12.07.2015 Views

Single Atoms on Demand for Cavity QED Experiments

Single Atoms on Demand for Cavity QED Experiments

Single Atoms on Demand for Cavity QED Experiments

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<str<strong>on</strong>g>Single</str<strong>on</strong>g> atoms <strong>on</strong> demand<strong>for</strong> cavity <strong>QED</strong> experimentsIgor DotsenkoB<strong>on</strong>n 2007


<str<strong>on</strong>g>Single</str<strong>on</strong>g> atoms <strong>on</strong> demand<strong>for</strong> cavity <strong>QED</strong> experimentsDissertati<strong>on</strong>zurErlangung des Doktorgrades (Dr. rer. nat.)derMathematisch-Naturwissenschaftlichen FakultätderRheinischen Friedrich-Wilhelms-Universität B<strong>on</strong>nvorgelegt v<strong>on</strong>Igor DotsenkoausCherkassy (Ukraine)B<strong>on</strong>n 2007


Angefertigt mit Genehmigung der Mathematisch-Naturwissenschaftlichen Fakultätder Rheinischen Friedrich-Wilhelms-Universität B<strong>on</strong>n1. Referent: Prof. Dr. Dieter Meschede2. Referent: Prof. Dr. Le<strong>on</strong>id YatsenkoTag der Promoti<strong>on</strong>: 22.05.2007Diese Dissertati<strong>on</strong> ist auf dem Hochschulschriftenserver der ULB B<strong>on</strong>nhttp://hss.ulb.uni-b<strong>on</strong>n.de/diss_<strong>on</strong>line/ elektr<strong>on</strong>isch publiziert


Abstract<strong>Cavity</strong> quantum electrodynamics (cavity <strong>QED</strong>) describes electromagnetic fields in ac<strong>on</strong>fined space and the radiative properties of atoms in such fields. The simplest example ofsuch system is a single atom interacting with <strong>on</strong>e mode of a high-finesse res<strong>on</strong>ator. Besidesobservati<strong>on</strong> and explorati<strong>on</strong> of fundamental quantum mechanical effects, this system bearsa high potential <strong>for</strong> applicati<strong>on</strong>s in quantum in<strong>for</strong>mati<strong>on</strong> science such as, e.g., quantumlogic gates, quantum communicati<strong>on</strong> and quantum teleportati<strong>on</strong>.In this thesis I present an experiment <strong>on</strong> the deterministic coupling of a single neutralatom to the mode of a high-finesse optical res<strong>on</strong>ator. In Chapter 1 I describe our basictechniques <strong>for</strong> trapping and observing single cesium atoms. As a source of single atoms weuse a high-gradient magneto-optical trap, which captures the atoms from background gasin a vacuum chamber and cools them down to millikelvin temperatures. The atoms arethen transferred without loss into a standing-wave dipole trap, which provides a c<strong>on</strong>servativepotential required <strong>for</strong> experiments <strong>on</strong> atomic coherence such as quantum in<strong>for</strong>mati<strong>on</strong>processing and metrology <strong>on</strong> trapped atoms. Moreover, shifting the standing-wave patternallows us to deterministically transport the atoms (Chapter 2). In combinati<strong>on</strong> with n<strong>on</strong>destructivefluorescence imaging of individual trapped atoms, this enables us to c<strong>on</strong>troltheir positi<strong>on</strong> with submicrometer precisi<strong>on</strong> over several millimeters al<strong>on</strong>g the dipole trap.The cavity <strong>QED</strong> system can distinctly display quantum behaviour in the so-calledstr<strong>on</strong>g coupling regime, i.e., when the coherent atom-cavity coupling rate dominates dissipati<strong>on</strong>in the system. This sets the main requirements <strong>on</strong> the res<strong>on</strong>ator’s properties: smallmode volume and high finesse. Chapter 3 is devoted to the manufacturing, assembling,and testing of an ultra-high finesse optical Fabry-Perot res<strong>on</strong>ator, stabilized to the atomictransiti<strong>on</strong>. In Chapter 4 I present the transportati<strong>on</strong> of single atoms into the cavity andtheir coupling to the cavity mode. The str<strong>on</strong>g coupling manifests itself in a str<strong>on</strong>g reducti<strong>on</strong>of the cavity transmissi<strong>on</strong> probed by a weak external laser. The atoms remaintrapped and coupled to the cavity mode <strong>for</strong> several sec<strong>on</strong>ds until we move them out of thecavity <strong>for</strong> final analysis of their number and positi<strong>on</strong>.Parts of this thesis have been published in the following journal articles:1. Y. Miroshnychenko, D. Schrader, S. Kuhr, W. Alt, I. Dotsenko, M. Khudaverdyan,A. Rauschenbeutel, and D. Meschede, C<strong>on</strong>tinued imaging of the transport of a singleneutral atom, Optics Express 11, 3498 (2003)2. I. Dotsenko, W. Alt, M. Khudaverdyan, S. Kuhr, D. Meschede, Y. Miroshnychenko,D. Schrader, and A. Rauschenbeutel, Submicrometer positi<strong>on</strong> c<strong>on</strong>trol ofsingle trapped neutral atoms, Physical Review Letters 95, 033002 (2005)


