl1-at(x). With a known number of atoms in the MOT, the measurement of the fluorescence from n atoms in the MOT can be used to update information about the fluorescence rate average, R, and standard deviation, σR, from one atom with Bayes theorem. With a noisy measurement, y as in section 4.2.3, the noise in a measurement must have value ζ = y − (B + nr) , (D.1) where B is the (assumed constant) background fluorescence rate, n the number of atoms in a MOT, and r is the single-atom fluorescence rate. The probability for the noise to have this value is [ ] − √ 1 − (y − (B + nr)) 2 p (ζ = y B + nr) = exp , 2πσ2 2σ 2 ζ ζ which is similar to Equation 4.8 with multiple atoms in the MOT. Then, Bayes’ theorem just says that the single atom signal fluorescence evolves according to Fl1-at(r)p (ζ = y −B + nr) Fl1-at(r)→ ∫ ∞ Fl1-at(r)p (ζ = y −B + nr) dr −∞ 222 The normalization function is ∫ ∞ 1 Fl1-at(r)p (ζ = y −B + nr) dr = × −∞ 2πσRσzeta ∫ ∞ [ [ ]− ](r −R)2 − (y − (B +Nr))2 exp exp dr, 2 −∞ 2σR 2σ 2 ζ which integrates to [ ] 2 √ 1 − (y( ) exp ( − (B +NR))) . (D.2) 2 2 2 2π σ2N2 + σ2 2 σRN + σζR ζ Thus, the single-atom fluorescence probability evolves according to [ ] [ ] → √ 1 − (r −R) 2 − (y − (B + nr))2 Fl1-at(r) exp exp × 2 2 2σ2 2σ 2σ 2π R σζ R ζ (σ2 2 2 [ r n +σζ) ] − (y(− (B + nR))) 2 exp 2 σ2 n2 + σ2R ζ  [ ] 2 − − (Rσ 2 ζ+ynσ 2 R−Bnσ 2  R)r 2 21  n2σ +σ √ × R ζ = √ exp σ2 σ2  .σ2 σ2 2π R ζ 2 R ζ  2 2 2 n2σ2 +σ2 n σR+σζ R ζ This is, of course, just a Gaussian. Starting at i, and evolving to i+1 while measuring data point yi+1, the average and variance evolve as: R σ2i ζ+(yi−B)nσ 2 R R,ii+1 = n2σ2R,i+σ2ζ (D.3) σ2 2 σ2 = R σ i ζ R,i+1 n2 . σ2 2R,i+σζ It is good to note that if n = 0, then R 2 2i+1 = Ri and σR,i+1 = σR,i. This should be the case as with no atoms in the MOT, no information can be gained about the 223 fluorescence rate from a single atom. For this reason, when there are no atoms in the MOT, our algorithm instead updates the background fluorescence rate average and standard deviation as discussed in Section 4.2.6. 224 APPENDIX E GAUSSIAN SAMPLED OSCILLATION AMPLITUDE Photons are measured by the APD as both a background rate, β, which is Gaussian-distributed with average rate B and a fluorescence rate from the atom, α which is Gaussian distributed with variance σα and whose average oscillates in time as described in the text: α(t) = α0 [1 +mα cos(2πfαt)] . (6.22) These two random values are sampled together, so that their total rate f and rate variance σ2 are just the sum of the two, fl(t) = 〈Fl〉 [1 +  cos(2πfαt)] (E.1) σ2 = σ2 + σ2B α, with average total background fluorescence and (dimensionless) total fluorescence oscillation amplitude defined as 〈Fl〉 = B + α0, and (E.2)  = α0mα/ (B + α0) . For single-atom fluorescence, as shown in Figure 4.3, the fluorescence is relatively close to a Poisson distribution and can be written after time T as σ2(T ) = 〈fl(T )〉 (1 + η) , (E.3) 225 where η is a parameter that compares the variance to the mean rate. Because of the Gaussian assumption for the photon