ABSTRACT Title of dissertation: EXPERIMENTS IN TWO-MODE CAVITY QED David Glenn Norris, Doctor of Philosophy, 2011 Dissertation directed by: Professor Luis Orozco Department of Physics Cavity quantum electrodynamics (QED) allows the study of light-matter in- teractions at the most basic level, through precise identi cation of the coherent and incoherent (dissipative) parts of the system evolution. We present measurements of light from a cavity QED system consisting of a high- nesse optical resonator coupled to a beam of cold Rb atoms. The novelty of the design lies in the interplay of two degenerate and orthogonal polarization modes. One mode (driven) behaves as the canonical cavity mode of the Jaynes-Cummings Hamiltonian, coherently exciting the atoms with a modest coupling strength; the other mode (undriven) collects a small fraction of spontaneously emitted light and provides a probe of the dissipative processes. We rst demonstrate the ability to detect individual atoms passing through the cavity modes in real time by coincidence detection of photons from the un- driven mode. Calculation of statistics and correlation functions from the complete photon detection record allows the determination of detection probabilities and the reconstruction of atomic trajectories. We next present evidence of quantum coher- ence that is created, modi ed, and measured in the excitation-spontaneous emission cycle. The coherence appears as a long-lived quantum beat at the ground-state Larmor frequency, visible in the intensity autocorrelation function of the undriven mode. Quantum jumps of the atomic state, occurring in between the detections of photons from the cavity, result in substantial changes in the frequency and spectral width of the beats. We present the results of a full quantum Monte Carlo calculation in order to quantitatively explain the measurements. EXPERIMENTS IN TWO-MODE CAVITY QED by David Glenn Norris Dissertation submitted to the Faculty of the Graduate School of the University of Maryland, College Park in partial ful llment of the requirements for the degree of Doctor of Philosophy 2011 Advisory Committee: Professor Luis A. Orozco, Chair/Advisor Professor Howard J. Carmichael Professor Mario Dagenais Professor William D. Phillips Professor Steven L. Rolston c Copyright by David Glenn Norris 2011 Acknowledgments The work presented in this thesis is the product of six years of collaborative e ort. The largest debt of gratitude is owed to my advisor and mentor, Professor Luis Orozco, who has most greatly shaped my development as a physicist. I know of no one more committed to science and to inculcating a respect for science in his students. Another great deal is owed to Professor Howard Carmichael, whose year with us proved crucial to understanding the physics involved in this work. The instruction happened chie y over co ee, and I am privileged to count such a brilliant person as teacher and friend. Many others have contributed to this project over the years, including current and former members of the UMD cavity QED group (Michael Scholten, Basudev Roy, Rebecca Olson Knell, Matt Terraciano, and Andres Cimmarusti, in whose ca- pable hands I leave the apparatus); our former postdoctoral researcher, Jietai Jing; our undergraduate researchers, Josh Crawford and Eric Cahoon; and a host of vis- iting student researchers. In addition, we have bene ted from theory collaboration with Pablo Barberis in Mexico, and Perry Rice and James Clemens in Ohio. Finally, I expressly acknowledge the contributions of the National Science Foundation, which has supported me monetarily for the past six years through a Graduate Research Fellowship, grants to the cavity QED program, and establish- ment of the Physics Frontier Center at the Joint Quantum Institute. Their generous funding made this research possible. ii Table of Contents List of Figures iv List of Abbreviations v 1 Two-Mode Cavity Quantum Electrodynamics 1 1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 1.2 Theoretical models . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2 1.2.1 The Jaynes-Cummings model . . . . . . . . . . . . . . . . . . 2 1.2.2 JC model versus reality . . . . . . . . . . . . . . . . . . . . . . 4 1.2.3 Two modes and three levels . . . . . . . . . . . . . . . . . . . 7 1.3 Apparatus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9 1.3.1 Physics cavity . . . . . . . . . . . . . . . . . . . . . . . . . . . 9 1.3.2 Atomic source . . . . . . . . . . . . . . . . . . . . . . . . . . . 11 1.3.3 Laser system . . . . . . . . . . . . . . . . . . . . . . . . . . . 15 1.3.4 Detection system . . . . . . . . . . . . . . . . . . . . . . . . . 18 2 Detecting Single Atoms 20 2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20 2.2 Detection scheme . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22 2.3 Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25 2.4 Preliminary data analysis . . . . . . . . . . . . . . . . . . . . . . . . 28 2.4.1 Faraday rotation . . . . . . . . . . . . . . . . . . . . . . . . . 28 2.4.2 Photons per atom . . . . . . . . . . . . . . . . . . . . . . . . . 29 2.5 Analysis and optimization of detection . . . . . . . . . . . . . . . . . 33 2.6 Intensity autocorrelation . . . . . . . . . . . . . . . . . . . . . . . . . 35 3 Ground-State Quantum Beats 43 3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43 3.2 Theoretical model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44 3.3 Measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52 4 Backward Light Shifts and Quantum Jumps 61 4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61 4.2 Experimental system . . . . . . . . . . . . . . . . . . . . . . . . . . . 62 4.3 Theoretical model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66 4.4 Data analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69 4.5 Comparisons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77 5 Conclusions and Future Directions 79 Bibliography 82 iii List of Figures 1.1 The optical cavity QED system . . . . . . . . . . . . . . . . . . . . . 2 1.2 85Rb level diagram and simpli ed three-level model . . . . . . . . . . 5 1.3 Vacuum chamber containing optical cavity and LVIS apparatus . . . 10 1.4 E ects of optical pumping beam . . . . . . . . . . . . . . . . . . . . . 13 1.5 Schematic of laser system . . . . . . . . . . . . . . . . . . . . . . . . 15 2.1 Schematic of coincidence measurement apparatus . . . . . . . . . . . 22 2.2 Diagram of atomic levels used for detection . . . . . . . . . . . . . . . 23 2.3 Measured photon statistics and linear ts . . . . . . . . . . . . . . . . 31 2.4 Counts detected from a single atom . . . . . . . . . . . . . . . . . . . 32 2.5 Signal-to-background ratios . . . . . . . . . . . . . . . . . . . . . . . 34 2.6 g(2)( ) vs. Faraday rotation . . . . . . . . . . . . . . . . . . . . . . . 36 2.7 g(2)( ) and t . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38 2.8 g(2)( ) vs. drive . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39 2.9 Atom number and count rates . . . . . . . . . . . . . . . . . . . . . . 41 3.1 Experimental system for measuring quantum beats . . . . . . . . . . 45 3.2 Simpli ed model of quantum eraser process . . . . . . . . . . . . . . . 46 3.3 g(2)( ) vs. atom number . . . . . . . . . . . . . . . . . . . . . . . . . 51 3.4 g(2)( ) and power spectrum . . . . . . . . . . . . . . . . . . . . . . . 53 3.5 g(2)( ) vs. magnetic eld . . . . . . . . . . . . . . . . . . . . . . . . . 54 3.6 g(2)( ) vs. drive . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55 3.7 g(2)( ) vs. waveplate mixing angle (composite) . . . . . . . . . . . . . 56 3.8 g(2)( ) vs. waveplate mixing angle (series) . . . . . . . . . . . . . . . 57 3.9 Calculated g(2)( ) and constituent terms . . . . . . . . . . . . . . . . 59 4.1 Four-level atom . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63 4.2 Schematic of apparatus for measuring g(2)( ) . . . . . . . . . . . . . . 63 4.3 Frequency shifts in calculated g(2)( ) . . . . . . . . . . . . . . . . . . 65 4.4 Frequency shifts in measured g(2)( ) . . . . . . . . . . . . . . . . . . . 65 4.5 Frequency of beat resonance vs. magnetic eld . . . . . . . . . . . . . 70 4.6 Frequency of beat resonance vs. drive . . . . . . . . . . . . . . . . . . 71 4.7 FWHM of beat resonance vs. drive . . . . . . . . . . . . . . . . . . . 73 4.8 Sample peaks from FFT amplitude spectrum . . . . . . . . . . . . . . 74 4.9 Frequency and FWHM vs. drive detuning . . . . . . . . . . . . . . . 76 iv List of Abbreviations AM amplitude modulation AOM acousto-optic modulator APD avalanche photodiode DAVLL dichroic absorption vapor laser lock DC direct current EOM electro-optic modulator FM frequency modulation FSR free spectral range FWHM full width at half maximum HWP half-wave plate JC Jaynes-Cummings LVIS low-velocity intense source MOT magneto-optical trap NIM nuclear instrumentation module PBS polarizing beam splitter PDH Pound-Drever-Hall PID proportional-integral-di erential PMT photomultiplier tube PZT piezo-electric transducer QED quantum electrodynamics RF radio frequency TEM transverse electromagnetic TTL transistor-transistor logic v Chapter 1 Two-Mode Cavity Quantum Electrodynamics 1.1 Introduction A cavity quantum electrodynamic (QED) system consists of an atom or other quantum object coupled to a small number of modes of a cavity [1]. Such a system provides a simple structure in which to investigate fundamental aspects of matter- light interactions, and may also be used for applications in quantum information science [2, 3, 4, 5]. Several cavity QED geometries operating over a wide range of the electromagnetic spectrum have now been demonstrated experimentally (see for example [1, 6, 7, 8].) This thesis presents a collection of experiments performed on a particular cav- ity QED system, a two-mode optical cavity and cold atomic source rst described in [9] and [10]. We begin with a theoretical overview of the system and the mea- surement goals. As the basic setup is common to all of the experiments discussed in this thesis, we give a thorough account of the technical details in this chapter, with additional details for each speci c measurement in the subsequent chapters. 1 g ? ? ? drive Figure 1.1: The optical cavity QED system. g: atom-photon dipole coupling frequency; : free-space atomic decay rate; : cavity eld mode decay rate; Edrive: coherent driving eld amplitude. 1.2 Theoretical models 1.2.1 The Jaynes-Cummings model The Jaynes-Cummings (JC) model provides the simplest realization of quan- tized light-atom interactions [11]. The model considers a single two-level atom, a saturable absorber, coupled to a mode of the electromagnetic eld, a harmonic os- cillator of equally-spaced energy levels; the mode is typically de ned by the mirror boundaries in an optical cavity (see Fig. 1.1.) Coherent electric-dipole coupling between the atom and a single photon (an excitation quantum of the mode) takes place at the Rabi frequency g = E= h = h s h! c 2 0V; (1.1) where is the atomic transition dipole moment and E is the amplitude of the electric eld of a single photon, with mode volume V and frequency !c. The system 2 Hamiltonian without dissipation is HJC = h!a z + h!caya i hg( +a ay ): (1.2) Here !a is the natural atomic transition frequency, a is the annihilation operator for the cavity eld, and ;z are the Pauli spin operators for a two-level atom. The evolution in this case is simply a continuous, reversible exchange of energy between the atom and the cavity mode. Dissipation may be added by considering separate reservoirs coupled to the cavity mode and atomic radiation pattern. Once traced over, non-reversible decay into these reservoirs brings lifetimes 1= and 1= for the cavity eld and atomic ex- cited state population, respectively [12]. The e ect of a single photon on the atomic state is characterized by the saturation photon number, nsat 2=(8g2), which may reach values much less than one for su ciently strong coupling. The e ect of an atom on the cavity eld is similarly characterized by the single-atom cooperativity parameter, C1 g2=( ), which expresses the ratio of coherent coupling to dissipa- tion rates. For the case of N atoms coupled identically to the mode, constructive interference enhances the emissions into the cavity mode, and the cooperativity parameter increases to C = NC1 [13]. Driving the cavity on resonance with a laser eld of amplitude Edrive (outside the cavity) causes the build-up of a coherent eld inside the cavity, the amplitude of which depends on C. The steady-state behavior of the system is given by the state equation, rst introduced in the context of optical bistability [13]: Y = X(1 + 2C=(1 +X))2; (1.3) 3 where Y is the intracavity photon number without atoms, and X is the number with atoms (both are normalized to nsat.) This form is exact for a plane-wave mode in a traveling-wave (ring) cavity, and it exhibits bistable behavior for large values of C and X. In the limit of weak driving (X 1), the equation takes on the simpler form: Y = X(1 + 2C)2: (1.4) This equation shows that X is smaller than Y by the factor 1=(1 + 2C)2, primarily due to the drive light radiated into free space by the atoms. 1.2.2 JC model versus reality Experimental realization of the JC Hamiltonian entails a number of approxi- mations. True two-level atomic systems are rare. Even the alkalis, a staple of atomic physics experiments because of their relatively simple spectra, exhibit complicated magnetic hyper ne structure and are susceptible to optical pumping. An example is the 85Rb D2-line F = 3!F0 = 4 manifold shown in gure Fig. 1.2a. An e ective two-level atom in such a level structure can be realized experimentally by optically pumping to the F = 3;m = +3( 3) ground state and using a + ( ) circularly polarized cavity mode to drive the F = 3;m = +3( 3) !F0 = 4;m0 = +4( 4) cycling transition, provided that light polarization and magnetic elds are carefully controlled. Real cavity mode structure is likewise complicated. In a standing-wave cavity of length l and in the transverse electromagnetic (TEM) basis, longitudinal modes 4 1 3 2 g G? ? 