VIII3. L. Förster, W. Alt, I. Dotsenko, M. Khudaverdyan, Y. Miroshnychenko,D. Meschede, S. Reick, and A. Rauschenbeutel, Number-triggered loading and collisi<strong>on</strong>alredistributi<strong>on</strong> of neutral atoms in a standing wave dipole trap, New Journalof Physics 8, 259 (2006)4. M. Khudaverdyan, W. Alt, I. Dotsenko, T. Kampschulte, K. Lenhard, A. Rauschenbeutel,S. Reick, K. Schörner, A. Widera and D. Meschede, C<strong>on</strong>trolled inserti<strong>on</strong>and retrieval of atoms coupled to a high-finesse optical res<strong>on</strong>ator, New Journalof Physics 10, 073023 (2008)


C<strong>on</strong>tentsAbstractVIIIntroducti<strong>on</strong> 11 Trapping and observing single atoms 31.1 Introducti<strong>on</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31.2 A single-atom magneto-optical trap . . . . . . . . . . . . . . . . . . . . . . . 41.2.1 Principle of operati<strong>on</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . 41.2.2 Experimental setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51.2.3 Fluorescence detecti<strong>on</strong> of single atoms . . . . . . . . . . . . . . . . . 81.3 A standing-wave optical dipole trap . . . . . . . . . . . . . . . . . . . . . . 91.3.1 Dipole potential . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101.3.2 Multi-level cesium atom . . . . . . . . . . . . . . . . . . . . . . . . . 111.3.3 Standing-wave trap . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111.3.4 Experimental setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131.4 Loading of the DT with atoms . . . . . . . . . . . . . . . . . . . . . . . . . 141.5 Imaging single atoms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151.6 C<strong>on</strong>clusi<strong>on</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172 Submicrometer positi<strong>on</strong> c<strong>on</strong>trol of single atoms 192.1 Introducti<strong>on</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 192.2 Measurement of the positi<strong>on</strong> of an atom . . . . . . . . . . . . . . . . . . . . 202.3 Measurement of the separati<strong>on</strong> between two atoms . . . . . . . . . . . . . . 252.4 The optical c<strong>on</strong>veyor belt . . . . . . . . . . . . . . . . . . . . . . . . . . . . 302.5 Active positi<strong>on</strong> c<strong>on</strong>trol . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 332.6 Calibrati<strong>on</strong> of the image scale . . . . . . . . . . . . . . . . . . . . . . . . . . 352.7 C<strong>on</strong>clusi<strong>on</strong> and discussi<strong>on</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . 393 The high-finesse optical res<strong>on</strong>ator 413.1 An optical res<strong>on</strong>ator <strong>for</strong> cavity <strong>QED</strong> experiments . . . . . . . . . . . . . . . 413.2 High-reflectivity mirrors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 443.2.1 Mirror design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 443.2.2 Portable cavity ring-down setup . . . . . . . . . . . . . . . . . . . . 453.2.3 Reflectivity measurement . . . . . . . . . . . . . . . . . . . . . . . . 473.2.4 Mirror birefringence . . . . . . . . . . . . . . . . . . . . . . . . . . . 503.2.5 <strong>Cavity</strong> ringing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 533.3 High-finesse cavity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 563.3.1 Assembly . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 573.3.2 Frequency stabilizati<strong>on</strong> . . . . . . . . . . . . . . . . . . . . . . . . . 583.3.3 Detecti<strong>on</strong> of the cavity transmissi<strong>on</strong> . . . . . . . . . . . . . . . . . . 613.3.4 Characterizati<strong>on</strong> of the cavity . . . . . . . . . . . . . . . . . . . . . . 64IX