rates, as time progresses both the variance and the mean for the fluorescence distribution grows linearly. Thus, the parameter η is constant over all times. The variances as measured in Figure 4.3 should then hold for estimates of the oscillation parameter (m) for the atom. Photon arrivals between time t = 0 and t = T produce a pulse chain, ∑ p(t) = δ(t− ti) (E.4) i which has a power spectrum [132] 1 [ { } ] S(f) = 〈fl(T )〉+ 〈fl2(T )〉 − 〈fl(T )〉 〈ei2πf(tc−tc′ )〉 , (E.5) T where the brackets donate statistical averages and the exponentials result from Fourier transforms of the photon arrivals [144]. With a variance defined relative to the average rate, it is possible to write 〈fl2〉 = σ2 − 〈fl〉2 = 〈fl〉+ 〈fl〉η − 〈fl〉2, (E.6) which simplifies the power spectrum to 1 [ { } ] S(f) = 〈fl(T )〉+ 〈fl(T )〉2 + η〈fl(T )〉 〈ei2πf(tc−tc′ )〉 . (E.7) T 226 To calculate the statistical average of the exponential, integrate over both tc and tc′ with both exponentials normalized by average photon rate. This produces ∣∣∫ ∣∣T 2 〈 1ei2πf(tc−tc′ )〉 = ∣〈 〉 ∣ f(t)e −i2πftdt∣ , (E.8) f(T ) 2 ∣0 so that the spectrum becomes [ ∣ ∣ ] ∣∣∫ ∣∣∫ 2 T 1 T 2 ∣ f(t)e−i2πftdt∣ ∣ ∣ S(f) = 〈f(T )〉+ ∣ f(t)e− ∣ i2πft η 0dt∣∣ + 〈 〉 . (E.9)T 0 T f(T ) The first term is identical to the Poisson-distributed rate calculated by Matzner and Bar-Gad [132], while the second term corresponds to the Gaussian modification made here. The added noise from the Gaussian-distributed fluorescence then increases the spectral power over the Poisson-distributed signal. The average fluorescence is given by ∫ T 〈f(T )〉 = f(t)dt = 〈Fl〉T [1 +  sinc(2πfαT )] . (E.10) 0 In the limit of small  and at f = fα, the last term is ∣∣∫ ∣2∣ T ∣η fl(t)e−i2πftdt∣ [0 2 〈 〉 = η〈Fl〉× + sinc (πfαT ) +  sinc (πfαT ) +T fl(T ) 4 ] 2 sinc(2πfαT ) {1 + 4 sinc(πfαT ) + sinc (2πfαT )} .4 (E.11) All of the oscillating terms decay rapidly at high frequencies (or long times), so they can be dropped. Inserting the Poisson-distributed spectrum, gives a final form for the power spectrum at f = fα 〈Fl〉2T2 2〈Fl〉 S (f = fα) = 〈Fl〉+ + η . (E.12) 4 4 227 Similarly, far from the driving frequency, fα, the spectrum becomes just S(f →∞) = 〈Fl〉, (E.13) which is just the average measured fluorescence rate (from the background and an atom). Now, from the power spectrum from the APD, the oscillation amplitude can be measured as √ 4 [S (f = fα)− 〈Fl〉]  = 〈 〉 (E.14)Fl 2T + η〈Fl〉 Just setting η = 0 returns the Poisson-distributed result of Matzner and Bar-Gad [132] for the oscillation parameter mp. Comparing these results gives √ 〈Fl〉T  = 〈 〉 mp. (E.15)Fl T + η The quantity 〈Fl〉T is the total number of photons counted in time T without any oscillations. This value should be much larger than the added super-Poissonian noise, η. Using the simpler Poisson result of Matzner and Bar-Gad is justified. The atom’s fluorescence is responsible for the oscillations. Then writing the Poisson form of  (setting η = 0) in terms of the atomic oscillation amplitude gives √ 2 S (f = fα)− (B + α0) mα = . (E.16) α0 T 228 APPENDIX