5S 1/2 ,F = 3, m = 5P 3/2 ,F' = 4, m' = 0 0 1-1 -1-4 -3 -2 -2-3 2 3 1 2 3 4 85 Rb a) b) Figure 1.2: a) Diagram of 85Rb levels used in the experiments. b) Simpli- ed three-level atomic model with coupling constants for the two cavity modes. are separated by a free spectral range, FSR = c=2l, with each of these split into multiple transverse modes. The requirement of a single spatial mode in the JC model means that all other modes must be separated from the mode of interest by several cavity linewidths. In addition, each transverse mode supports two orthogonal polarizations. The frequency splitting between the modes is determined by the rate of cross-coupling between polarizations due to cavity mirror birefringence. True single-mode excitation is accomplished by driving one of the polarization eigenmodes (generally elliptically polarized), but for su ciently small birefringence an arbitrary polarization may be used. A greater and still outstanding problem experimentally is keeping a single atom localized in the cavity mode within a region of much less than one half-wavelength, such that the value of g, which varies spatially with the mode structure of E, is 5 constant and known. Early cavity QED experiments operated with a thermal beam of atoms, giving a random and changing distribution of coupling strengths and atom number, and non-adiabatic interaction dynamics due to the high speeds [14]. Individual neutral atoms trapped inside a cavity mode have been demonstrated, but still with time-varying values of g and a substantial warping of the level structure due to light shifts [15, 16]. Single ions trapped in a cavity hold more promise, but face their own technical challenge owing to the close presence of dielectric mirrors [17]. Our goal in this work is to explore the dynamics of a real cavity QED system in full quantitative detail beyond the limits of the single-mode JC Hamiltonian. The atoms are multitudinous and freely moving, but can be reduced in number to allow detection and manipulation of individual atoms. The magnetic structure entails all sixteen levels shown in Fig. 1.2a, and supports the creation of ground- state coherence that does not exist for two-level atoms. The cavity birefringence is small enough to permit two degenerate polarization modes, but large enough to give substantial contributions to the measured signals. We rst consider a simple extension to the JC model in order to motivate the measurements that follow. In the subsequent chapters we will introduce greater complexity into the model, requiring a quantum Monte Carlo simulation for the full calculations. 6 1.2.3 Two modes and three levels As a rst complication, we consider a three-level -type atom (Fig. 1.2b) inside a cavity with two orthogonal and degenerate polarization eigenmodes, with annihilation operators a and b ([10]). The j1i-to-j2i transition dipole couples to a photon in mode a with Rabi frequency g, and thej2i-to-j3itransition dipole couples to a photon in mode b with Rabi frequency G. The total spontaneous emission rate is the sum of the emission in the two channels, tot = + . The atom starts in state j1i and the excitation is su ciently weak such that the probability of being transferred to statej3iand subsequently re-excited by the driving light is negligibly small. This avoids the problem that the atom in Fig. 1.2b gets pumped to a dark state of the driving laser ( polarization), while the real level structure in Fig. 1.2a has no dark states. The cavity is driven with a coherent eld (inside the cavity) of amplitude " in mode a. The e ective Hamiltonian in between quantum jumps and in the rotating- wave approximation is [12, 10]: Hmode = " a ay +i hg a NX i=1 j2iih1j ay NX i=1 j1iih2j ! +i hG b NX i=1 j2iih3j by NX i=1 j3iih2j ! i h aya+byb i h + 2 NX i=1 j2iih2j; (1.5) where the sums are over the N atoms, each considered to be identically coupled to the mode. We now have two single-atom cooperativity parameters, C1 = g2=( tot) for 7 the driven mode and eC1 = G2=( tot) for the orthogonal mode. The e ect of eC1 on the driven mode is to reduce the cooperativity by an amount commensurate with emission into the orthogonal mode; instead of C = NC1, we de ne a new driven mode cooperativity, C0 = C1N1 + 2 eC 1 : (1.6) De ning separate values of the intracavity photon number X for the driven (Xk) and orthogonal (X?) modes, the state equations become [10]: Xk = Y 1(1 + 2C0)2 Y (1 4C0); (1.7) X? = Y " 2 eC 1 1 + 2 eC1 #" C0 (1 + 2C0)2 # Y2 eC1C0; (1.8) where the lowest-order approximations are valid for 2C 1 and 2 eC1 1, and we assume weak drive (X 1). The population of the driven mode decreases with C0 as before, while the population of the orthogonal mode increases with C0; in other words, X? contains a fraction of the light Y Xk that would otherwise have been radiated into free space. This thesis studies light emission from the two modes for a variety of ex- perimental parameters. In particular, we study time-resolved photon correlation measurements of light from X?, which reveal information on atomic dynamics and coherences that are invisible in the steady-state average. 8 1.3 Apparatus 1.3.1 Physics cavity We perform measurements on light escaping from a high- nesse Fabry-Perot interferometer (referred to below as the \physics cavity".) The mirrors that form the cavity have re ective coatings from Research Electro-Optics of Boulder, CO, with transmissions T1 = 15 ppm and T2 = 270 ppm at a wavelength of = 780 nm, where the larger transmission T2 corresponds to the mirror on the output side [9]. The spacing between the re ective surfaces is l = 2:2 mm, and the mirrors have radii of curvature 25 cm and 10 cm, respectively. The cavity supports a TEM00 mode with 1=e eld waist w0 = 56 m and Rayleigh range w20= = 1:2 cm, i.e. the mode diameter changes little over the distance between the mirrors. The cavity free spectral range is FSR = c=(2l) = 68 GHz. Barring additional losses, we expect cavity nesse F = 2 =(T1 + T2) = 22 000 and resonance full width at half maximum (FWHM) 2 =(2 ) = FSR=F = 3:1 MHz. We measure instead a FWHM of 5.8 MHz and a nesse of F = 12 000. We infer from this an additional loss channel approximately equal in magnitude to the larger mirror transmission. This loss is most likely due to residual glue deposited on the mirror surface during construction. When inferring the photon number inside the cavity, we increase our estimate by a factor of two due to this loss. This agrees with the results of full Quantum Monte Carlo simulations of the system discussed in Chapter 3. The cavity also has a small intrinsic birefringence, which results in a slight (less than 200 kHz) splitting between the two (elliptical) polarization 9 cavity mirror MOT beams 7.5 cm 1.5-mm hole Figure 1.3: Schematic (not to scale) of vacuum chamber containing op- tical cavity and LVIS apparatus. eigenmodes. Because this splitting is much less than =(2 ), we can e ectively choose an arbitrary polarization basis for driving the cavity (typically vertical and horizontal) and su er only a small background of coherent light in the undriven mode due to birefringence coupling. The cavity length is controlled by two shear piezo-electric transducers (PZT) to which the mirrors are glued. The PZT?s are glued to a stainless steel base, which rests on a stack of alternating layers of Sorbothane (a synthetic polymer from Sorbothane, Inc.) and lead for acoustic isolation [18]. The unit resides in a small stainless steel vacuum chamber at 10 8 Torr (Fig. 1.3), with electrical feedthroughs 10 for the PZT?s. One PZT is driven by the ampli ed output of a Stanford Research Systems SIM960 Proportional-Integral-Di erential (PID) controller, with a feedback bandwidth of about 1 kHz. The other PZT is connected to a high-voltage direct current (DC) power supply to correct for slow thermal drift. A LabVIEW program monitors the output of the PID controller and adjusts the DC voltage to keep the error signal output near zero. The power supply voltage spans about three cavity FSR?s. 1.3.2 Atomic source We extract a continuous cold beam of 85Rb atoms from a modi ed magneto- optical trap (MOT). The atoms are supplied by a Rb dispenser from SAES Get- ters, Inc., operated at a current of 3.4 A. The trap uses a pair of coils in near \anti-Helmholtz" con guration (8 cm diameter and 5 cm separation) to produce a magnetic eld gradient of 10 G/cm. We cool the atoms with three orthogonal retro- re ected beams at 780 nm tuned below the F = 3 ! F0 = 4 cycling transition. Each beam has a power of approximately 15 mW and a diameter of 2 cm. The detuning from resonance is approximately -20 MHz, which we vary by changing the frequency of a double-passed acousto-optic modulator (AOM). The exact parame- ters are optimized to give the greatest ux of atoms. A repumping beam resonant with transitions from the F =2 ground state to the F0 = 3 or 4 excited states is mixed with the MOT beams on a polarizing beam splitter (PBS). The atoms exit via a 1.5 mm-diameter hole in the vertical beam retro-re ecting 11 optic, which consists of a quarter-wave plate ( =4) atop a gold mirror, placed at the bottom of the MOT chamber. The radiation imbalance above the hole pushes out a continuous beam of atoms in a Low-Velocity Intense Source (LVIS) con guration [19]. The atoms have a mean longitudinal velocity of about 20 m/s, determined from comparison of measured data with Monte Carlo simulation (see Chapter 3.) The beam intersects the cavity mode at near normal incidence, 3-4 cm below the hole in the mirror. We estimate a maximum ux of about 1 107 atoms s 1 that pass through the cavity mode, corresponding to a mean mode occupation of 50 atoms with an average transit time of 5 s. For some measurements, we reduce the ux of atoms in order to have a mean mode occupation of less than one atom. This is accomplished by changing the detuning or power of the MOT beams. As part of the measurements, we require control of the magnetic eld inside the cavity, in particular the components parallel to the input laser polarization (vertical) and parallel to the cavity axis (horizontal). To adjust the eld in the vertical direction, we implement a third coil below the \anti-Helmholtz" coil pair. This allows us to vary the eld in the cavity between 12 G (corresponding to a splitting between ground-state Zeeman levels of approximately =2) while a ecting the eld in the MOT region only slightly. We use an additional pair of coils oriented with axis parallel to the cavity axis to control the magnetic eld in that direction as well. Coming out of the MOT in LVIS con guration with a vertical magnetic eld applied, the atoms are left primarily in the m = 3 ground state, with the exact distribution sensitive to laser alignment and detuning. Before they enter the cavity 12 0 2 4 6 8 10 12 14 16 -20 -15 -10 -5 0 5 10 15 20 U n d r i v e n m o d e c o u n ts (1000?s / s e c ) Detuning (MHz) with optical pumping no optical pumping 0 0.2 0.4 0.6 0.8 1 1.2 -20 -15 -10 -5 0 5 10 15 20 X / Y Detuning (MHz) with optical pumping no optical pumping a) b) Figure 1.4: a.) Example of absorption fraction X=Y in driven mode as a function of common detuning of laser and cavity frequency from central m =0 to m0 =0 transition. The optical pumping beam shifts the population toward the center. The magnetic eld value is approximately 5 G and drive strength is approximately one photon in the cavity, with drive polarization parallel to the magnetic eld. Solid lines are least- square ts of the data points to a Lorentzian line shape. b.) Count rates from the undriven mode measured concurrently with data in (a), with Lorentzian ts. 13 mode, we optically pump the atoms to the m = 0 state using a beam resonant with the F = 3!F0 = 3 transition and with polarization vector parallel to the vertical magnetic eld. The beam is combined with light from the repumper laser in a 50/50 polarization-maintaining single-mode ber splitter, and collimated to a waist diameter of 0.5 mm in order to pass between the top of the cavity mirrors and the upper edge of the viewport. Due to strong scattering into the photon detectors from multiple re ections, we are unable to use the beam in a retro-re ected con guration. Therefore, the beam imparts a net momentum kick to the atoms, which must be compensated for with atomic beam alignment. Alternatively, a con guration of MOT beams may be found which leaves a broader distribution of population over the ground states. In this case the optical pumping beam acts as a lter, ejecting those atoms which are not already in m = 0. The intensity of the optical pumping beam and repumper are optimized as a compromise between moving most of the atoms to m = 0 while not ejecting too many from the beam. Figure 1.4 shows a typical measurement sequence used for optimization of the optical pumping beam. The e ect of the beam is to shift and narrow the line shapes for both absorption and emission, with the goal of having narrow Lorentzian peaks centered around 0 MHz (the resonance frequency of the F = 3;m = 0 ! F0 = 4;m0 = 0 transition.) The distributions do not always shift completely to the center (an e ect evident in Fig. 1.4); this is likely due to some combination of incomplete optical pumping and a small ( 1 MHz) drift in the frequency setpoint of the saturated absorption spectroscopy reference, arising from temperature-dependent amplitude modulation (AM) in the electro-optic modulator (EOM) that generates 14 820 nm transfer cavity Pump laser 532 nm Ti:Sapph 780 nm Lock system for stabilizing Ti:Sapph Rb vapor cell LVIS Repumping laser system physics cavity Figure 1.5: Schematic of the laser system used for stabilizing and probing the optical cavity, and for cooling and trapping the atoms. frequency modulation (FM) sidebands for the error signal (see below). 