List of Figures1.1 Principle of operati<strong>on</strong> of a magneto-optical trap . . . . . . . . . . . . . . . . 51.2 Side view of an optical table with the vacuum system and the optical traps 61.3 Hyperfine structure of the cesium D 2 transiti<strong>on</strong> . . . . . . . . . . . . . . . . 71.4 Schematic experimental setup of the atom traps . . . . . . . . . . . . . . . . 81.5 APD signal of the MOT fluorescence . . . . . . . . . . . . . . . . . . . . . . 91.6 Three-dimensi<strong>on</strong>al view of the standing-wave dipole potential . . . . . . . . 121.7 Histograms of fluorescence counts from the MOT . . . . . . . . . . . . . . . 141.8 Images of atoms in the optical traps . . . . . . . . . . . . . . . . . . . . . . 162.1 Determinati<strong>on</strong> of the positi<strong>on</strong> of a single trapped atom . . . . . . . . . . . . 212.2 Measurement of the phase variati<strong>on</strong>s of the standing-wave dipole trap . . . 242.3 Determinati<strong>on</strong> of the distance between atoms . . . . . . . . . . . . . . . . . 262.4 Histogram of the measured distances modulo λ/2 . . . . . . . . . . . . . . . 272.5 Cumulative distributi<strong>on</strong> of separati<strong>on</strong>s between atoms in the dipole trap . . 292.6 Histogram of the averaged atomic separati<strong>on</strong>s modulo λ/2 . . . . . . . . . . 302.7 Time dependence of the main transportati<strong>on</strong> parameters . . . . . . . . . . . 312.8 Experimental setup of the optical c<strong>on</strong>veyor belt . . . . . . . . . . . . . . . . 322.9 Absolute positi<strong>on</strong> c<strong>on</strong>trol of single trapped atoms . . . . . . . . . . . . . . . 342.10 Calibrati<strong>on</strong> of the image scale using atomic separati<strong>on</strong>s modulo λ/2 . . . . 372.11 Fourier trans<strong>for</strong>mati<strong>on</strong> of the distributi<strong>on</strong> of atomic separati<strong>on</strong>s . . . . . . . 383.1 Basic elements and rates in a cavity <strong>QED</strong> system . . . . . . . . . . . . . . . 423.2 High-reflectivity mirror <strong>for</strong> cavity <strong>QED</strong> . . . . . . . . . . . . . . . . . . . . 443.3 Portable cavity ring-down setup . . . . . . . . . . . . . . . . . . . . . . . . . 483.4 <strong>Cavity</strong> ring-down signal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 493.5 Birefringence of high-reflectivity mirrors . . . . . . . . . . . . . . . . . . . . 513.6 <strong>Cavity</strong> ringing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 543.7 Theoretical cavity ringing signal . . . . . . . . . . . . . . . . . . . . . . . . 553.8 <strong>Cavity</strong> ringing and birefringence . . . . . . . . . . . . . . . . . . . . . . . . 553.9 Schematic view of the cavity integrated into the atom-trapping experiment 563.10 <strong>Cavity</strong> holder in the vacuum setup . . . . . . . . . . . . . . . . . . . . . . . 573.11 Assembled cavity integrated into the vacuum glass cell . . . . . . . . . . . . 583.12 Frequency stabilizati<strong>on</strong> of the high-finesse cavity . . . . . . . . . . . . . . . 593.13 Schematic optical cavity setup . . . . . . . . . . . . . . . . . . . . . . . . . 62XI