F PARAMETRIC RESONANCE DERIVATION A parametric resonantor is one where the value for the restoring spring constant oscillates [131]. This means that the resonant frequency also oscillates. . Taking the simplest equation for a damped, harmonic oscillator and allowing it to become parametric gives the differential equation for its position, z β κ z̈ + ż + z [1 + cos(ωt)] = 0 (F.1) m m where m is the mass of the oscillator, κ is the restoring force spring constant, β is the damping coefficient, and ω is the oscillation frequency of the parametric spring constant. This frequency is not necessarily the same as the constant oscillation √ frequency, ω0 = κ/m. Assume that the differential equation can be solved by an equation z(t) given by z(t) = a(t) cos(νt) + b(t) sin(νt) [ ] ż(t) = [ȧ(t) + νb(t)] cos(νt) + ḃ(t)− νa(t) sin(νt) (F.2) [ ] [ ] z̈(t) = ä(t) + 2νḃ(t)− ν2a(t) cos(νt) + b̈(t)− 2νȧ(t)− ν2b(t) sin(νt) for some frequency ν. Putting these equations into the differential equation F.1 gives [ ] 0 = ä+ 2νḃ− 2 βν a+ ȧ+ ω20νb+ ω20a+ ω20a cosωt cos(νt) + [ m ] b̈− β β2νȧ− ν2b+ ḃ− νa+ ω20b+ ω20b cosωt sin(νt).m m 229 Replacing the two frequencies with the natural frequency of the oscillator and a small deviation frequency ζ: ω → 2ω0 + ζ (F.3) ν → ω + 10 ζ,2 and noting that cos(ωt) cos(νt) = cos [(ω + ν)t] + cos [(ω − ν)t] = cos [3νt] + cos [νt] cos(ωt) sin(νt) = sin [(ω + ν)t]− sin [(ω − ν)t] = sin [3νt]− sin [νt] , the differential equation becomes [ ] β β 0 = ä+ 2νḃ− ν2a+ ȧ+ νb+ ω2 2 m m 0 a+ ω0a cos(νt) + [ ] b̈− β β2νȧ− ν2b+ ḃ− νa+ ω2 20b− ω0b sin(νt),m m after dropping the quickly rotating 3ν terms. Further, assume that ȧ ∼ ζa (F.4) ḃ ∼ ζb 230 Dropping terms that scale as the very small ζ2, the following changes are made to the differential equations: ä ≈ 0 b̈ ≈ 0 ν2 ≈ ω0(ω0 + ζ) νȧ ≈ ω0ȧ νḃ ≈ ω0ḃ. These changes give a final differential equation of the form [ ] − a βȧ βνb0 = 2ḃ ζa+ ω0 + + cos(νt) + [ 2 ω0m ω0m ] − b βνa βḃ2ȧ+ ζb+ ω0 + − sin(νt). 2 ω0m ω0m Under the assumption of small damping, β ∼ ζ, the differential equation can be simplified to [ ] a β 0 = 2ḃ− ζa+ ω0 + b cos(νt) + [ 2 m ] − b β2ȧ+ ζb+ ω0 + a sin(νt). 2 m This is solved only when the coefficients for both cos(νt) and sin(νt) vanish. Forcing both to vanish produces coupled differential equations for ȧ and ḃ that follow the 231 vector equation    ( )    a   β 1d ω + ζ a  − 2m 4 0 2   = dt ( )   . (F.5) b 1ω ζ β 4 0 − b 2 2m Assuming an eigenvalue solution, the equations of motion are a(t) = a e−λt0 b(t) = b e−λt0 with eigenvalue √ β 1 ζ2 λ = ∓ 2ω20 − . (F.6)2m 16 4 A non-parametric resonator will be damped to z = 0. This solution will behave similarly unless λ < 0. This exponential growth of the oscillator amplitude occurs if 2 ζ2 1 < 2ω20 − β . 