1.3.3 Laser system Figure 1.5 shows a schematic of the apparatus, focusing on the laser sys- tem. The primary source of light at 780 nm is a Titanium:Sapphire MBR-110 laser (Ti:Sapph), driven by a 10 W pump laser at 532 nm (Verdi V-10), both from Co- herent, Inc. A reference cavity-based lock internal to the MBR housing narrows the Ti:Sapph linewidth to less than 100 kHz. Long-term frequency stability and abso- lute frequency referencing are accomplished by saturated-absorption spectroscopy in a magnetically-shielded Rb vapor cell. The error signal is derived from the 85Rb F = 3 ! (F0 = 3;F0 = 4) hyper ne crossover transition. We generate the signal by imprinting FM sidebands at 12 MHz with an EOM, detecting the saturated ab- 15 sorption with a fast photodiode, and mixing the ampli ed photodiode output with the radio frequency (RF) local oscillator. Locking is accomplished with a SIM960 PID controller, the output of which is fed to the external scan input on the MBR control box. The light output of the MBR at this frequency is approximately 1 W. We stabilize the length of the physics cavity with a Toptica DL-100 diode laser at 820 nm, which interacts negligibly with the atomic beam but still re ects strongly from the cavity mirrors. The diode laser frequency itself is stabilized to the Ti:Sapph frequency with a transfer cavity technique. First, we adjust the current to the Toptica laser diode such that both lasers are simultaneously resonant with a TEM00 mode of the physics cavity, changing the cavity length if necessary to nd a suitable point. We then nd a similar resonance overlap for a 25 cm confocal (transfer) cavity into which light from both lasers has been injected, adjusting the length of the transfer cavity as necessary. We use the Ti:Sapph light to lock the transfer cavity length with the Pound-Drever-Hall (PDH) technique [20], a PID controller, and PZT?s. Finally, we stabilize the diode laser frequency to the transfer cavity resonance with another PDH signal, derived from FM sidebands at 9 MHz generated by directly modulating the laser diode current. The rest of the light from the Toptica diode laser is used to generate a PDH signal for stabilizing the length of the physics cavity. Before reaching the cavity, the light goes through an AOM in double-passed con guration and enters a single-mode polarization-maintaining ber. The AOM allows for scanning the cavity frequency by 40 MHz. A monitor photodiode after the ber output is read by an analog-to- digital sampling card and monitored by a LabVIEW program, which controls the 16 RF drive power to the AOM in order to maintain the light output from the ber at a constant power. This is especially important when scanning the cavity frequency, as residual AM sidebands from modulating the diode current give a power-dependent DC o set to the PDH error signal for the physics cavity; this results in a nonlinear scan due to the frequency-dependent AOM di raction e ciency. We drive the physics cavity with Ti:Sapph light that passes through a 780 nm single-mode ber electro-optical modulator (EOSpace) in Mach-Zender con gura- tion to imprint AM sidebands at 150 MHz from the carrier. One sideband is near resonance with the atomic transition and the cavity, becoming the system drive, while the carrier and all other sidebands are re ected from the front cavity mirror. Changing the sideband frequency allows us to probe the frequency response of the cavity-atom system without misalignment of the input laser coupling. Polarization elements (better than 5 10 5 extinction) and mode-matching optics prepare the driving laser before it enters the cavity. Light at 780 nm used for repumping atoms in the MOT is generated from a Sharp laser diode housed in a ThorLabs mount. The laser is slaved to a New Focus Vortex diode laser, which is frequency stabilized to the repumping transition with a Dichroic Absorption Vapor Laser Lock (DAVLL) and PID controller [21]. This system generates about 20 mW of repumper light. 17 1.3.4 Detection system A lens at the cavity output collimates the escaping light, and a zero-order half- wave plate (HWP) aligns the polarization to a Wollaston prism, which separates vertically and horizontally polarized light into separate beams. Additional optics allow the choice of sending both beams to the detectors, or of blocking the vertically polarized beam and splitting the horizontally polarized beam between two paths for coincidence measurements. We split o a small fraction of the vertically polarized beam rst in order to monitor the intensity of transmitted light at 820 nm; the rest of this light is removed from the signal with 5-nm wide interference lters centered at 780 nm (Semrock MaxLine), such that no detectable level of light at 820 nm reaches the detectors. Passing through spatial ltering optics to remove background light, the signal beams enter a dark box where they are detected by two photomultiplier tube (PMT) units (Hamamatsu R636-10), which measure the average light intensity; or, by remotely activating motorized ip mirrors, to two avalanche photodiode (APD) units (Perkin Elmer SPCM-AQR-12 and -13) for time- resolved photon counting, each with quantum e ciency greater than 50%. The total detection e ciency of light from the cavity is about 30%, as calibrated by an optical power meter and neutral density lters of known attenuation. The electronic output pulses from the two APDs are rst duplicated in a transistor-transistor logic (TTL) voltage divider circuit. One copy of each goes to a counter unit (SR400) for measuring average count rates, or to a nuclear instru- mentation module (NIM) logical gate module for coincidence measurements. The 18 other copy of each goes to a computer time-stamp correlator card (Becker and Hickl DPC-230), which records both channels of TTL events with 164 ps resolution. For measuring photon correlation functions, we record the arrival times of all photons and compute the detector cross-correlation in post-processing with a C++ program, with typical user-de ned time resolution of 16.4 ns bins. 19 Chapter 2 Detecting Single Atoms 2.1 Introduction Many applications in atomic physics and quantum information hinge on the ability to detect the presence of individual neutral atoms or molecules in order to perform subsequent state manipulation protocols [22, 23]. Optical detection of atoms involves identifying a change in average light intensity due to the presence of the atom. This typically happens by one of two methods: uorescence detection, in which the atom adds light on top of a background level; or absorption detection, in which the atom removes light from an incident eld. Dispersive detection can e ectively operate in either of these limits, depending on whether the signal with no atoms is bright or dark. All these techniques require the integration of photon ux until a targeted signal-to-noise ratio is reached; this sets the probability of obtaining a false positive for atom detection. Fluorescence detection requires high-numerical aperture optics and careful suppression of background light, and generally operates on time scales of tens of milliseconds. The integration time limits the bandwidth with which new atoms can be identi ed and then manipulated, making this method much better suited for identifying stationary trapped atoms [24, 25, 26]. When detection speed is more important (as when detecting freely moving particles that spend a limited time in 20 an interaction region), a more useful technique is uorescence burst detection [27], which looks for above-average count rates over short time intervals [28, 29]. One particularly impressive example [30] showed detection of falling atoms in free space in 60 s using highly-e cient mirrors and lenses. In contrast to free-space methods that collect light from as many spatial modes as possible, cavity-based methods limit the optical interactions to one or a few resonant modes, yet give signal-to-noise ratios that may be orders of magnitude larger. Collecting uorescence in an optical cavity with axis perpendicular to the driving laser gains the bene t of Purcell-enhanced emission [31] into the cavity mode [32, 33, 34]. Experiments based on changes in cavity transmission, which require strong atom-cavity coupling [35], have achieved single atom detection times of 20 s for moving atoms [7, 32, 36, 37], and as low as 10 s [34, 38, 39] for trapped atoms. Both methods enable strong suppression of background light through spatial ltering of the output quasimode of the cavity. We have developed a cavity technique for atom detection that combines ele- ments of driven cavity QED detection and spontaneous emission collection [40]. The technique allows rapid identi cation of atoms with near maximal coupling to the cavity mode, and their subsequent emission dynamics. This chapter presents our approach, in particular autocorrelation studies of the transmitted light. We identify contributions to the autocorrelation signal from di erent physical processes, in the short time (atomic decay) and in the long time (transit of atoms through the mode). 21 2.2 Detection scheme The basic components of our system are shown in Fig. 2.1. We resonantly drive a TEM00 mode of the optical cavity, previously discussed, with a laser sideband tuned to the 85Rb F = 3 to F0 = 4 transition. Intersecting the cavity axis at near- normal incidence is the beam of cold atoms, each of which interacts with the cavity mode for some time before leaving. Because of the large di erence in time scales (5 s transit time through the mode versus 26 ns excited state lifetime), the atoms undergo dozens of spontaneous emission cycles while in the mode volume. Atomic Beam PBS BSB APD A (start) APD B (stop) Figure 2.1: Schematic of the apparatus with the basic optical elements necessary for coincidence measurements. Vertically polarized laser light (parallel) drives the high nesse cavity traversed by a continuous beam of cold 85Rb atoms from a magneto-optical trap. The output light passes through a polarizing beam splitter (PBS) that separates the horizontal (perpendicular) polarization, sending it to a nonpolarizing beam split- ter (BS) for coincidence measurements using two avalanche photodiodes (APD). The direction of the magnetic eld (B) inside the cavity can be parallel to the exciting polarization (yellow arrow, -polarized drive) or parallel to the direction of light propagation (blue arrow, -polarized drive). 22 B Detection of spontaneous emission B Detection of forward scattering ? ? + ? ? 5P 3/2 , F?=4 5S 1/2 , F=3 m = n-1 n+1n m? = n 5P 3/2 , F?=4 5S 1/2 , F=3 m = n-1 n+1n m? = n ? + ? ? a) b) Figure 2.2: a) A linearly polarized driving eld excites the atom, but a transition to a ground state level with di erent m results in emission in the orthogonal polarization. b) Simpli ed diagram of 85Rb levels relevant to Faraday rotation in an applied magnetic eld. The crux of our detection scheme is the separation of the cavity output eld into two distinct polarization components relative to the polarization of the exciting laser: one parallel (driven mode) which is little a ected by the transiting atoms (as the cooperativity parameter is relatively small), and the other perpendicular (undriven mode) which in the ideal case is populated only by scattered light from the atoms. This con guration in cavity QED was rst used in [41]. As the coupling of the two modes by mirror birefringence is small, the two eigenfrequencies are split by less than 10% of the cavity linewidth, and we can use the two modes for excitation and signal collection, respectively. We consider two di erent con gurations for excitation. The rst has a weak (< 10 G) magnetic eld parallel to the incident polarization, forming the quan- tization axis. The light drives ( m=0) transitions in the atoms in this basis 23 (see Fig. 2.2a). An excited atom can return to its original ground-state m-level through spontaneous or stimulated emission into the parallel (driven) mode; how- ever, there is also some probability to relax into a di erent m-level through a + or ( m = 1) spontaneous emission transition, which produces light that is circularly polarized with respect to the magnetic eld axis and may be collected by the perpendicular (undriven) cavity mode. Although the cavity coupling strength is modest, the amount of collected light increases measurably from that in free space [10]. The second con guration has a weak magnetic eld along the cavity axis, while keeping the same polarization arrangement (see Fig. 2.2b). The linearly polarized drive can then be considered as the sum of two circular components with di erent indices of refraction due to the Zeeman shift. This results in a Faraday rotation of the drive into the perpendicular mode. Though the single-pass Faraday rotation is small, the high cavity nesse allows the light to compound thousands of rotations before escaping, and the e ect becomes appreciable. A polarizing beam splitter (PBS) sends the two modes to separate detectors, which can resolve individual photon events. From the record of photon detections, we analyze the statistics of emission to look for signs of single atom transits. One advantage of collection into a cavity mode is that we can spatially lter out most sources of background light. However, there are remaining contributions to the perpendicular mode from cavity mirror birefringence and background light that cannot be completely eliminated. The optimal parameters for operation minimize the background in uence on the atom detection con dence. 24 2.3 Theory To include the e ects of Faraday rotation, we consider an atom with one ex- cited state (m0 = 0) and three ground states (m = 1;0). (See Fig 2.2b). A magnetic eld B along the direction of light propagation lifts the ground state de- generacy through the Zeeman e ect, resulting in a shift in the dispersion curves and a di erence in the phase velocities of the two circular components of light, causing Faraday rotation [42]. We excite the system with vertically polarized light, which is an equal superposition of right- and left-circularly polarized light. The light in- teracts with the atom and becomes elliptically polarized; the plane of polarization rotates through an angle. The light then passes again through the atom after re- ection on the mirror and compounds the rotation. When the light exits the cavity, the total rotation is that of a single pass multiplied by the number of times the light passes through the atom, indicated here by the presence of the cooperativity C (which is proportional to F): = 2gL BB= h tot1 + (2g L BB= h tot)2 C; (2.1) where gL is the Land e factor and B is the Bohr magneton. The addition of this e ect changes Eq. 1.8 (assuming j j 1) into: X? = Y 2 eC1C +j j2 : (2.2) Equation 2.2 shows that the signal in the perpendicular mode now consists of two parts: a background level of spontaneous emission (which may be considered as ( + ) polarization if the drive is ( + + ), in full analogy to the basis with 25 drive), and Faraday-rotated drive which increases from zero to a maximum value of X?Faraday = YC2=4 when B = h tot=(2gL B). For simplicity in what follows we treat the atom-cavity coupling in the basis of drive. We analyze the two-time intensity autocorrelation function of the light from the perpendicular mode to understand the dynamics of photon emissions that con- tribute to the steady-state value of X?. We do not distinguish between the two components in Eq. 2.2, treating both as arising from near-resonant scattering of incident light (resonance uorescence) in which the cavity serves only as a mode in which to collect the light. For a stream of photons scattered from a single two-level atom undergoing sta- tionary, resonant excitation, the intensity autocorrelation function takes the familiar form for resonance uorescence antibunching [12]: g(2)A ( ) = 1 e (3 =4) (cosh + 3 4 sinh ); (2.3) where = ( =4)p1 8Y is a function of the drive intensity Y. The antibunching referred to here means an increasing slope for increasing times: dg(2)( )=d j >0 > 0. The condition g(2)(0) < 1 indicates that the photons obey sub-Poisson statistics, and is not a necessary condition for the photons to exhibit antibunching [43]. The modi cation of Eq. 2.3 for the case of an atomic beam passing through a localized excitation region appears in Ref. [44], which adds a number of e ects: i., the multiplication of Eq. 2.3 by a window function to include