XIILIST OF FIGURES3.14 <strong>Cavity</strong> ring-down of the high-finesse cavity . . . . . . . . . . . . . . . . . . . 663.15 Birefringent modes of the cavity . . . . . . . . . . . . . . . . . . . . . . . . 673.16 Lifetime of trapped atoms placed between the cavity mirrors . . . . . . . . 753.17 Probe transmissi<strong>on</strong> and compensati<strong>on</strong> of the cavity thermal drifts . . . . . . 763.18 Compensating <strong>for</strong> the cavity thermal drifts . . . . . . . . . . . . . . . . . . 773.19 Re-thermalizati<strong>on</strong> of the locked cavity during an experimental run . . . . . 794.1 Eigenstates of the res<strong>on</strong>ant atom-cavity system . . . . . . . . . . . . . . . . 834.2 Vacuum Rabi splitting from steady-state soluti<strong>on</strong> of the master equati<strong>on</strong> . . 864.3 Energy levels of the coupled atom-cavity system . . . . . . . . . . . . . . . 874.4 Expected cavity transmissi<strong>on</strong> <strong>for</strong> an atom oscillating al<strong>on</strong>g the cavity axis . 894.5 Locating the cavity mode using optical pumping by the probe laser . . . . . 924.6 Locating the cavity mode by detecting atom losses . . . . . . . . . . . . . . 944.7 Locating the cavity mode by inducing the losses of many atoms . . . . . . . 964.8 Aligning the DT <strong>on</strong>to the cavity mode using many atom losses . . . . . . . 984.9 Vanishing cavity transmissi<strong>on</strong> <strong>for</strong> several atoms in the cavity mode . . . . . 1004.10 Influence of lock laser and repumping laser <strong>on</strong> transmissi<strong>on</strong> decrease . . . . 1014.11 Storage time in the cavity depending <strong>on</strong> probe laser power . . . . . . . . . . 1024.12 <strong>Cavity</strong> transmissi<strong>on</strong> <strong>for</strong> different numbers of coupled atoms . . . . . . . . . 1034.13 Storage time in the cavity depending <strong>on</strong> the atom number . . . . . . . . . . 1044.14 Multiple transport of a large number of atoms into the cavity . . . . . . . . 1054.15 Light shift of the atomic excited and ground states . . . . . . . . . . . . . . 1074.16 The rapid changes of the atom-cavity coupling . . . . . . . . . . . . . . . . 1094.17 Hopping of atoms al<strong>on</strong>g the DT . . . . . . . . . . . . . . . . . . . . . . . . . 1104.18 Dipole potential <strong>for</strong>med by crossed DT and lock laser . . . . . . . . . . . . 1124.19 <strong>Cavity</strong> transmissi<strong>on</strong> <strong>for</strong> different powers of the lock laser . . . . . . . . . . . 1134.20 <strong>Cavity</strong> transmissi<strong>on</strong> <strong>for</strong> different orientati<strong>on</strong>s of the guiding magnetic field . 116A.1 Level scheme of the first excited states in 133 Cs . . . . . . . . . . . . . . . . 124B.1 Theoretical cavity ringing signal . . . . . . . . . . . . . . . . . . . . . . . . 126


List of Tables2.1 Sources of noise influencing the precisi<strong>on</strong> of the positi<strong>on</strong> detecti<strong>on</strong> . . . . . 253.1 Fit results <strong>for</strong> the oscillating CRD signal . . . . . . . . . . . . . . . . . . . . 523.2 Measured mirror losses and mode coupling efficiency . . . . . . . . . . . . . 703.3 Properties of the high-finesse optical cavity . . . . . . . . . . . . . . . . . . 71A.1 Some physical properties of the 133 Cs atom . . . . . . . . . . . . . . . . . . 123A.2 Dipole matrix elements <strong>for</strong> the 133 Cs atom . . . . . . . . . . . . . . . . . . . 124XIII