4 m2 In terms of the frequencies, this parametric resonance is excited for frequencies √ √ 2ω − ω0 2 − 2 ω00 TH < ω < 2ω0 + 2 − 2TH (F.7)2 2 if the parametric strength is above a threshold strength 2 4β 2 TH = . (F.8)m2ω20 232 REFERENCES CITED [1] T. Hänsch and A. Schawlow, Optics Communications 13, 68 (1975). [2] V. S. Letokhov, V. G. Minogin, and B. D. Pavlik, Journal of Experimental and Theoretical Physics 45, 698 (1977). [3] S. Chu, L. Hollberg, J. E. Bjorkholm, A. Cable, and A. Ashkin, Phys. Rev. Lett. 55, 48 (1985). [4] E. L. Raab, M. Prentiss, A. Cable, S. Chu, and D. E. Pritchard, Phys. Rev. Lett. 59, 2631 (1987). [5] A. Höpe, D. Haubrich, G. Müller, W. G. Kaenders, and D. Meschede, Europhysics Letters 22, 669 (1993). [6] Z. Hu and H. J. Kimble, Optics Letters 19, 1888 (1994). [7] S. B. Hill and J. J. McClelland, Applied Physics Letters 82, 3128 (2003). [8] K. M. Fortier, S. Y. Kim, M. J. Gibbons, P. Ahmadi, and M. S. Chapman, Physical Review Letters 98 (2007). [9] M. J. Gibbons, C. D. Hamley, C.-Y. Shih, and M. S. Chapman, Phys. Rev. Lett. 106, 133002 (2011). [10] L. Förster, W. Alt, I. Dotsenko, M. Khudaverdyan, D. Meschede, Y. Miroshnychenko, S. Reick, and A. Rauschenbeutel, New Journal of Physics 8, 259 (2006). [11] Z. Zuo, M. Fukusen, Y. Tamaki, T. Watanabe, Y. Nakagawa, and K. Nakagawa, Opt. Express 17, 22898 (2009). [12] D. Schrader, I. Dotsenko, M. Khudaverdyan, Y. Miroshnychenko, A. Rauschenbeutel, and D. Meschede, Phys. Rev. Lett. 93, 150501 (2004). [13] Y. Miroshnychenko, D. Schrader, S. Kuhr, W. Alt, I. Dotsenko, M. Khudaverdyan, A. Rauschenbeutel, and D. Meschede, Opt. Express 11, 3498 (2003). [14] D. B. Hume, I. Stroescu, M. Joos, W. Muessel, H. Strobel, and M. K. Oberthaler, Phys. Rev. Lett. 111, 253001 (2013). [15] B. Ueberholz, S. Kuhr, D. Frese, V. Gomer, and D. Meschede, Journal of Physics B 35 (2002). 233 [16] N. Spethmann, F. Kindermann, S. John, C. Weber, D. Meschede, and A. Widera, Phys. Rev. Lett. 109, 235301 (2012). [17] F. Schmidt, D. Mayer, T. Lausch, D. Adam, Q. Bouton, M. Hohmann, F. Kindermann, J. Koch, J. Nettersheim, and A. Widera, Physica Status Solidi B Early View, 1800710. [18] C. Y. Chen, Y. M. Li, K. Bailey, T. P. O’Connor, L. Young, and Z.-T. Lu, Science 286 (1999). [19] J. Welte, F. Ritterbusch, I. Steinke, M. Henrich, W. Aeschbach-Hertig, and M. K. Oberthaler, New Journal of Physics 12, 065031 (2010). [20] F. Ruschewitz, D. Bettermann, J. L. Peng, and W. Ertmer, Europhysics Letters 34, 651 (1996). [21] Y. Choi, S. Yoon, S. Kang, W. Kim, J.-H. Lee, and K. An, Phys. Rev. A 76, 013402 (2007). [22] J. Weiner, V. S. Bagnato, S. Zilio, and P. S. Julienne, Rev. Mod. Phys. 71, 1 (1999). [23] V. Gomer, F. Strauch, B. Ueberholz, S. Knappe, and D. Meschede, Phys. Rev. A 58, R1657 (1998). [24] V. Gomer, B. Ueberholz, S. Knappe, F. Strauch, D. Frese, and D. Meschede, Applied Physics B 67, 689 (1998). [25] H. J. Metcalf and P. van der Straten, Laser Cooling and Trapping (Springer-Verlag, New York, 1999). [26] P. Kohns, P. Buch, W. Sptitz, C. Csambal, and W. Ertmer, Europhysics Letters 22, 517 (1993). [27] A. M. Steane and C. J. Foot, Europhysics Letters 14, 231 (1991). [28] C. Tuchendler, A. M. Lance, A. Browaeys, Y. R. P. Sortais, and P. Grangier, Phys. Rev. A 78, 033425 (2008). [29] A. Fuhrmanek, A. M. Lance, C. Tuchendler, P. Grangier, Y. R. P. Sortais, and A. Browaeys, New Journal of Physics 12, 053028 (2010). [30] P. D. Lett, R. N. Watts, C. I. Westbrook, W. D. Phillips, P. L. Gould, and H. J. Metcalf, Phys. Rev. Lett. 61, 169 (1988). [31] A. Kastberg, W. D. Phillips, S. L. Rolston, R. J. C. Spreeuw, and P. S. Jessen, Phys. Rev. Lett. 74, 1542 (1995). 