the shape of the mode function encountered by each transiting atom; ii., the scaling of Eq. 2.3 by the inverse of the steady-state mean number of interacting atoms, N; iii., the addition 26 of a term jg(1)A ( )j2 to include the beating of elds emitted by di erent atoms (the Hanbury Brown and Twiss e ect [45]); iv., the addition of a constant o set of one arising from uncorrelated emissions from di erent atoms; and v., the reduction of the correlation factor due to the presence of completely uncorrelated background light. Further generalizations appear in Refs. [46, 47], where the authors also consider the product of the elds that come from atomic emission with a temporally uncorrelated but spatially mode-matched eld from background light. We are able to describe the total correlation function as: g(2)( ) = 1 + 1(1 +R b=Rs)2 f( )g(2)A ( ) N +F( ;Rb=Rs;N) (2.4) where Rb and Rs are the average count rates for background and signal, respectively, and f( ) = [cos( )e + 1]e ( =T)2 (2.5) is an empirically-determined window function describing motion of the atom through the cavity mode with Gaussian temporal width T, where a small tilt of the atomic beam with respect to the cavity axis normal results in motion through the standing- wave lobes at frequency . These oscillations decay at a rate much faster than the transit time, due to the spread in velocities from the atomic source. The function F corresponds to any additional terms related to the beating of signals from di erent atoms or of signal with background light; since this term represents classical noise sources, it can only contribute light which is bunched around = 0 [48]. Any observation of antibunching in our signal must come from the single-atom term g(2)A ( ). An explicit observation of F would require a high atom density (N 1) 27 since the width is comparable to that of the antibunching signal. (A technique in which the antibunching width was arti cially made wider was demonstrated in [49] in order to see this term.) 2.4 Preliminary data analysis 2.4.1 Faraday rotation The Stokes parameter formalism [50] allows a measurement of the rotation angle if the amplitudes of the electric eld in the two polarizations "? and "k are known: j j= j"k"?jj" kj2 j"?j2 "? "k ; (2.6) where the approximation is made forj"kj j"?j. We measure the count rates in the perpendicular mode as a function of applied magnetic eld along the cavity axis to determine the location of maximum Faraday rotation. For the measurements in this chapter we do not optically pump the atoms to m = 0 before they enter the cavity. The peak rotation is di erent depending on the sign of the magnetic eld, probably due to an asymmetry in the population distribution. We obtain a maximum of j j = 0:035 0:007 rad from the measured increase in the count rate with a eld magnitude of 3.3 Gauss (calibrated from Zeeman shifts in absorption.) To make explicit the e ects of the cavity, we estimate the e ective number of maximally- coupled atoms by using Eq. 1.7 and measuring the driven mode transmission. This was greater than 0.95, giving a bound of C <0.01 or N <0.1 e ective maximally- 28 coupled atoms (for transitions from m = 0). Calculating the free-space absorption length for 0.1 atoms and our cavity waist, we would expect a maximum rotation of the order of 3 rad [42]. The enhancement by a factor of approximately 10,000 is a result of the cavity nesse. 2.4.2 Photons per atom The formula of Mandel relates the photon number distribution P(n) with the atom number distribution Patom(m): P(n) = X m P(njm)Patom(m); (2.7) where P(njm) = ( m)n exp ( m)=n! is the conditional probability of detecting n photons when there are m atoms in the cavity volume, each contributing a mean of photons to the signal with a Poisson distribution of number [51]. This shows that the super-Poisson uctuations of the light arise from the combined Poisson uctuations in atom number and photons scattered, with an extractable parameter. We follow the method of Ref. [32] to relate the mean and mean-squared photon number (hni, hn2i) to : hn2i hni 1 =hnigaa + ; (2.8) where the atom-atom correlation function is gaa = (hm2i hmi)=hmi2, which for Poisson atomic uctuations reduces to gaa = 1. Figure 2.3 shows a detail of the plot of the experimental values for the left side of equation (2.8) with the two di erent detection methods: Faraday rotation on the 29 top and no Faraday rotation on the bottom. We calculate the statistics based on 300 s time series of photon detections that we divide into a series of time bins of equal width. Longer bin sizes have a larger average number of counts (horizontal axis). The intercept gives the value of . A linear t for bins from 50 s to 100 s (hni= 1.81 to 3.63, o the scale of the plot in Fig. 2.3) in the Faraday case gives a vertical axis intercept of = 0:196 0:003 photons detected per atom, and a slope of 1.033 0.001. The intercept in the no Faraday case (for a t fromhni= 0.50 to 1.00) is = 0:036 0:002 photons detected per atom and a slope of 1:026 0.003. The drive in both cases is two photons. The values of the slopes indicate that the assumption of Poisson statistics for the atom number is well justi ed. Figure 2.4 shows the extracted values of as a function of cavity photon number (driving intensity) spanning more than two orders of magnitude. There are data points (rhombs) that come from Faraday rotation and one with spontaneous emission alone (circles). The latter is smaller by almost a factor of ve. increases linearly with drive until atomic saturation intensity (nsat = 5:3 photons for the m = 0 to m0 = 0 transition), where in addition to saturation of the scattering rate, optical pumping e ects also start to enter. Each point on Fig. 2.4 comes from a least-squares t similar to those in Fig. 2.3. The count rates in the perpendicular mode and the values of give an atomic ux through the mode of approximately 160,000 atoms s 1. (The ux is set much smaller than the maximum available in order to isolate primarily single-atom events.) 30 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 0.0 0.1 0.2 0.3 0.4 0.00 0.05 0.10 0.15 0.20 0.25 0.30 /-1 (no Faraday) /-1 (Faraday) 0.0 Figure 2.3: Examples of data (circles) and least-squares ts (continuous line) to Eq. (2.8) in the large-hni region for Faraday rotation (top) and for no Faraday rotation (bottom) in the cavity as a function of average number of photons in a time bin (hni). 31 0 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.01 0.1 1 10 ? (photons/atom) /n sat (driven mode photons) Figure 2.4: Counts detected from a single atom (rhombs for Faraday, circle for non Faraday) as a function of intensity of the exciting laser normalized by the saturation photon number. The arrows indicate the points extracted from the plots of Fig. 2.3. The statistical error bars are smaller than the size of the data points. 32 2.5 Analysis and optimization of detection Three identi able sources of background counts contribute to the measurement of the perpendicular mode: intrinsic detector dark counts (approximately 300 s 1 each), scattered light from the MOT beams (approximately 2000 s 1), and the birefringence of the cavity (less than 5 10 5 of the driven mode count rate). The light from the MOT dominates the background count rate at low driving intensity; however, at higher intensity, light from cavity birefringence is the main contributor to the degradation of the signal-to-background ratio. A single photon escaping in the perpendicular mode most likely comes from an atom in the cavity; however, there is at best still a 4% probability that it comes from background counts. We suppress the probability of a false atom detection by requiring photon coincidences in a time window smaller than the transit time of an atom across the cavity mode [7]. We have implemented this in real time for preliminary measurements using a NIM electronic coincidence module. The rst perpendicularly polarized photon detection (\start") opens a gate of variable width which allows counting pulses from the \stop" APD. Detection of one or more of these yields an atom detection event. Figure 2.5 shows the signal-to-background ratio calculated from photon time series for the rate of single photon detection (open rhombs), two-photon coincidence (open squares), and three-or-more-photon coincidence events (open triangles) in a 1 s window as a function of driving intensity. The maximum occurs at the same driving intensity for all three. Two photons within a 1 s window improves the signal-to-background ratio by more than an order of magnitude compared to de- 33 /n sat (driven mode photons) signal-to-background ratio 10 0 10 1 10 2 10 3 0.1 1 10 10 4 Figure 2.5: Signal-to-background ratio as a function of normalized driv- ing laser intensity for single detection (open rhombs), double coincidence detection (open squares), and triple coincidence detection (open trian- gles) in a 1 s window with Faraday rotation of the drive. tection of single photons. The three-photon coincidence gives a better ratio but signi cantly decreases the rate of detection as evident in the error bars associated with each point. The three sets achieve their maxima just before the atomic satu- ration intensity, due to background counts from cavity birefringence that continue to increase linearly with drive. The contribution to the coincidences from detector afterpulsing is small (less than 1%) and does not a ect the results. To evaluate the probability of making a false detection for various coincidence gate times, we extract the delity of detection F versus gate length for two-photon coincidence based on the time record. We calculate the waiting time distribution 34 of photon arrivals on a single APD by making a histogram of the time separation between consecutive detections. The integral of the waiting time distribution gives the number of two-photon coincidences c in a given time window. We calculate c with and without atomic ux to obtain: F = 1 cwithout=cwith: (2.9) The delity (one minus the probability of error) reaches more than 99% at 0.1 s, and 99.7% at 1 s with Faraday rotation, but only 96.7% at 1 s with detection of spontaneous emission alone. The delity decreases at longer times. The delity is optimal for times between 1 and 5 s, determined by the distribution of atom transit times. 2.6 Intensity autocorrelation We ensure that our coincidence detection scheme is sensitive to single atoms by measuring the intensity autocorrelation function (g(2)( )) of the perpendicular mode under very weak driving intensity at the same atomic beam density used during detection measurements. If we consider the light in the perpendicular mode as coming from resonance uorescence, the antibunching can only arise from photon pairs from the same atom. The observation of antibunching in our signal is strong indication of our ability to detect photon coincidences from a single atom in less than 1 s. Figure 2.6 shows two examples of an autocorrelation function around = 0 with and without Faraday rotation for the same driving intensity (0.4 nsat). The 35 -1.0 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1.0 1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.4 2.6 2.8 g (2) ( ? ) ? (?s) Figure 2.6: Intensity autocorrelation function (g(2)( )) of the perpendic- ular mode with (triangles) and without (diamonds) Faraday rotation, at a driving intensity of 0.4 nsat. antibunching is visible and lasts for a time of the order of the excited state lifetime (26 ns). The area under the curve is larger with Faraday rotation, indicating a substantial increase in photon ux from individual atoms. The antibunching is less pronounced without the rotation because of the lower signal-to-noise ratio. Since we record the photon detections as continuous time series, we can also look at the long-term behavior of the autocorrelation in the context of resonance uorescence as presented in Ref. [44, 46, 47]. We t the data to Eq. 2.4 with free parameters N, T, , and , excluding the term F( ) and the region j j <50 ns 36 where it is important. We nd excellent agreement in this outer region (reduced 2=1.03 for points out to 5 s). Fig. 2.7 shows the t to the data taken at a driving intensity of 0.24 nsat. The damped oscillation at =1.5 MHz corresponds to motion through the standing wave with a period of 0.67 s (a velocity of 0.58 m/s along the cavity axis) and a damping time of 1= =0.29 s. Monte Carlo simulations of the transit show good agreement with this damping time based on the geometric collimation of transverse velocities from the source [19]. The Gaussian background gives a mean atom number of N=0.88, and a 1/e waist of T=2.7 s (a mean velocity of 20.7 m/s across the Gaussian mode). The ratio of the velocities gives a beam tilt of 1.6 with respect to the cavity axis normal. Although we take care to align the cavity under the exit hole of the atomic source, the cavity mode is not necessarily centered on the two mirrors, and the transverse beam width allows for small inclinations of the beam. The full structure of the correlation function depends not only on the above parameters (properties of the atomic beam), but also on the intensity of the drive, which can change the atomic response or increase the amount of background light from cavity birefringence. Fig. 2.8 shows the evolution of the autocorrelation func- tion as the driving intensity increases by two orders of magnitude. We observe a transition from antibunching to bunching with higher drive. The behavior comes from the term F( ) in Eq. 2.4 which includes beating of the birefringence light with the atomic emissions. This is in contrast to the transition seen in Ref. [49], where the atomic density is increased such that the constant jg(1)A ( )j2 term (part of our F( )) becomes visible. In separate experiments at low intensity we have also 37 -5 -4 -3 -2 -1 0 1 2 3 4 5 0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.4 2.6 ? (?s) g (2) ( ? ) Figure 2.7: Intensity autocorrelation function (g(2)( )) of the perpen- dicular mode at 0.24 nsat. The continuous line is the theoretical t to Eq. 2.4. 