Introducti<strong>on</strong>In 1946 Edward Purcell noticed that the rate of sp<strong>on</strong>taneous emissi<strong>on</strong> of an atom canbe significantly enhanced by coupling it to an electrical circuit res<strong>on</strong>ant with the atomicradio-frequency transiti<strong>on</strong> [1]. This publicati<strong>on</strong> gave birth to a new field of research –cavity quantum electrodynamics (C<strong>QED</strong>), which describes the electromagnetic field in ac<strong>on</strong>fined space and the radiative properties of atoms in such field. Obviously, <strong>on</strong>e of thesimplest systems in the scope of C<strong>QED</strong> is a single atom interacting with a single phot<strong>on</strong> ina cavity. The theory describing this system was first developed by Jaynes and Cummingsin 1963 [2], providing the most basic model to C<strong>QED</strong> experiments.The main obstacles in the experimental realizati<strong>on</strong> of a coupled atom-cavity system aresp<strong>on</strong>taneous emissi<strong>on</strong> of the atom and damping of the cavity field. Thus, the most relevantregime <strong>for</strong> observing quantum phenomena of the atom-field dynamics is reached whenthe strength of the coherent atom-field coupling exceeds these dissipati<strong>on</strong>s. Fortunately,the tremendous progress in c<strong>on</strong>trolling single atoms as well as in manufacturing highfinessecavities accomplished over the last 15 years has allowed experimentalists to achievethe str<strong>on</strong>g coupling with single atoms in cavities in two spectral domains. At microwavefrequencies highly excited Rydberg atoms are coupled to the field of a superc<strong>on</strong>ductingcavity with a very high Q-factor while crossing the cavity mode <strong>on</strong>e by <strong>on</strong>e [3, 4]. Str<strong>on</strong>gcoupling in the optical domain has been reached using ultra-cold ground-state atoms andcavities with small mode volume [5, 6].About 10 years ago a particular interest in C<strong>QED</strong> systems arouse due to their possibleapplicati<strong>on</strong>s in quantum in<strong>for</strong>mati<strong>on</strong> processing [7], where quantum c<strong>on</strong>cepts can lead todramatic speed up in solving certain classes of computati<strong>on</strong>al problems, such as primefactoring [8] and database search [9]. Using individual atoms as carriers of quantum in<strong>for</strong>mati<strong>on</strong>(qubits), two-qubit logic operati<strong>on</strong>s can be per<strong>for</strong>med in a cavity by the coherentexchange of cavity phot<strong>on</strong> between the atoms [10]. Since a cavity emits in a well-definedmode and a radiated phot<strong>on</strong> carries in<strong>for</strong>mati<strong>on</strong> <strong>on</strong> an atom, C<strong>QED</strong> systems could alsobe efficiently used <strong>for</strong> building quantum logic network [11]. Here, a phot<strong>on</strong> emitted by thecavity transfers quantum in<strong>for</strong>mati<strong>on</strong> between qubits located at spatially separated nodesof a network.Besides str<strong>on</strong>g coupling, <strong>on</strong>e of the most important requirement <strong>for</strong> using C<strong>QED</strong> systems<strong>for</strong> quantum in<strong>for</strong>mati<strong>on</strong> processing is permanent localizati<strong>on</strong> of single atoms in thecavity. The required degree of c<strong>on</strong>finement can be provided with i<strong>on</strong>s in i<strong>on</strong> traps [12].<str<strong>on</strong>g>Single</str<strong>on</strong>g> neutral atoms can be trapped inside the cavity by using intra-cavity dipole traps[13], external dipole traps [14], or near-res<strong>on</strong>ant cavity fields [15, 16]. Besides two-qubit1


2 Introducti<strong>on</strong>quantum gates, the realizati<strong>on</strong> of a quantum computer also requires the ability to initializeand to measure the states of the qubits [17]. We have already dem<strong>on</strong>strated thata string of neutral atoms stored in a standing-wave dipole trap can serve as a quantumregister, where atomic qubits can be individually addressed and coherently manipulatedmuch faster than the relevant decoherence times [18]. Besides, we are able to transportthe atoms over several millimeters while partly retaining (single-atom) quantum coherence[19].Till recently the two basic approaches to deliver single ground-state atoms into a cavityhave been either to let them freely fall from a magneto-optical trap (MOT) [20], or to ejectthem from below by means of an atomic fountain [21]. In order to increase the efficiency ofthe atom delivery, <strong>on</strong>e can use a dipole trap beam <strong>for</strong> guiding atoms between a MOT anda cavity while c<strong>on</strong>fining them transversally [22, 14]. Recently, the submicr<strong>on</strong> positi<strong>on</strong>ingof single atoms inside a microcavity has been dem<strong>on</strong>strated using a standing-wave trap[14]. However, this experiment does provide neither the possibility to prepare quantumstates of atoms be<strong>for</strong>e placing them into the cavity nor to analyze their final states.In this thesis I present an experiment <strong>on</strong> the deterministic coupling of a single neutralatom to a mode of a high-finesse optical res<strong>on</strong>ator. A desired number of ultra-coldcesium atoms (from 1 to 20) is prepared in a MOT and then transferred without lossinto a standing-wave dipole trap. This trap provides a c<strong>on</strong>servative potential required<strong>for</strong> experiments <strong>on</strong> quantum in<strong>for</strong>mati<strong>on</strong> as well as a tight c<strong>on</strong>finement of atoms. Ourapproach <strong>for</strong> c<strong>on</strong>trolling the positi<strong>on</strong> of individual atoms is based <strong>on</strong> the “optical c<strong>on</strong>veyorbelt” technique [23, 24]. Combining it with fluorescence imaging of single trapped atoms,I have measured and c<strong>on</strong>trolled their positi<strong>on</strong> with submicrometer precisi<strong>on</strong> [25, 26]. Theoptical cavity I have assembled is composed of two mirrors of high reflectivity placed closeto each other and is stabilized to the atomic transiti<strong>on</strong>. A finesse of about a milli<strong>on</strong> anda small mode volume enable us to operate the atom-cavity system in the str<strong>on</strong>g-couplingregime. I was able to couple single atoms to the cavity mode <strong>on</strong> demand. Together withthe possibility to prepare their initial states as well as to analyze their final states, thiswork represents an important step towards deterministic quantum computing with neutralatoms [27].