234 [32] W. Alt, D. Schrader, S. Kuhr, M. Müller, V. Gomer, and D. Meschede, Phys. Rev. A 67, 033403 (2003). [33] W. Ketterle and N. V. Druten, Advances In Atomic, Molecular, and Optical Physics 37, 181 (1996). [34] R. Grimm, M. Weidemuller, and Y. B. Ovchinnikov, Advances In Atomic, Molecular, and Optical Physics 42, 95 (2000). [35] M. Weber, J. Volz, K. Saucke, C. Kurtsiefer, and H. Weinfurter, Phys. Rev. A 73, 043406 (2006). [36] F. M. V. Gheorghe and G. Werth, Charged Particle Traps (Springer-Verlag, Heidelberg, 2005). [37] J. C. Bergquist, W. M. Itano, and D. J. Wineland, Phys. Rev. A 36, 428 (1987). [38] A. D. Ludlow, M. M. Boyd, T. Zelevinsky, S. M. Foreman, S. Blatt, M. Notcutt, T. Ido, and J. Ye, Phys. Rev. Lett. 96, 033003 (2006). [39] J. B. Yi Deng and N. R. Forde, Journal of Optics A: Pure and Applied Optics 9, S256 (2007). [40] S. Friebel, C. D’Andrea, J. Walz, M. Weitz, and T. W. Hänsch, Phys. Rev. A 57, R20 (1998). [41] A. Görlitz, M. Weidemüller, T. W. Hänsch, and A. Hemmerich, Phys. Rev. Lett. 78, 2096 (1997). [42] R. Loudon, The Quantum Theory of Light, 3rd ed. (Oxford University Press, New York, 2008). [43] M. A. Nielsen and I. L. Chuang, Quantum Computation and Quantum Information, ninth ed. (Cambridge University Press, 2007). [44] W. K. Wootters, Phys. Rev. Lett. 80, 2245 (1998). [45] R. Wagner and J. P. Clemens, Phys. Rev. A 79, 042322 (2009). [46] R. Wagner and J. P. Clemens, J. Opt. Soc. Am. B 27, A73 (2010). [47] H.J.Carmichael and K. Kim, Optics Communications 179, 417 (2000). [48] D. A. Steck, Quantum and atom optics (2017), [Online; accessed February, 6 2018. http://steck.us/teaching]. [49] R. Shankar, Principles of Quantum Mechanis, 2nd ed. (Springer, New York, 1994). 235 [50] Y. Kaluzny, P. Goy, M. Gross, J. M. Raimond, and S. Haroche, Phys. Rev. Lett. 51, 1175 (1983). [51] T. H. Stievater, X. Li, D. G. Steel, D. Gammon, D. S. Katzer, D. Park, C. Piermarocchi, and L. J. Sham, Phys. Rev. Lett. 87, 133603 (2001). [52] E. Cook, Laser Cooling and Trapping of Neutral Strontium for Spectroscopic Measurements of Casimir-Polder Potentials, Ph.D. thesis, University of Oregon (2017). [53] J. B. Mackrory, T. Bhattacharya, and D. A. Steck, Phys. Rev. A 94, 042508 (2016). [54] P. W. Milonni, The Quantum Vacuum (Academic Press, Inc., San Diego, 1994). [55] G. K. Campbell and W. D. Phillips, Philosophical Transactions of the Royal Society A 369, 4078 (2011). [56] G. Santarelli, P. Laurent, P. Lemonde, A. Clairon, A. G. Mann, S. Chang, A. N. Luiten, and C. Salomon, Phys. Rev. Lett. 82, 4619 (1999). [57] O. Arcizet, P. F. Cohadon, T. Briant, M. Pinard, and A. Heidmann, Nature 444, 71 (2006). [58] S. Gigan, H. R. Böhm, M. Paternostro, F. Blaser, G. Langer, J. B. Hertzberg, K. C. Schwab, D. Bäuerle, M. Aspelmeyer, and