38 -8 -6 -4 -2 0 2 4 6 8 10 -1 10 0 10 1 0.5 1.0 1.5 2.0 2.5 ? (?s) n/n sat g (2) ( ? ) Figure 2.8: Intensity autocorrelation function (g(2)( )) of the perpendic- ular mode for di erent values of photon number (intensity) in the driven mode. followed the disappearance of the antibunching as we increase the number of atoms. Even when the actual density of atoms is kept constant, the e ective excitation volume and thus the value of N increases with power, as atoms near the nodes and wings of the mode start to interact [52]. This process reaches a maximum above saturation intensity, when the only unsaturated atoms are too weakly coupled to emit into the mode. Fig. 2.9a. shows the extracted values of N versus driving 39 intensity for the autocorrelations in Fig. 2.8. The line shows the expected shape of the saturation, assuming that the probability of collecting a photon from an atom is proportional to the product of a constant g2(x;y;z) for collection and a saturating strength of excitation ([(n=nsat)g2(x;y;z)]=[1+(n=nsat)g2(x;y;z)]), which is consistent with the expression Y2 eC1C in Eq. 1.8, modi ed for strong driving. We obtain the theoretical line by integrating this product over a volume much larger than the cavity mode, and using a least-squares t to scale the vertical axis and nsat. Using the extracted parameters of mean atom number and mean transit time from the two-APD correlation measurements, together with the extracted values of from the one-APD measurements, we can predict the expected macroscopic count rates as Rs = N =2T. Fig. 2.9b shows that we obtain excellent agreement for low intensities, while the highest intensity drive (largest birefringence background) disagrees signi cantly. This comes from the large contribution from the beating be- tween background and signal, which generates the large central peak in the measured autocorrelation. We emphasize that although the mean number of interacting atoms is ap- proximately one, this is not the same as the e ective number of maximally-coupled atoms, since most are weakly coupled. To explore this numerically, we use the work of Carmichael and Sanders [53]. We distribute atoms randomly and uniformly across the cavity mode function such that the mean density is 0.88 atoms within two mode waists (similar to the measured value for low driving intensity). Allowing the atom number to uctuate with a Poisson distribution, we sum the individual coupling 40 0 1 2 3 4 5 6 7 8 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 n/n sat atoms 0 1 2 3 4 5 6 7 8 9 10 0.01 0.1 1 10 n/n sat 10 -4 counts/s a) b) Figure 2.9: a) Measured mean number of interacting atoms, and theory curve showing the expected saturation. b) Measured ( lled rhombs) count rates in the perpendicular mode together with those predicted (open squares) from the extracted mean number of atoms, photons per atom, and transit time. 41 strengths for each con guration of atoms and weight by the atomic number distri- bution to obtain the mean steady-state e ective number of atoms, which in this case is 0.04. Using the mean value of C1=0.09 for transitions from the ground state, this gives a driven mode transmission of approximately 0.99 from Eq. 1.7, consis- tent with the experimental conditions for single atom detection. A rapid decrease in probability above one e ective atom indicates that we operate in a density regime of single-atom coupling, with a low probability of having two atoms simultaneously coupled by more than half of the maximum value. This numerical result together with the antibunching in the perpendicular mode counts con rms that the measured signals indeed arise primarily from single-atom emission bursts. 42 Chapter 3 Ground-State Quantum Beats 3.1 Introduction Quantum beats are oscillations in the radiation intensity of an ensemble of ex- cited atoms due to interfering emission pathways. They must be counted amongst the earliest predictions of quantum mechanics [54]. So-called \Type-I" atoms ex- hibit beats at the separation frequency of two excited states which, prepared in a superposition, decay to a common ground state. The preparation may be achieved in a number of ways, e.g., through pulsed optical excitation [55, 56] or cascade emission [57]. In all cases the coherence lasts for the excited state lifetime, unlike ground-state coherence, which can last long enough to be interrogated later and, for this reason, is favored by the eld of quantum information. We consider an unusual situation where ground-state coherence gives rise to a long-lived quantum beat in spontaneous emission [58]. As repeatedly noted [54, 59, 60], QED predicts no beat in the decay of a \Type-II" atom to non-degenerate ground states, since they are orthogonal. Nevertheless, while beats may be absent from the mean intensity, they can still lie hidden in the uctuations. In an elegant experiment in the 1950?s, Forrester et al. [61] showed this for the (classical) beating of light from a pair of incoherent sources. We proceed in similar spirit; we recover a long-lived quantum beat from uctuations. 43 In contrast to recent experiments [62, 63], which aim for deterministic quantum control, the ground-state coherence in our experiment is both prepared and read out by spontaneous emission. Moreover, our measured beat is di erent from that seen by Schubert et al. [64], where an interference occurs in absorption rather than emission. Creation of coherence through spontaneous emission, so-called spontaneously generated coherence, has been discussed in the theoretical literature [65, 66, 67] and indirect experimental evidence reported for spontaneous creation of electron spin coherence in charged GaAs quantum dots [68]. We detect only spontaneous emission, and therefore make a direct and unambiguous observation of spontaneously generated coherence. We realize, in a continuously driven variation, the scheme of Zajonc [69] for generating ground-state quantum beats on the principle of the quantum eraser [70]. 3.2 Theoretical model Beginning with an outline of the scheme, we consider rst one idealized atom then a realistic atomic ensemble. Consider an atom with Zeeman structure in its ground and excited states (Fig. 3.1a) interacting with degenerate, orthogonally po- larized cavity modes, H and V; a weak magnetic eld sets the quantization axis in the V direction, and mode V is weakly and continuously driven (Fig. 3.1b). The atom is prepared in state jg0i from which it is excited to je0i by the V mode (Fig. 3.2A). It may return to the ground state emitting a ; + or photon, or 44 Figure 3.1: Experimental system for observation of ground-state quan- tum beats. (a) -excitation of an F = 3 to F0 = 4 transition with scattering of a rst (red) and second (blue) photon into the H mode; b) schematic of the apparatus, HWP: half-wave plate, PBS: polarizing beam-splitter, BS: beam-splitter, APD: avalanche photodiode. 45 e -1 ? ? ? ? e 0 e +1 e -1 e 0 e +1 e -1 e 0 e +1 e -1 e 0 e +1 g -1 g 0 g +1 g -1 g 0 g +1 g -1 g 0 g +1 g -1 g 0 g +1 A B C D } ? e }? g Figure 3.2: Simpli ed model of quantum eraser process. A) excitation from g0 to e0; (B) Spontaneous decay by transitions to superposition of g 1 and g+1; (C) excitation to superposition of e 1 and e+1; (D) Spontaneous decay by transitions back to g0. 46 any linear combination conserving angular momentum. In the given geometry, only + or light couples to the H mode, with the helicity undetermined; thus, if the emitted photon escapes the cavity before being reabsorbed, its detection places the atom in the superposition j 0i=jg 1i+jg+1i (Fig. 3.2B). The atom is now in the ground state with angular momentum perpendicular to the magnetic eld, and thus performs Larmor precession. When subsequently reexcited by the driven V mode, state j 0i= ei ( )je 1i+e i ( )je+1i (3.1) is reached, with phase ( ) gained through its precession in the ground state (Fig. 3.2C). From here the atom can decay back tojg0i, emitting a second H-mode photon (Fig. 3.2D). The probability to do so depends on (t), giving rise to quantum beats. In summary, there are two paths for scattering a pair of photons into the H mode: jg0i!je0i!jg+1i!je+1i!jg0i and jg0i!je0i!jg 1i!je 1i! jg0i. The phase gained from the ground-state Zeeman shift (precession) di ers along the two paths, which interfere to produce oscillations in the rate of delayed coincidences|i.e., in the correlation function g(2)( ). Note that after the rst pho- ton is detected \which path" information is available, since jg+1i and jg 1i are distinguishable in principle. This information is erased by the second detection. What we have presented is a single-atom idealization. It neglects the pres- ence of more than one atom in the cavity (not admissible for an atomic beam), spontaneous emission to non-cavity modes, the nite cavity decay rate, and the full 47 complement of magnetic sublevels for the employed F = 3 to F0 = 4 transition (Fig. 3.1a). These features are included in a full quantum trajectory treatment, including a Monte Carlo simulation of an atomic beam [71]. Two approximations are adopted: (i) the driven mode is treated semiclassically (with absorption still taken into account), and (ii) reabsorption of H-mode photons is neglected. The approximations are justi ed by our moderate dipole coupling strength. Consider rst an atom prepared injg0ithat has not yet su ered a spontaneous emission. Letjaii,ja0ii, andja00iidenote unnormalized states expanded, respectively, over the mi = 0, mi = 1, and mi = 0; 2 sub-manifolds, as indicated by the black, red, and black plus blue levels of Fig. 3.1a. These states correlate with the scattering of zero, one, and two photons into the H mode. Should the atom undergo a spontaneous emission (to non-cavity modes), the expansion manifolds, after the quantum jump, are unchanged for a -emission but move one step to the right or left for a -emission (mi!mi 1). Keeping track of these shifts, the system state is expanded as j i = j0ijAi+j1i NX i=1 ja0iijAii ! +p2j2i NX i=1 0 @1 2 NX j6=i=1 ja0iija0jijAiij +ja00iijAii 1 A; where j0i, j1i, and j2i denote zero, one, and two photons in the H mode, N is the (time-varying) number of interacting atoms,jAi=ja1ija2i:::jaNi,jAii is the state jAi with jaii omitted from the product, and jAiij is the state jAi with jaii and jaji omitted from the product. This base state evolves under the coherent drive and coupling to the H mode, the coming and going of atoms as they transit the 48 cavity, and spontaneous emission. From it, at regular sample times, we initiate the collapsed state j 0i = ^bj i, where ^b annihilates an H-mode photon; thus, the ground-state coherence is prepared as an entangled state of N atoms, which then evolves in parallel with j i: j 0i = j0i N0X i=1 jb0iijAii +j1i N0X i=1 0 @ NX j6=i=1 ja0jijb0iijAiij +jb00iijAii 1 A; (3.2) where jb0ii = ja0ii and jb00ii = 2ja00ii at the start of a sample, after which jb0ii (jb00ii) andja0ii(2ja00ii) di er due to their correlation with zero (one) rather than one (two) H-mode photons; N0 N is the number of surviving entangled atoms, i.e., those remaining in the cavity. The source of the quantum beat is the atomic ground-state coherence preserved in the vacuum of the cavity, i.e., the rst term in Eq. (3.2). The system ground state|both atoms and cavity|does not decay, and through -excitation drives a sustained excited-state oscillation in the atom mirroring Eq. (3.1): gmi 1ei g i(mi 1) jemi 1i+ gmi+1e i g +i(mi + 1) jemi+1i; (3.3) with = e g, where g ( e) are ground-state (excited-state) Zeeman detunings, gmi 1 are Clebsch-Gordan coe cients, and is the excited state linewidth; mi tracks the state reached by atom i through possible spontaneous emissions. From Eq. (3.3), oscillation at the ground-state frequency g is passed to the probability amplitudes for emitting a second H-mode photon through jb00ii and ja0jijb0ii in Eq. (3.2); note thatjb0iiis a source term driving the equation of motion forjb00ii(the blue wavy lines 49 plus -excitation in Fig. 3.1a). Only the amplitude and phase, but not the frequency of the oscillation, will be a ected by a detuning of the drive from the atom. The probability for emitting a second H-mode photon contains a term pro- portional to hb00ijb00ii, summed over all surviving entangled atoms. It accounts for the scattering of a rst and second photon by the same atom and shows the quan- tum beat introduced above. There is also a term computed from the norm of ja0jijb0ii+ja0iijb0ji, i6= j, which adds probability amplitudes for \a rst photon from atom i and a second from atom j," and \a rst photon from atom j and a second from atom i." If the scattered elds were classical,Ebi at time t andEai at time t+ , one would have the intensityjEajEbi +EaiEbjj2, where the interference 2Re(Ea j EbjEb i Eai ) disallows assignment of the rst detection (superscript b) to the intensity of either source (similarly the second). An assignment may be made in principle, however, since one and only one atom changes its ground state when the photon is detected. The change could be seen if one looked, thus providing \which-path" information. As we do not look, the two quantum paths, \atom i then j" and \atom j then i", interfere|after a second detection both atoms have changed state, so the \which- path" information is erased. This interference also creates a quantum beat, as a term analogous to the Hanbury Brown and Twiss thermal bunching peak [45], or to the beating of light from independent atoms observed in Ref. [61]. 50 -4 -3 -2 -1 0 1 2 3 4 1.0 1.2 1.4 1.6 1.8 2.0 b 1.0 1.2 1.4 1.6 1.8 2.0 a ?(?s) g (2) ( ? ) g (2) ( ? ) Figure 3.3: Intensity correlation function of the H mode for 4{6 photons in the V mode (when no atoms are present) and a 5G magnetic eld: a) N = 0:2 and b) N = 2:0. 51 3.3 Measurements Figure 3.3a displays a measured correlation function at a magnetic eld of 5G, with a beat frequency of 4:9MHz. The V mode is populated with 4{6 photons, on average, when no atoms are present, and the atomic ux corresponds to N 0:2 e ective maximally coupled atoms; most of the time there are no atoms well-coupled to the mode [53]. The oscillation has low visibility and sits atop a raised Gaussian background whose correlation time is given by the transit time of an atom ( 4 s). The dominant beat for small N is that arising from the emission of two photons by the same atom (see Fig. 3.9c). Figure 3.3b displays the measured correlation function with N larger by a factor of 10 but otherwise similar conditions. The beat visibility is improved. The background is also removed, evidence that there is now an equal (at least) two-atom quantum beat (see Fig. 3.9d). The beat frequency is reduced to 4:7MHz, an indirect e ect of increased absorption by the additional atomic ux, which reduces the V-mode photon number|by a factor of two|and thus also the light shifts. A detailed study of light shifts is presented in Chapter 4. Figure 3.4 shows a similar measured correlation function for a magnetic eld of 4 G (A) and its calculated FFT power spectrum (B). The main peak near 3.5 MHz corresponds to the quantum beat resonance. A smaller peak at half this frequency is also visible, the result of homodyne interference with drive light mixed in by cavity birefringence (see discussion below). The small sidebands on the main peak correspond to a slight modulation of the beat envelope visible in (A). This is the result of the small (1{2 ) deviation from normal incidence at which the atomic beam 52 g (2) ( ? ) 1.5 1.4 1.3 1.2 1.1 1.0 0.9 ? [?s] 0.0 2.0-2.0-4.0 4.0 6.0-6.0 0.0 2.0 4.0 6.01.0 3.0 5.0 0.0 1.0 0.2 0.4 0.6 0.8 Power [arb. units] ? [MHz] A B Figure 3.4: Intensity correlation function and spectrum. A) g(2)( ) mea- sured at 4 G magnetic eld. B) FFT power spectrum of data in (A). The peak at 3.5 MHz corresponds to approximately twice the Larmor frequency for the ground states in a 4 G eld. 53 -5 -4 -3 -2 -1 0 1 2 3 4 5 0 1 2 3 4 5 6 7 8 0.8 1.0 1.2 1.4 1.6 1.8 magnetic field (G) g (2) ( ? ) ? ( ? s) Figure 3.5: Evolution of g(2)( ) with increasing magnetic eld. traverses the cavity mode standing wave, resulting in a sinusoidally varying value of g and amplitude modulation in the spontaneous emission rate. (This is the same modulation envelope from Fig. 2.7.) Figure 3.5 illustrates the change of the autocorrelation function with increas- ing magnetic eld. The beat frequency increases while the envelope from the atomic transits stays constant. The beat visibility decreases for large magnetic eld. Fig. 3.6 shows the e ect of stronger driving on the beats. The amplitude of oscillation damps until at high drive only a thermal bunching peak remains. At even higher drives this peak also disappears, the result of the larger fraction of coherent light in the count rate, due to cavity birefringence and possibly the splitting of the sponta- 54 ?3 ?2 ?1 0 1 2 3 2 1 1.5 ?3 ?2 ?1 0 1 2 3 2 1 1.5 ?3 ?2 ?1 0 1 2 3 2 1 1.5 ?3 ?2 ?1 0 1 2 3 2 1 1.5 ?3 ?2 ?1 0 1 2 3 2 1 1.5 n=0.4 n=1.2 n=4.2 n=13 n=46 g (2) ( ? ) ?