Chapter 1Trapping and observing singleatoms1.1 Introducti<strong>on</strong>To per<strong>for</strong>m any experiment <strong>on</strong> single atoms we need a set of tools providing a necessarydegree of c<strong>on</strong>trol over them. Moreover, the atoms should be well isolated from envir<strong>on</strong>ment,which, permanently probing the atom system, would lead to its decoherence andthus would complicate the study of any coherent process. A widely used approach overcomingthis problem is to experiment with atoms in vacuum or in dilute gases instead of inand <strong>on</strong> solids or liquids. However, high velocities at room or even cryogenic temperaturesprevent l<strong>on</strong>g observati<strong>on</strong> times of flying atoms.The great breakthrough in atom studies has been made with the inventi<strong>on</strong> of lasercooling [28, 29, 30] and magneto-optical trapping of atoms [31]. This allowed not <strong>on</strong>lycooling of atomic species to sub-millikelvin temperatures but also storing them <strong>for</strong> manysec<strong>on</strong>ds and even minutes. The creati<strong>on</strong> of Bose-Einstein c<strong>on</strong>densate became the triumphof these developments [32, 33]. However, although a magneto-optical trap (MOT) allows usto efficiently capture, cool, and store even single atoms [34, 35, 36], it does not completelysolve the problem of the isolati<strong>on</strong> of atoms from the envir<strong>on</strong>ment and does not allow usto store atoms in a specific quantum state since the operati<strong>on</strong> of the MOT is based <strong>on</strong>nearly-res<strong>on</strong>ant scattering of the MOT lasers.In c<strong>on</strong>trast to MOTs, c<strong>on</strong>servative dipole traps rely <strong>on</strong> off-res<strong>on</strong>ant interacti<strong>on</strong> ofpolarizable particles with the laser field [37, 38] resulting in low light scattering. Alsoknown as optical tweezers [39], they are successfully used in different research areas rangingfrom biology to nanofabricati<strong>on</strong> <strong>for</strong> the precise manipulati<strong>on</strong> of various dielectric particles:from microspheres and biological cells to DNA and single atoms. Another applicati<strong>on</strong>of dipole traps is to <strong>for</strong>m optical lattices by the interference of laser beams. They areextensively used <strong>for</strong> studying quantum gases [40], in quantum in<strong>for</strong>mati<strong>on</strong> processing [41],in metrology [42], etc. We have experimentally shown that a standing-wave dipole trappreserves the atomic coherence [19] and can be used as a holder of a quantum register <strong>for</strong>storing quantum in<strong>for</strong>mati<strong>on</strong> [18].3