A. Zeilinger, Nature 444, 67 (2006). [59] J. A. E. A. (JAXA), Small solar power sail demonstrator ’ikaros’ confirmation of photon acceleration (2010), [Press Release] Last accessed Aug. 2018. Originally available at http://global.jaxa.jp/press/2010/07/20100709 ikaros e.html. [60] Y. Tsuda, O. Mori, R. Funase, H. Sawada, T. Yamamoto, T. Saiki, T. Endo, K. Yonekura, H. Hoshino, and J. Kawaguchi, Acta Astronautica 82, 183 (2013). [61] P. W. Milonni and J. H. Eberly, Lasers (Wiley-Interscience., New York, 1988). [62] D. A. Steck, Rubidium 87 d line data (2001), [Online; accessed December, 2 2017. http://steck.us/alkalidata]. [63] D. M. Brink and C. V. Sukumar, Phys. Rev. A 74, 035401 (2006). [64] J. W. Jun, S. Chang, T. Y. Kwon, H. S. Lee, and V. G. Minogin, Phys. Rev. A 60, 3960 (1999). [65] A. L. Migdall, J. V. Prodan, W. D. Phillips, T. H. Bergeman, and H. J. Metcalf, Phys. Rev. Lett. 54, 2596 (1985). 236 [66] W. Petrich, M. H. Anderson, J. R. Ensher, and E. A. Cornell, Phys. Rev. Lett. 74, 3352 (1995). [67] M. H. Anderson, J. R. Ensher, M. R. Matthews, C. E. Wieman, and E. A. Cornell, Science 269, 198 (1995). [68] W. W. Erickson and D. S. Steck, In preparation (2019). [69] T. Bergeman, G. Erez, and H. J. Metcalf, Phys. Rev. A 35, 1535 (1987). [70] Y. Shevy, D. S. Weiss, P. J. Ungar, and S. Chu, Phys. Rev. Lett. 62, 1118 (1989). [71] J. Dalibard and C. Cohen-Tannoudji, J. Opt. Soc. Am. B 6, 2023 (1989). [72] S. Chang, T. Yong Kwon, H. Seong Lee, and V. Minogin, Phys. Rev. A 60, 2308 (1999). [73] E. A. Schoene, Cold Atom Control with an Optical One-way Barrier, Ph.D. thesis, University of Oregon (2010). [74] J. Thorn, Dissipative Control and Imaging of Cold Atoms, Ph.D. thesis, University of Oregon (2012). [75] T. Li, Manipulation of Cold Atoms using an Optical One-way Barrier, Ph.D. thesis, University of Oregon (2008). [76] M. Fuchs, Development of a High power Stabilized Diode Laser System, Master’s thesis, University of Oregon (2006). [77] D. Haubrich, H. Schadwinkel, F. Strauch, B. Ueberholz, R. Wynands, and D. Meschede, Europhysics Letters 34, 663 (1996). [78] M. Meucci, E. Mariotti, P. Bicchi, C. Marinelli, and L. Moi, Europhysics Letters 25, 639 (1994). [79] A. Gozzini, F. Mango, J. H. Xu, G. Alzetta, F. Maccarrone, and R. A. Bernheim, Il Nuovo Cimento D 15, 709 (1993). [80] C. Marinelli, A. Burchianti, A. Bogi, F. Della Valle, G. Bevilacqua, E. Mariotti, S. Veronesi, and L. Moi, The European Physical Journal D - Atomic, Molecular, Optical and Plasma Physics 37, 319 (2006). [81] C. Klempt, T. van Zoest, T. Henninger, O. Topic, E. Rasel, W. Ertmer, and J. Arlt, Phys. Rev. A 73, 013410 (2006). [82] J. H. Xu, A. Gozzini, F. Mango, G. Alzetta, and R. A. Bernheim, Phys. Rev. A 54, 3146 (1996). 