(?s) Figure 3.6: Evolution of g(2)( ) with increasing photon number in V. 55 Figure 3.7: Evolution of the quantum beat as V-mode light is mixed with the H mode. Upper and lower extremes correspond to approximately six times more detector counts from the V mode than spontaneous emission. The parameters are 1:2 photons in the V mode, a magnetic eld of 4G, and N = 1. neous emission spectrum with strong driving. The mechanism for the disappearance of the beats in these two gures will be explained in Chapter 4. Figures 3.7 and 3.8 illustrate the change in the observed beat when the po- larization presented to the detector is not taken orthogonal to the polarization of the drive but is allowed to rotate by a few degrees. The rotation is controlled by changing the angle of the HWP placed between the cavity and the PBS (Fig. 3.1b). This mixes a small amount of drive light with the scattered light. With increas- ing fraction of drive light, the beat is eventually dominated by a homodyne term (Fig. 3.9e) arising from the correlation of a photon scattered into the H mode with a photon from the drive; thus, as in the two-atom case, interfering time orders yield a quantum beat. This beat oscillates at half the frequency and allows the correla- tion function to dip below one. Generally, some drive light is coupled into the H 56 ?3 ?2 ?1 0 1 2 3 0.5 1 1.5 2 ?3 ?2 ?1 0 1 2 3 0.5 1 1.5 2 ?3 ?2 ?1 0 1 2 3 0.5 1 1.5 2 ?3 ?2 ?1 0 1 2 3 0.5 1 1.5 2 ?3 ?2 ?1 0 1 2 3 0.5 1 1.5 2 0.0 ? 0.5 ? 1.0 ? 1.5 ? 2.0 ? ?(?s) g (2) ( ? ) Figure 3.8: Evolution of g(2)( ) with homodyne mixing. Parameters as in Fig. 3.7, with angle of HWP labeled with respect to position of maximum extinction. 57 mode through a small birefringence of the cavity mirrors. In Fig. 3.7, the evident asymmetry with respect to angle is probably due to a small residual magnetic eld component in the direction of the cavity axis. Quantum trajectory simulations agree well with the measurements. Figure 3.9a displays a computed correlation function overlaying the data of Fig. 3.3b. A mean velocity of 22 m/s ts the decay of coherence well. Other parameters, such as the delity of the optical pumping, atomic beam tilt, and background from birefringence or elsewhere are more di cult to accurately determine (no background correction is made to the data). Note that with a background light amplitude from birefrin- gence, the post-detection state is j 0i+ j i, and carries noise from the atomic beam and spontaneous emission (absorption on the V mode). Plausible parameters are used for the plots of Fig. 3.9, with the quality of the t primarily determined by the atomic beam density, which controls the relative size of the one- and two-atom quantum beats, and the strength of the drive, which controls the level of sponta- neous emission. For the parameters of Fig. 3.9a, an atom passing near the cavity axis (within half a mode waist) typically undergoes 10 spontaneous emissions to non-cavity modes during its transit, yet the coherence, merely passed between di erent mi-multiplets, is preserved. In contrast, spontaneous emission decoheres an excited-state beat. Figure 3.9b displays an example of a correlation function with mixed drive light (as in Fig. 3.7), together with its breakdown into three contributing quantum beats: one-atom interference (frame c), two-atom interference (frame d), and homodyne interference (frame e). The pieces lie in one-to-one correspondence with known 58 0.0 0.1 0.2 c 1 I 0.0 0.4 0.8 d 2 I -3 -2 -1 0 1 2 3 -0.3 0.0 0.3 e ? ? ( s) I 3 1.0 1.2 1.4 1.6 1.8 2.0 b ( 2 ) ( ) ? g 0.8 1.0 1.2 1.4 1.6 1.8 2.0 ( 2 ) a ( ) ? g Figure 3.9: Calculated intensity correlation function with the birefrin- gence background set at 1% a) and 10% b) of the H-mode photon num- ber; c), d), e) the three interference terms contributing in b); curve a) is plotted against the data of Fig. 3.3 b). Parameters are N = 4, 8 photons in the V mode with no atoms present, and an atomic beam tilt of 1:3 degrees. 59 terms in the intensity correlation function for a source comprised of many scatterers and a coherent background, e.g. Eq. (11) of [46], where the correlation function is the sum of a single-atom term, g(2)A ( ), two-atom term, jg(1)A ( )j2, and a homodyne term, Re[g(1)A ( )]. 60 Chapter 4 Backward Light Shifts and Quantum Jumps 4.1 Introduction Spontaneous emission is a fundamental damping mechanism for coherence stored in optically driven atoms [72]. At random intervals, the atomic dipole be- tween ground and excited energy levels disappears in a sudden projection onto a single state, an event referred to as a \quantum jump" [12]. Under certain condi- tions quantum jumps are directly observable in the measurement record, as when an atom jumps intermittently between \bright" and \dark" states [73, 74]; experimen- tal measurements of switching in atomic [75, 76, 77] and photonic [8] systems provide clear visualization of the discrete nature of the dynamics. Although quantum jumps often bring decoherence and spectral broadening{they give breadth to the natural linewidth of resonance uorescence{certain coherences may also be preserved or even created in these processes under appropriate conditions [58, 65]. In contrast to the dynamical energy changes associated with quantum jumps in a driven atom, light coupled to an atomic transition can also induce a structural change in the energy levels through the AC Stark shift [78]. The energy eigenvalues move in a direction determined by the sign of the light detuning from resonance, but populations and coherences among the levels are preserved. For example, the ground states of Fig. 4.1 exhibit equal but opposite AC Stark shifts when the laser 61 is tuned halfway between the two transitions; the energy ofjg+idecreases and the energy of jg i increases, reducing the ground-state energy separation (a negative di erential light shift.) When preservation of coherence is paramount, quantum state manipulation with lasers is typically performed far from resonance, where the AC Stark shift is large but the spontaneous emission rate is small [79, 80]. Perhaps for this reason, the evolution of coherence in the opposite regime, where AC Stark shifts coexist with a large spontaneous emission rate, has not been exhaustively investigated. In this chapter we present measurements on a system near resonance where high rates of spontaneous emission and quantum jumps not only preserve but dy- namically modify a coherence, e ectively changing the sign of the ground-state di er- ential light shift of the con guration illustrated in Fig. 4.1. Our analysis generalizes a decoherence mechanism in Rayleigh scattering recently reported in the literature [81]. We show that it encompasses a coherent as well as incoherent aspect, depending on the size of the detunings. 4.2 Experimental system We measure di erential light shifts in a beam of cold 85Rb atoms passing through an optical cavity, using a ground-state interference technique [58]. A verti- cally polarized cavity mode, V, is weakly driven with coherent light at 780 nm and oriented with polarization vector parallel to a weak magnetic eld (see Fig. 4.2). 62 e + e - g - g + ? e ? g ? g ? e g? g? Figure 4.1: Four-level atom with unequal Zeeman shifts in the ground ( g) and excited ( e) states. A coherent eld is resonant with the un- shifted transition between ground and excited states (dashed lines) and drives the two transitions with equal coupling constants g. transi- tions (not shown) allow the spontaneous creation of an initial coherence between the ground states, starting from an unshifted and driven m = 0 to m0 = 0 transition. PBS Magnetic field BS APDs V H Figure 4.2: Schematic of the apparatus for measuring g(2)( ), with a single atom represented in green and the magnetic eld inside the cavity aligned to the exciting polarization; PBS: polarizing beam-splitter, BS: beam-splitter, APD: avalanche photodiode. 63 Photons in V excite transitions in the atoms between the F = 3 and F0 = 4 manifolds, and also cause light shifts when a detuning exists between a transition frequency and the laser frequency. A degenerate but horizontally polarized mode, H, collects photons emitted spontaneously on transitions between the same two manifolds; these photons have a bandwidth similar to that of the cavity, and can be distinguished from those of the drive by polarization. Detection of a photon from H spontaneously creates a coherent superposition among the magnetic ground states, and conditional detection of a second photon from H reads out the coherence. This setup enables a direct measurement of the Zeeman energy splitting of the ground states as the frequency of a long-lived quantum beat in the intensity autocorrela- tion function g(2)( ). The measurement is sensitive to a di erential light shift of the ground states, which appears as an intensity-dependent change in the beat fre- quency; and also to decoherence of the spontaneously-created superpositions, which appears as a broadening of the resonance in the power spectrum of g(2)( ). Examples of the beat signals taken from a full quantum Monte Carlo simulation of our system appear in Fig.4.3, and from recorded data in Fig.4.4. Both show the frequency of oscillation increasing as a function of average photon number in V, evidently the result of a light shift. However, this behavior is contrary to that expected from the AC Stark shift; our system roughly corresponds to the situation depicted in Fig. 4.1, since most atoms are in a superposition of the m = 1 ground states while the laser is resonant with the unshifted m = 0 to m0 = 0 transition. Stronger driving rapidly damps the coherence, bringing faster decay times to the oscillations. 64 ?(?s) 0.0 0.5 1.0 1.5 2.0 2.5 3.0 1.0 1.2 1.4 1.6 1.8 2.0 n=0.003 n=0.550 n=2.630 n=6.170 g (2) ( ? ) Figure 4.3: Quantum Monte Carlo simulation of intensity autocorrela- tion signal from H for four di erent photon numbers n in the driven mode V. (B) Composite experimental autocorrelation data as in (A) with amplitude color scaled to show detail. -2.0 -1.5 -1.0 -0.5 0.0 0.5 1.0 1.5 0 4 8 12 16 ? (?s) photons 1.0 1.1 1.2 1.3 1.4 1.5 2.0 g (2) (?) Figure 4.4: Composite experimental autocorrelation data as in Fig. 4.3 with amplitude color scaled to show detail. 65 4.3 Theoretical model We explain the origin of the light shifts and damping with the four-level atom shown in Fig. 4.1, using the quantum trajectory formalism [12]. Our analysis applies generally to any quantum system of this type and does not explicitly require a cavity, which essentially provides a spatial mode for the optical excitation and detection. The system is prepared at time t = 0 in the ground-state superposition j i= 1p 2(jg i+jg+i). (This occurs spontaneously in our apparatus after detection of a single photon from H.) The weak coherent drive couples the ground and excited states, with coe cients satisfying the equations: _ce = 2ce i ece +g cg ; _cg = i gcg g ce ; (4.1) where g is the Rabi frequency, and g and e are ground and excited-state Zeeman shifts. The ground-state superposition evolves in time as: j g(t)i= 1p2(ei( g+ AC)tjg i+e i( g+ AC)tjg+i); (4.2) where AC = g 2j j2 ( =2)2 + 2 (4.3) is the ground-state AC Stark shift obtained from diagonalizing Eqs. 4.1 to lowest order in g2j j2, and = e g is the magnitude of the laser detuning from each transition as depicted in Fig. 4.1. The di erential ground-state AC Stark shift is 2 AC, and it reduces the ground-state Zeeman splitting. 