237 [83] B. P. Anderson and M. A. Kasevich, Phys. Rev. A 63, 023404 (2001). [84] M. H. T. Extavour, L. J. LeBlanc, T. Schumm, B. Cieslak, S. Myrskog, A. Stummer, S. Aubin, and J. H. Thywissen, AIP Conference Proceedings 869, 241 (2006). [85] G. Telles, T. Ishikawa, M. Gibbs, and C. Raman, Phys. Rev. A 81, 032710 (2010). [86] S. N. Atutov, R. Calabrese, V. Guidi, B. Mai, A. G. Rudavets, E. Scansani, L. Tomassetti, V. Biancalana, A. Burchianti, C. Marinelli, E. Mariotti, L. Moi, and S. Veronesi, Phys. Rev. A 67, 053401 (2003). [87] E. Mimoun, L. D. Sarlo, D. Jacob, J. Dalibard, and F. Gerbier, Phys. Rev. A 81, 023631 (2010). [88] K. M. Birnbaum, Ultra high vacuum chambers (2005), [Online; Last access unknown. Originally available at http://www.its.caltech.edu/qoptics/Vacuum/UHV chambers.pdf]. [89] Z. T. Lu, K. L. Corwin, M. J. Renn, M. H. Anderson, E. A. Cornell, and C. E. Wieman, Phys. Rev. Lett. 77, 3331 (1996). [90] C. Y. Park, M. S. Jun, and D. Cho, J. Opt. Soc. Am. B 16, 994 (1999). [91] C. D. Wallace, T. P. Dinneen, K. Y. N. Tan, A. Kumarakrishnan, P. L. Gould, and J. Javanainen, J. Opt. Soc. Am. B 11, 703 (1994). [92] A. M. Steane, M. Chowdhury, and C. J. Foot, J. Opt. Soc. Am. B 9, 2142 (1992). [93] P. G. Pappas, M. M. Burns, D. D. Hinshelwood, M. S. Feld, and D. E. Murnick, Phys. Rev. A 21, 1955 (1980). [94] P. E. Gaskell, J. J. Thorn, S. Alba, and D. A. Steck, Review of Scientific Instruments 80, 115103 (2009), https://doi.org/10.1063/1.3250825 . [95] J. J. Thorn, E. A. Schoene, T. Li, and D. A. Steck, Phys. Rev. Lett. 100, 240407 (2008). [96] E. A. Schoene, J. J. Thorn, and D. A. Steck, Phys. Rev. A 82, 023419 (2010). [97] J. D. Jackson, Classical Electrodynamics, 3rd ed. (John Wiley & Sons, Inc., New York, 1999). [98] J. H. Moore, C. C. Davis, M. A. Coplan, and S. C. Greer, Building Scientific Apparatus, 4th ed. (Cambridge University Press, New York, 2009). 238 [99] P. C. D. Hobbs, Building Electro-optical Systems, 2nd ed. (John Wiley & Sons, Inc., Hoboken, New Jersey, 2009). [100] K. Inc., Ate full rack power supply operator’s manual (2010). [101] E. H. Hall, American Journal of Mathematics 2, 287 (1879). [102] M. Briel, Design and Characterization of a Custom Aspheric Lens System for Single Atom Imaging, Master’s thesis, University of Oregon (2012). [103] C.-S. Chuu, Direct Study of Quantum Statistics in a Degenerate Bose Gas, Ph.D. thesis, The University of Texas at Austin (2006). [104] J. L. Hanssen, Controlling Atomic Motion: From Single Particle Classical Mechanics to Many Body Quantum Dynamics, Ph.D. thesis, The University of Texas at Austin (2004). [105] W. Alt, Optik - International Journal for Light and Electron Optics 113, 142 (2002). [106] W. Alt, Optical control of single neutral atoms, Ph.D. thesis, University of Bonn (2004). [107] Altera, De2 development and education board user manual (2012). [108] D. Branning, S. Bhandari, and M. Beck, American Journal of Physics 77, 667 (2009), https://doi.org/10.1119/1.3116803 . [109] M. Beck, Coincidence counting units (ccus) (2014), [Online; accessed December 2, 2017. http://people.whitman.edu/ beckmk/QM/circuit/circuit.html]. [110] S. Kuhr, A controlled quantum system of individual neutral atoms, Ph.D. thesis, University of Bonn (2004). [111] S. Yoon, Y. Choi, S. Park, J. Kim, J.-H. Lee, and K. An, Applied Physics Letters 88, 211104 (2006). [112] W. Ketterle, K. B. Davis, M. A. Joffe, A. Martin, and D. E. Pritchard, Phys. Rev. Lett. 70, 2253 (1993). [113] G. Labeyrie, F. Michaud, and R. Kaiser, Phys. Rev. Lett. 96, 023003 (2006). [114] T. Walker, D. Sesko, and C. Wieman, Phys. Rev. Lett. 64, 408 (1990). [115] N. Schlosser, G. Reymond, I. Protsenko, and P. Grangier, Nature 411, 1024 (2001). [116] S. Yoon, Y. Choi, S. Park, W. Ji, J.-H. Lee, and K. An, Journal of Physics: Conference Series 80, 012046 (2007). 239 [117] L. Mandel, Opt. Lett. 4, 205 (1979). [118] R. Short and L. Mandel, Phys. Rev. Lett. 51, 384 (1983). [119] T. W. Hodapp, G. W. Greenlees, M. A. Finn, and D. A. Lewis, Phys. Rev. A 41, 2698 (1990). [120] R. J. Cook, Optics Communications 35, 347 (1980). [121] Z. Hu, Quantum Optics with Cold Atoms – Nonlinear Spectroscopy and Road towards Single-Atom Trap, Ph.D. thesis, Californian Institute of Technology (1995). [122] O. Jacobs, Introduction to control theory (Oxford University Press, London, 1974). [123] T. Bhattacharya, S. Habib, and K. Jacobs, Physical Review Letters 85 (2000). [124] G. Milburn, Quantum Semiclassical Optics 8 (1996). [125] K. Jacobs and D. A. Steck, Contemporary Physics 47 (2006). [126] M. E. Rose, Elementary theory of angular momentum (Wiley, New York, 1957). [127] J. H. Wilkinson and C. Reinsch, Handbook for Automatic Computation : Volume II: Linear Algebra, 1st ed. (Springer-Verlag, Berlin Heidelberg, 1971). [128] L. Team, Lapack-linear algebra package (2017), [Online; Last access Aug. 2018. Available at http://www.netlib.org/lapack/]. [129] E. Raab, Trapping Sodium with Light, Ph.D. thesis, Massachusetts Institute of Technology (1988). [130] S. Marksteiner, K. Ellinger, and P. Zoller, Phys. Rev. A 53, 3409 (1996). [131] L. Landua and E. Lifshitz, Mechanics, 2nd ed. (Peramon Press, Oxford, 1969). [132] M. A. and B.-G. I., PLoS Computational Biology 11, e1004252 (2015). [133] M. Hohmann, F. Kindermann, T. Lausch, D. Mayer, F. Schmidt, and A. Widera, Phys. Rev. A 93, 043607 (2016). [134] W. W. Erickson, In Preparation, Ph.D. thesis, University of Oregon (2020). [135] L. Pedersen and H. Flyvbjerg, Phys. Rev. Lett. 98, 189801 (2007). [136] K. Kim, H.-R. Noh, Y.-H. Yeon, and W. Jhe, Phys. Rev. A 68, 031403 (2003). [137] K. Kim, H.-R. Noh, H.-J. Ha, and W. Jhe, Phys. Rev. A 69, 033406 (2004). 240 [138] K. Kim, H.-R. Noh, and W. Jhe, Optics Communications 236, 349 (2004). [139] G. B. Arfken, H. J. Webber, and F. E. Harris, Mathematica Methods for Physicists, seventh ed. (Academic Press, Waltham, MA, 2013). [140] R. E. Berg. and D. G. Stork, The Physics of Sound, 3rd ed. (Pearson Education, San Francisco, 2005). [141] T. NESLAB, Rte-101 bath circulator - installation, operation and service manual (2000). [142] F. Kreith and M. S. Bohn, Principles of Heat Transfer (Harper and Row, New York, 1986). [143] C. N. de Castro, S. Li, A. Nagashime, R. Trengove, and W. Wakeham, Journal of Physical and Chemical Reference Data 15, 1073 (1986). [144] Y. Pinhasi and Y. Lurie, Phys. Rev. E 65, 026501 (2002) 241