66 Under weak driving and with no quantum jumps, the amplitudes in Eq. 4.2 drive a steady-state superposition in the excited state: j e(t)i= g p2 ! ei( g+ AC)t =2 i je i+ e i( g+ AC)t =2 +i je+i ! : (4.4) The amplitudes of Eq. 4.4 oscillate at the ground-state splitting; the excited-state splitting enters only through the factors =2 i , which imply a phase shift. Inter- ference of the amplitudes in Eq. 4.4 produces the quantum beats which we measure. We show now that spontaneous emission in the form of quantum jumps adds both a coherent and incoherent contribution to the time evolution, changing the measured di erential light shift from 2 AC to 2 AC, while at the same time decohering the ground-state superposition. (Here for simplicity we consider only jumps occurring in between the measured emissions; additional jumps serve only to pump the coherence elsewhere in the manifold, changing Rabi frequencies and adding additional detunings.) At jump rate = g2j j2=(( =2)2 + 2), the driven dipole between ground and excited states turns o and the amplitudes of Eq. 4.4 are transferred to the ground states; in place of Eq. 4.2: j g(t)i= 1p2 2 4 =2 +i q ( =2)2 + 2 ei( g+ AC)tjg i+ =2 i q ( =2)2 + 2 e i( g+ AC)tjg+i 3 5; (4.5) where the ground-state superposition has acquired a phase advance. The process repeats for additional quantum jumps, so the phase advance accumulates over time. For a jump rate , we average thejg+ihg j-coherence against a Poisson distribution of mean t, to obtain the expectation value of the ground-state coherence: g+;g = e 2i( g+ AC)t12 1X n=0 ( =2 i )2 ( =2)2 + 2 !n ( t)n n! e t 67 = e 2i( g+ AC)t12e ( decoh+i2 jump)t: (4.6) The imaginary part of the exponent contains a new term, 2 jumpt, where jump = =2( =2)2 + 2 = 2g2j j2 =2 ( =2)2 + 2 !2 : (4.7) This term represents an additional frequency shift arising from the mean rate of phase accumulation from quantum jumps. It is opposite in sign to the AC Stark shift brought by the same driving laser (Eq. 4.3), and bigger than the AC Stark shift by a factor of two when the detuning is small (i.e. when =2). The net di erential ground-state light shift in this case becomes 2 light = 2( AC + jump) 2 AC; (4.8) with the sign of the di erential AC Stark shift e ectively reversed. The e ect requires at least the structure of a four-level atom, since jumps bring only an unob- servable overall phase shift in a two-level atom. The exponent in Eq. 4.6 also contains a damping term, decoht, which deco- heres the quantum beats at a rate decoh = 2 2 ( =2)2 + 2 = 2g 2j j2 ( =2)2 + 2 !2 : (4.9) The origin of the decoherence is phase di usion of the ground-state superposition, which accompanies the net phase drift that constitutes the frequency shift. Both aspects are expected from the stochastic nature of the jump process. Noting that decoh=(2 jump) = 2 = , we may speak of a well-resolved frequency shift when =2. When this condition is not met, a single jump can result in a phase 68 change of , smearing out the fringes over multiple measurements. In general, we require the product of the jump number uctuation with the phase shift per jump to be much less than in order to maintain coherence in the averaging process. Increasing the magnetic eld ( ) or the drive strength ( ) causes this product to increase; this is the origin of the beat decoherence presented in Chapter 3. The above analysis is valid for a four-level atom with constant coupling strength to a weak coherent drive. Our experimental system brings many additional com- plications. These include optical pumping within the full level structure of 85Rb, a nite interaction time with the laser and uctuations in coupling strength due to the atomic beam traversing the cavity mode, (weak) multi-atom and cavity e ects, and saturation behavior with strong driving. For these reasons we implement the theory as a full quantum Monte Carlo calculation in order to make quantitative comparison with our measurements [58]. 4.4 Data analysis We analyze the data by calculating the Fast Fourier Transform (FFT) am- plitude spectrum of each measured g(2)( ) and extracting the center frequency and FWHM of each ground-state beat resonance peak. Figure 4.5 shows the weak drive region for four di erent magnetic eld values, with linear ts to the data as dashed lines; the slope increase is approximately linear with magnetic eld (detuning). We extract the natural (non-light-shifted) Larmor frequency for each magnetic eld value by extrapolating this portion of data to n=0. This point corresponds to the 69 0.0 0.1 0.2 0.3 0.0 1.0 2.0 3.0 photons Shift (MHz) 5 G 2 G 8 G 11 G Figure 4.5: Center frequency of beat resonance at low photon number for four magnetic eld values. The dashed lines are linear ts to the data used to extrapolate the unshifted beat frequency. In all the following gures, error bars show one standard deviation con dence intervals of Voigt pro le ts to the resonance peaks. Experimentally determined photon numbers have an uncertainty of 20%. 70 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0 2 4 6 8 10 12 14 16 18 20 Freq uen cy shift (M Hz ) photons 4.9 Figure 4.6: Center frequency of beat resonance versus photon number in V from FFTs of measured data (red lled triangles) and simulations (blue open squares). The solid line is a t to a simple saturation func- tional form. unshifted beat frequency 2 g = 2 BBmag=3, where B=1.4 MHz/G is the Bohr magneton and Bmag is the magnitude of the magnetic eld. From this frequency we extract the value of Bmag and the detuning, = g=2. Figure 4.6 tracks the frequency shift (red lled triangles) for the data of Fig. 4.4, where a shift of 0 MHz corresponds to the natural Larmor frequency of 4.7 MHz at 5 G magnetic eld. The results taken from the quantum Monte Carlo calculations are plotted as blue empty squares. The di erence between the two is primarily due 71 to the experimental uncertainty in intracavity photon number. We nd that despite the complications, the data points in the high intensity region are described well by considering saturation of the jump rate as in a two-level atom, i.e. ! ( =2)(n=nsat)=(1 + n=nsat), where nsat 2=(8g2) is the saturation photon number for a two-level atom in a cavity. In the absence of optical pumping e ects, the e ective saturation photon number for an atom in the m = 1 ground states of our system should be nsat = 2=(3g2max) = 5:7 photons, where gmax is the peak value of g at the center of an antinode of the cavity Gaussian standing wave mode function, and a factor of 3/8 has been included to account for the averaging of g2 over the mode structure. To match the frequency shift data in Fig. 4.6, we use the simple saturation form for together with Eq. 4.8 at lowest order in : 2 light = Ashift (n=n 0 sat) 1 +n=n0sat : We expect the e ective saturation photon number n0sat to be larger than nsat = 5:7 as optical pumping puts the population in states with weaker coupling to the mode and larger detunings. The scale factor of Ashift is to include any other optical pumping and detuning e ects. The two-parameter least-squares t shown by the solid line in Fig. 4.6 gives Ashift = 0:81 0:01, n0sat = 9:2 0:2, with a reduced 2 = 0:32. The statistical error bars in all the gures are the 68.3% con dence intervals for the given parameter (center frequency or linewidth) from a least-squares t of each FFT resonance peak to a Gaussian pro le. Figure 4.7 shows the onset of decoherence with stronger driving, as extracted 72 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 0 2 4 6 8 10 12 14 16 18 20 F W H M ( M H z) photons i ii iii Figure 4.7: FWHM of beat resonance versus photon number in V for measured data (red lled circles) and simulations (blue open squares). The solid line is a t to the data for a simple saturation functional form. 73 0.0 0.5 1.0 4.0 4.5 5.0 5.5 Frequency (MHz) Amplitude (arb. units) i ii iii Figure 4.8: Three sample peaks from FFT amplitude spectra showing shifts and broadening with increased drive; the labels refer to the three points in Fig. 4.7 (i, ii, and iii). 74 from the widths of the resonance peaks from measured data (red lled circles) and from the quantum Monte Carlo simulation (blue empty squares). The zero-photon linewidth (0.23 MHz) arises from the nite transit time of the atoms traversing the Gaussian cavity mode. To match the data, we apply a similar saturation form as above to Eq. 4.9, at lowest order in : decoh = Awidth (n=n 0 sat)(2 = ) 2 1 +n=n0sat : We convolve this Lorentzian width with the Gaussian FWHM arising from the atom transits, 2 gauss = 2 0:23 MHz, in a Voigt pro le, using the numerical approximation: 2 voigt 0:5346(2 decoh) + q 0:2166(2 decoh)2 + (2 gauss)2: A two-parameter least-squares t to the composite width (solid line) gives Awidth = 1:31 0:03, n0sat = 14:2 0:5, with a reduced 2 = 1:6. Fig. 4.8 shows the broadening and shift of three sample resonance peaks from the experimental data. The results of the four-level model can be extended to include a detuning of the drive laser from the unshifted (zero magnetic eld) transition frequency (dashed lines in Fig. 4.1), denoted below as . The formulas for the di erential ground-state light shift and decoherence rate become: 2 light = 2( AC + jump) = 2g2j j2 ( =2) 2 2 + 2 [( =2)2 + ( + )2][( =2)2 + ( )2] (4.10) and decoh = 2g2j j2 2 [( =2)2 + ( + )2][( =2)2 + ( )2]: (4.11) 75 -0.05 0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 -10 -8 -6 -4 -2 0 2 4 6 8 10 Frequency (MHz) Laser detuning (MHz) Full width at half maximum Shift Figure 4.9: Sensitivity of light shift and linewidth to drive detuning from the central (zero magnetic eld) transition frequency, with 5 G magnetic eld and n =2 photons in V. Blue circles show shift in the center frequency from the zero-intensity value of 4.6 MHz, and red squares show the resonance FWHM. The solid lines are two-parameter ts to Eqs. 4.10 and 4.11. Error bars show one standard deviation con dence intervals of Gaussian pro le ts to the resonance peaks, and for the case of the linewidths are smaller than the squares used to plot the points. 76 Figure 4.9 presents the predicted and measured detuning dependence at a magnetic eld of 5 G. The photon number inside the cavity is kept constant at n = 2 photons, compensating for the o -resonant response of the cavity and atoms. The frequency shift (blue circles) decreases in both directions away from resonance. We t the data to Eq. 4.10 (blue line) with an amplitude scaling factor Ashift, while keeping the same value obtained before for n0sat, which e ectively scales g2 for the multi-level atomic structure. We nd Ashift = 0:71 0:01 with a reduced 2 = 7:6, due to a poor t at large detuning and some possible ne structure at small. The resonance peak width (red squares) presents a similar structure, reducing to the limiting value of 0.20 MHz (atom transit width) at a detuning of approx- imately 10 MHz, where the di erential phase shift from each quantum jump is approximately zero. The red line is a t to Eq. 4.11 convolved with transit broad- ening, with amplitude scaling as a free parameter. We nd Awidth = 0:64 0:01 for Eq. 4.11, with an order-of-magnitude discrepancy between the statistical error bars from the resonance peak ts and the uctuations about the theoretical form. 4.5 Comparisons Our results generalize the work of Uys et al. [81], who analyzed the e ect of Rayleigh scattering on the decoherence of trapped ion qubits. Their result is recovered from Eq. 4.11 in the limit =2, i.e. under conditions where jumps rarely occur but each carries a phase shift, fully destroying the ground-state coherence. More generally the character of the decoherence di ers depending on 77 the relative sizes of , , and =2. We have presented in the above an analysis of ground-state superpositions driven weakly by coherent light. A related situation was investigated by Cohen- Tannoudji [82, 83], who studied atoms excited by broadband thermal light. In this case, the ground-state superposition (induced by an external RF eld) jumps at random times completely to an excited-state superposition, with no coherent dipole between ground and excited states, and the Larmor frequency changes from 2 g to 2 e. Remaining in the excited state for an average time of 1= , the relative phase advance per jump is thus 2( e g)= = 2 = . This should be compared to the coherent driving case, where the relative phase advance per jump is 4 = (from Eq. 4.5 with =2.) The extra factor of two arises because the damping rate for the dipole is =2, in contrast to the damping rate for the population, which is ; however, there is no AC Stark shift for broadband light near resonance, so the net light shift in the two cases is exactly the same when = 0. This is a remarkable result, considering that the situations di er completely in detuning dependence and require entirely separate physical descriptions. We also note that the e ect with coherent drive is ve orders of magnitude larger than that observed previously with thermal light [82, 84, 85] and thus may need to be considered in some applications. 78 Chapter 5 Conclusions and Future Directions The experimental results reported in this thesis all rely on conditional mea- surements of light from one mode of an optical cavity that probe the atom-cavity interactions occurring in the other mode. Although the basic concept of the appa- ratus is simple, the complexity of the components{in particular the presence of an atomic beam, multi-level atoms, and a mode-matched source of background light{ has allowed observation of several surprising e ects that may not appear in simpler systems, e.g. in ones optimized for quantum information applications and precise state control. We demonstrated in Chapter 2 that a two-mode cavity QED system can func- tion as a high- delity detector of single atoms. Our method can operate with a wide range of photon uxes and works for atoms traveling as fast as 20 m/s across the mode. Because the cavity is frequency-selective and has a short lifetime (approxi- mately equal to the atomic excited-state lifetime), the method guarantees that the atom is in the F = 3 hyper ne ground state immediately after the detection of the second photon. The scheme is sensitive enough to work without strong coupling of the cavity and atoms, meaning that the detection volume can be kept large. We also demonstrated how autocorrelation measurements of the emitted light show the underlying dynamics of the atom as it traverses the Gaussian mode of the cavity 79 and encounters the standing wave. In Chapter 3 we presented observations of ground-state quantum beats in the spontaneous emission from a continuously driven atomic ensemble. Contrasting de- terministic manipulations [62, 63], we demonstrated the spontaneous creation and readout of ground-state coherence, where in the spirit of Forrester et al. [61], we re- trieve a hidden beat from the uctuations. Our theoretical treatment via a quantum Monte Carlo calculation, which is in good agreement with the measurements, de- composes the signal into a one-atom beat, a two-atom beat (interference of emission time order), and a homodyne beat due to interference with a drive photon mixed through birefringenece. In Chapter 4 we reported evidence of repeated quantum jumps modifying the spontaneously-generated coherence without destroying it, as a net drift guiding a stochastic phase di usion. We presented a microscopic quantum trajectory model that gives an intuitive explanation of the light shifts, and which also explains the state decoherence at large detuning reported in Ref. [81]. Moreover, our results sug- gest the tantalizing possibility of using dissipation as an additional tool for coherent control. Future upgrades to the experimental apparatus are planned. These include adding additional viewports to the cavity chamber to allow for better control of atomic state initialization as well as to give optical access to the cavity from the side; building a new cavity with slightly stronger coupling strength and much smaller absorptive photon losses; and implementing optical focusing of the atomic beam to increase the density of atoms in the cavity. Details are given in Ref. [18]. 80 We plan to study conditional quantum feedback in this system, similar to that rst demonstrated in Ref. [86], in which the conditional evolution of the atom-cavity system was frozen by application of a control pulse. Compared to the time scale of coherence in that experiment (set by the atom and cavity lifetimes), we have access to ground-state coherences that can last as long as the atom remains in the cavity, which relaxes some of the technical requirements for conditional control. We discuss details of one possible implementation in Ref. [87]. Other possibilities may arise with the additional optical access to the side of the cavity, such as measuring the correlation functions conditioned on a click directly from the (non-collective) atomic emission; this spontaneous emission arises from the same quantum jumps that give the measured light shifts. Another possibility is direct optical manipulation of the atomic states through the side of the cavity, using the output of one of the modes or some auxiliary control eld. The exibility of the system presents many opportunities, some of which are bound to yield even more surprising results. 81 Bibliography [1] P. R. Berman, ed., Cavity Quantum Electrodynamics (Academic Press, Boston, MA, 1994). [2] Q. A. Turchette, C. J. Hood, W. Lange, H. Mabuchi, and H. J. Kimble, \Mea- surement of Conditional Phase Shifts for Quantum Logic," Phys. Rev. Lett. 75, 4710 (1995). [3] J. I. Cirac, P. Zoller, H. J. Kimble, and H. Mabuchi, \Quantum State Transfer and Entanglement Distribution among Distant Nodes in a Quantum Network," Phys. Rev. Lett. 78, 3221 (1997). [4] K. M. Gheri, C. Saavedra, P. T orm a, J. I. Cirac, and P. Zoller, \Entanglement Engineering of One-Photon Wave Packets Using a Single-Atom Source," Phys. Rev. A 58, R2627 (1998). [5] C. Sch on, E. Solano, F. Verstraete, J. I. Cirac, and M. M. Wolf, \Sequen- tial Generation of Entangled Multiqubit States," Phys. Rev. Lett. 95, 110503 (2005). [6] A. Wallra , D. I. Schuster, A. Blais, L. Frunzio, R.-S. Huang, J. Majer, S. Kumar, S. M. Girvin, and R. J. Schoelkopf, \Strong Coupling of a Single Photon to a Superconducting Qubit Using Circuit Quantum Electrodynamics," Nature 431, 162 (2004). [7] T. Aoki, B. Dayan, E. Wilcut, W. P. Bowen, A. S. Parkins, T. J. Kippenberg, K. J. Vahala, and H. J. Kimble, \Observation of Strong Coupling between One Atom and a Monolithic Microresonator," Nature 443, 671 (2006). [8] S. Gleyzes, S. Kuhr, C. Guerlin, J. Bernu, S. Del eglise, U. B. Ho , M. Brune, J.-M. Raimond and S. Haroche, \Quantum Jumps of Light Recording the Birth and Death of a Photon in a Cavity," Nature 446, 297 (2007). [9] M. L. Terraciano, \Cross-Correlations and Entanglement in Cavity QED," Ph. D. Thesis, U of Maryland, College Park, 2006. http://hdl.handle.net/1903/3767 [10] M. L. Terraciano, R. Olson Knell, D. L. Freimund, L. A. Orozco, J. P. Clemens, and P. R. Rice, \Enhanced Spontaneous Emission into the Mode of a Cavity QED System," Opt. Lett. 32, 982 (2007). 82 [11] E. T. Jaynes and F. W. Cummings, \Comparison of Quantum and Semiclassical Radiation Theories with Application to the Beam Maser," Proc. IEEE 51, 89 (1963). [12] H. J. Carmichael, An Open Systems Approach to Quantum Optics, Lecture Notes in Physics 18 (Springer-Verlag, Berlin, 1993). [13] L. A. Lugiato, \Theory of Optical Bistability," in Progress in Optics, XXI, 69 (North-Holland, Amsterdam, 1984). [14] A. T. Rosenberger, L. A. Orozco, H. J. Kimble, and P. D. Drummond, \Absorp- tive Optical Bistability in Two-State Atoms," Phys. Rev. A 43, 6284 (1991). [15] C. J. Hood, M. S. Chapman, T. W. Lynn, and H. J. Kimble, \Real-Time Cavity QED with Single Atoms," Phys. Rev. Lett. 80, 4157 (1998). [16] P. M unstermann, T. Fischer, P. Maunz, P. W. H. Pinkse, and G. Rempe, \Dynamics of Single-Atom Motion Observed in a High-Finesse Cavity," Phys. Rev. Lett 82, 3791 (1999). [17] G. R. Guth ohrlein, M. Keller, K. Hayasaka, W. Lange, and H. Walther, \A Single Ion as a Nanoscopic Probe of an Optical Field," Nature 414, 49 (2001). [18] A. D. Cimmarusti , J. A. Crawford , D. G. Norris, and L. A. Orozco, \Enhance- ments to Cavity Quantum Electrodynamics System," Rev. Mex. Fis., TBD (2011). [19] Z. T. Lu, K. L. Corwin, M. J. Renn, M. H. Anderson, E. A. Cornell, and C. E. Wieman, \Low-Velocity Intense Source of Atoms from a Magneto-Optical Trap," Phys. Rev. Lett. 77, 3331 (1996). [20] R. W. P. Drever, J. L. Hall, F. V. Kowalski, J. Hough, G. M. Ford, A. J. Munley, and H. Ward, \Laser Phase and Frequency Stabilization Using an Optical Resonator," Appl. Phys. B 31, 97 (1983). [21] K. L. Corwin, Z. T. Lu, C. F. Hand, R. J. Epstein, and C. E. Wieman, \Frequency-Stabilized Diode Laser with the Zeeman Shift in an Atomic Va- por," Appl. Opt. 37, 3295 (1998). [22] L. Isenhower, E. Urban, X. L. Zhang, A. T. Gill, T. Henage, T. A. Johnson, T. G. Walker, and M. Sa man, \Demonstration of a Neutral Atom Controlled- NOT Quantum Gate," Phys. Rev. Lett. 104, 010503 (2010). 83 [23] T. Wilk, A. Gatan, C. Evellin, J. Wolters, Y. Miroshnychenko, P. Grangier, and A. Browaeys, \Entanglement of Two Individual Neutral Atoms Using Rydberg Blockade," Phys. Rev. Lett. 104, 010502 (2010). [24] N. Schlosser, G. Reymond, I. Protsenko, and P. Grangier, \Sub-Poissonian Loading of Single Atoms in a Microscopic Dipole Trap, Nature 411, 1024 (2001). [25] S. Kuhr, W. Alt, D. Schrader, M. Mueller, V. Gomer, and D. Meschede, \De- terministic Delivery of a Single Atom," Science 293, 278 (2001). [26] M. P. A. Jones, J. Beugnon, A. Ga etan, J. Zhang, G. Messin, A. Browaeys, and P. Grangier, \Fast Quantum State Control of a Single Trapped Neutral Atom," Phys. Rev. A 75, 040301 (R) (2007). [27] D. A. Lewis, J. F. Tonn, S. L. Kaufman, and G. W. Greenlees, \Photon-Burst Method in High-Resolution Laser Spectroscopy," Phys. Rev. A 19, 1580 (1979). [28] K. Peck, L. Stryer, A. N. Glazer, and R. A. Mathies, \Single Molecule Fluores- cence Detection: Autocorrelation Criterion and Experimental Realization with Phycoerythrin," Proc. Natl. Acad. Sci. USA 86, 4087 (1989). [29] Y.-H. Lee, R. G. Maus, B. W. Smith, and D. Winefordner, \Laser-Induced Fluorescence Detection of a Single Molecule in a Capillary," Anal. Chem. 66, 4142 (1994). [30] T. Bondo, M. Hennrich, T. Legero, G. Rempe, and A. Kuhn, \Time-Resolved and State-Selective Detection of Single Freely Falling Atoms, Opt. Comm. 264, 271 (2006). [31] E. M. Purcell, Phys. Rev. 69, 681 (1946). [32] I. Teper, Y.-J. Lin, and V. Vuletic c, \Resonator-Aided Single-Atom Detection on a Microfabricated Chip," Phys. Rev. Lett. 97, 023002 (2006). [33] S. Nussmann, K. Murr, M. Hijlkema, B. Weber, A. Kuhn, and G. Rempe, \Vacuum-Stimulated Cooling of Single Atoms in Three Dimensions," Nature Phys. 1, 122 (2005). [34] K. M. Fortier, S. Y. Kim, M. J. Gibbons, P. Ahmadi, and M. S. Chapman, \Deterministic Loading of Individual Atoms to a High-Finesse Optical Cavity," Phys. Rev. Lett. 98, 233601 (2007). 84 [35] H. Mabuchi, Q. A. Turchette, M. S. Chapman, and H. J. Kimble, \Real-Time Detection of Individual Atoms Falling through a High-Finesse Optical Cavity," Opt. Lett. 21, 1393 (1996). [36] A. Ottl, S. Ritter, M. K ohl, and T. Esslinger, \Correlations and Counting Statistics of an Atom Laser," Phys. Rev. Lett. 95, 090404 (2005). [37] M. Trupke, J. Goldwin, B. Darqui e, G. Dutier, S. Eriksson, J. Ashmore, and E. A. Hinds, \Atom Detection and Photon Production in a Scalable, Open, Optical Microcavity," Phys. Rev. Lett. 99, 063601 (2007). [38] A. D. Boozer, A. Boca, R. Miller, T. E. Northup, and H. J. Kimble, \Cooling to the Ground State of Axial Motion for One Atom Strongly Coupled to an Optical Cavity," Phys. Rev. Lett. 97, 083602 (2006). [39] T. Puppe, I. Schuster, A. Grothe, A. Kubanek, K. Murr, P. W. H. Pinkse, and G. Rempe, \Trapping and Observing Single Atoms in a Blue-Detuned Intracavity Dipole Trap," Phys. Rev. Lett. 99, 013002 (2007). [40] M. L. Terraciano, R. Olson Knell, D. G. Norris, J. Jing, A. Fern andez, and L. A. Orozco, \Photon Burst Detection of Single Atoms in an Optical Cavity," Nat. Phys. 5, 480 (2009). [41] K. M. Birnbaum, A. Boca, R. Miller, A. D. Boozer, T. E. Northup, and H. J. Kimble, \Photon Blockade in an Optical Cavity with One Trapped Atom," Nature 436, 87 (2005). [42] D. Budker, W. Gawlik, D. F. Kimball, S. M. Rochester, V. V. Yashchuk, and A. Weis, \Resonant Nonlinear Magneto-Optical E ects in Atoms," Rev. Mod. Phys. 74, 1153 (2002). [43] L. Mandel and E. Wolf, Optical Coherence and Quantum Optics (Cambridge University Press, New York, 1995). [44] H. J. Kimble, M. Dagenais, and L. Mandel, \Multiatom and Transit-Time Ef- fects on Photon-Correlation Measurements in Resonance Fluorescence," Phys. Rev. A 18, 201 (1978). [45] R. Hanbury Brown and R. Twiss, \A Test of a New Type of Stellar Interfer- ometer on Sirius," Nature 178, 1046 (1956). [46] H. J. Carmichael, P. Drummond, P. Meystre, and D. F. Walls, \Intensity Corre- lations in Resonance Fluorescence with Atomic Number Fluctuations," J. Phys. A: Math. Gen. 11 L121 (1978). 85 [47] H. J. Carmichael, P. D. Drummond, D. F. Walls, and P. Meystre, \Photon Antibunching in Resonance Fluorescence in the Presence of Atomic Number Fluctuations," Optica Acta 27, 581 (1980). [48] G. T. Foster, S. L. Mielke, and L. A. Orozco, \Intensity Correlations of a Noise-Driven Diode Laser," J. Opt. Soc. Am. B 15, 2646 (1998). [49] M. Hennrich, A. Kuhn, and G. Rempe, \Transition from Antibunching to Bunching in Cavity QED," Phys. Rev. Lett. 94, 053604 (2005). [50] G. Labeyrie, C. Miniatura, and R. Kaiser, \Large Faraday Rotation of Resonant Light in a Cold Atomic Cloud," Phys. Rev. A 64, 033402 (2001). [51] M. Wilzbach, D. Heine, S. Groth, X. Liu, T. Raub, B. Hessmo, and J. Schmied- mayer, \Simple Integrated Single-Atom Detector," Opt. Lett. 34, 259 (2009). [52] P. D. Drummond, \Optical Bistability in a Radially Varying Mode," IEEE J. Quant. Electron. 17, 301 (1981). [53] H. J. Carmichael and B. C. Sanders, \Multiatom E ects in Cavity QED with Atomic Beams," Phys. Rev. A 60, 2497 (1999). [54] G. Breit, \Quantum Theory of Dispersion (Continued). Parts VI and VII," Rev. Mod. Phys. 5, 91 (1933). [55] E. B. Aleksandrov, \Luminescence Beats Induced by Pulsed Excitation of Co- herent States," Opt. Spectrosc. 17, 522 (1964). [56] J. N. Dodd, W. J. Sandle, and D. Zissermann, \Study of Resonance Fluores- cence in Cadmium: Modulation E ects and Lifetime Measurements," Proc. Phys. Soc. 92, 497 (1967). [57] A. Aspect, J. Dalibard, P. Grangier, and G. Roger, \Quantum Beats in Con- tinually Excited Atomic Cascades," Opt. Commun. 49, 429 (1984). [58] D. G. Norris, L. A. Orozco, P. Barberis-Blostein, and H. J. Carmichael, \Ob- servation of Ground-State Quantum Beats in Atomic Spontaneous Emission," Phys. Rev. Lett. 105, 123602 (2010). [59] W. W. Chow, M. O. Scully, and J. O. Stoner Jr, \Quantum-Beat Phenomena Described by Quantum Electrodynamics and Neoclassical Theory," Phys. Rev. A 11, 1380 (1975). 86 [60] R. M. Herman, H. Grotch, R. Kornblith, and J. H. Eberly, \Quantum Elec- trodynamic and Semiclassical Interference E ects in Spontaneous Radiation," Phys. Rev. A 11, 1389 (1975). [61] A. T. Forrester, R. A. Gudmundsen, and P. O. Johnson, \Photoelectric Mixing of Incoherent Light," Phys. Rev. 99, 1691 (1955). [62] T. Wilk, S. C. Webster, A. Kuhn, and G. Rempe, \Single-Atom Single-Photon Quantum Interface," Science 317, 488 (2007). [63] B. Weber, H. P. Specht, T. M uller, J. Bochmann, M. M ucke, D. L. Moehring, and G. Rempe, \Photon-Photon Entanglement with a Single Trapped Atom," Phys. Rev. Lett. 102, 030501 (2009). [64] M. Schubert, I. Siemers, R. Blatt, W. Neuhauser, and P. E. Toschek, \Transient Internal Dynamics of a Multilevel Ion," Phys. Rev. A 52, 2994 (1995). [65] J. Javanainen, \E ect of State Superpositions Created by Spontaneous Emis- sion on Laser-Driven Transitions," Europhys. Lett. 17, 407 (1992). [66] A. K. Patnaik and G. S. Agarwal, \Cavity-Induced Coherence E ects in Spon- taneous Emissions from Preselection of Polarization," Phys. Rev. A 59, 3015 (1999). [67] S. E. Economou, R.-B. Liu, L. J. Sham, and D. G. Steel, \Uni ed Theory of Consequences of Spontaneous Emission in a System," Phys. Rev. B 71, 195327 (2005). [68] M. Gurudev Dutt, J. Cheng, B. Li, X. Xu, X. Li, P. R. Berman, D. G. Steel, A. S. Bracker, D. Gammon, S. E. Economou, R.-B. Liu, and L. J. Sham, \Stimulated and Spontaneous Optical Generation of Electron Spin Coherence in Charged GaAs Quantum Dots," Phys. Rev. Lett. 94, 227403 (2005). [69] A. G. Zajonc, \Proposed Quantum-Beats, Quantum-Eraser Experiment," Phys. Lett. A 96, 61 (1983). [70] M. O. Scully and K. Dr uhl, \Quantum Eraser: A Proposed Photon Correla- tion Experiment Concerning Observation and \Delayed Choice" in Quantum Mechanics," Phys. Rev. A 25, 2208 (1982). [71] L. Horvath and H. J. Carmichael, \E ect of Atomic Beam Alignment on Photon Correlation Measurements in Cavity QED," Phys. Rev. A 76, 043821 (2007). 87 [72] L. Allen and J. H. Eberly, Optical Resonance and Two-Level Atoms (John Wiley and Sons, Inc., New York, 1975). [73] H. Dehmelt, Bull. Am. Phys. Soc. 20, 60 (1975). [74] R. J. Cook, and H. J. Kimble, \Possibility of Direct Observation of Quantum Jumps," Phys. Rev. Lett. 54, 1023 (1984). [75] W. Nagourney, J. Sandberg, and H. Dehmelt, \Shelved Optical Electron Am- pli er: Observation of Quantum Jumps," Phys. Rev. Lett. 56, 2797 (1986). [76] J. C. Bergquist, R. G. Hulet, W. M. Itano, and D. J. Wineland, \Observation of Quantum Jumps in a Single Atom," Phys. Rev. Lett. 57, 1699 (1986). [77] T. Sauter, W. Neuhauser, R. Blatt, and P. E. Toschek, \Observation of Quan- tum Jumps," Phys. Rev. Lett. 57, 1696 (1986). [78] W. Happer, \Light Propagation and Light Shifts in Optical Pumping Experi- ments," Prog. Quant. Electron. 1, 51 (1970). [79] R. A. Cline, J. D. Miller, M. R. Matthews, and D. J. Heinzen, \Spin Relaxation of Optically Trapped Atoms by Light Scattering," Opt. Lett. 19, 207 (1994). [80] R. Ozeri, C. Langer, J. D. Jost, B. DeMarco, A. Ben-Kish, B. R. Blakestad, J. Britton, J. Chiaverini, W. M. Itano, D. B. Hume, D. Leibfried, T. Rosenband, P. O. Schmidt, and D. J. Wineland, \Hyper ne Coherence in the Presence of Spontaneous Photon Scattering," Phys. Rev. Lett. 95, 030403 (2005). [81] H. Uys, M. J. Biercuk, A. P. VanDevender, C. Ospelkaus, D. Meiser, R. Ozeri, and J. J. Bollinger, \Decoherence Due to Elastic Rayleigh Scattering, Phys. Rev. Lett. 105, 200401 (2010). [82] C. Cohen-Tannoudji, Ann. Phys. (Paris), 13e s erie 7, 423 (1962). [83] A. Kastler, \Displacement of Energy Levels of Atoms by Light," J. Opt. Soc. Am. 53, 902 (1963). [84] B. R. Bulos, A. Marshall, and W. Happer, \Light Shifts Due to Real Transitions in Optically Pumped Alkali Atoms," Phys. Rev. A 4, 51 (1971). [85] J. C. Lehmann, \Nuclear Orientation of Cadmium111 by Optical Pumping with the Resonance Line 51S0-51P1," Phys. Rev. 178, 153 (1969). 88 [86] W. P. Smith, J. E. Reiner, L. A. Orozco, S. Kuhr, and H. M. Wiseman, \Cap- ture and Release of a Conditional state of a Cavity QED System by Quantum Feedback," Phys. Rev. Lett. 89, 133601 (2002). [87] P. Barberis-Blostein, D. G. Norris, L. A. Orozco, and H. J. Carmichael, \From Quantum Feedback to Probabilistic Error Correction: Manipulation of Quan- tum Beats in Cavity QED," New J. of Phys. 12, 023002 (2010). 89