4 Chapter 1: Tools <strong>for</strong> single atom c<strong>on</strong>trolSpatial in<strong>for</strong>mati<strong>on</strong> <strong>on</strong> trapped atoms is essential <strong>for</strong> realizing complete c<strong>on</strong>trol overthem. The observati<strong>on</strong> of atoms in optical and magnetic traps is typically per<strong>for</strong>med viafluorescence imaging, in c<strong>on</strong>trast to, e.g., atomic <strong>for</strong>ce and scanning tunnelling microscopyused <strong>for</strong> imaging atoms <strong>on</strong> surfaces of solids [43]. The first image of an individual atomicparticle in a trap was obtained by recording the fluorescence from a single barium i<strong>on</strong> [44].Nowadays we are able to n<strong>on</strong>-destructively observe individual neutral atoms in the dipoletrap <strong>for</strong> more than a minute [45].1.2 A single-atom magneto-optical trapThe c<strong>on</strong>siderable progress in atomic physics observed over the last two decades would havehardly been possible without the inventi<strong>on</strong> of the laser cooling techniques allowing us toslow down, cool, and even Bose c<strong>on</strong>dense atomic gases. Proposed about 30 years ago byT. Hänsch and A. Shawlow [28], the three-dimensi<strong>on</strong>al laser cooling of atoms has been firstexperimentally realized by S. Chu [30] in 1985. By adding an inhomogeneous magneticfield, resulting in a positi<strong>on</strong>-dependent radiati<strong>on</strong> pressure, the first magneto-optical trappingof atoms has been dem<strong>on</strong>strated two years later [31]. Since then, magneto-opticaltraps became a widely used tool <strong>for</strong> cooling atoms down to sub-millikelvin temperaturesand keeping them trapped <strong>for</strong> l<strong>on</strong>g times. We use a specially designed MOT as a sourceof single cold cesium atoms <strong>for</strong> our experiments. The high-gradient magnetic field assistsus in trapping single or few atoms, whereas the MOT fluorescence allows us to infer theirexact number in real time.1.2.1 Principle of operati<strong>on</strong>The MOT’s speciality is the simultaneous cooling and trapping of neutral atoms [46, 47].Its operati<strong>on</strong> relies <strong>on</strong> a velocity-dependent cooling <strong>for</strong>ce and a positi<strong>on</strong>-dependent restoring<strong>for</strong>ce providing spatial c<strong>on</strong>finement of the atoms. The cooling is realized by threeorthog<strong>on</strong>al pairs of counter-propagating laser beams which are slightly red-detuned fromthe atomic transiti<strong>on</strong> frequency. An atom moving at the intersecti<strong>on</strong> of the beams preferentiallyabsorbs phot<strong>on</strong>s from the beam opposite to its directi<strong>on</strong> of moti<strong>on</strong>, frequencyof which is Doppler-shifted closer to the atomic res<strong>on</strong>ance. Since the subsequent sp<strong>on</strong>taneousemissi<strong>on</strong> is in random directi<strong>on</strong>, it does not change the average momentum of theatom and the total momentum transfer from the laser field to the atom is opposite tothe directi<strong>on</strong> of the atomic moti<strong>on</strong>. There<strong>for</strong>e this c<strong>on</strong>figurati<strong>on</strong> of the laser beams called“optical molasses” results in an effective fricti<strong>on</strong> <strong>for</strong>ce cooling the atom.Even at rest, the atom is c<strong>on</strong>tinuously excited by the near-res<strong>on</strong>ant laser field followedby phot<strong>on</strong> emissi<strong>on</strong> in a random directi<strong>on</strong> causing atom heating. The equilibrium temperaturebetween the molasses cooling and the heating by the sp<strong>on</strong>taneous emissi<strong>on</strong> is calledthe Doppler temperature and is given byT D = Γ2k B, (1.1)


1.2 A single-atom magneto-optical trap 5(a)m = -1J'0+1s +dw - d0E+10-1(b)s -s + s +s - s -IBzs -s +yxm =J00zIFigure 1.1: Principle of operati<strong>on</strong> of a magneto-optical trap. (a) Simplified level schemeof an atom interacting with two counter-propagating red-detuned laser beams with oppositecircular polarizati<strong>on</strong>. The linear magnetic field gradient lifts the degeneracy of theexcited state. If the atom is displaced from the trap center, it becomes res<strong>on</strong>ant withthe laser beam pushing it back to the zero of the field. (b) A three-dimensi<strong>on</strong>al MOTuses a pair of coils in anti-Helmholtz c<strong>on</strong>figurati<strong>on</strong> producing a quadrupole field and athree-dimensi<strong>on</strong>al optical molasses.where Γ is the sp<strong>on</strong>taneous emissi<strong>on</strong> rate from the excited state and k B is the Boltzmannc<strong>on</strong>stant. For a cesium atom, Γ = 2π × 5.22 MHz and T D = 125 µK.Trapping of atoms at a specific positi<strong>on</strong> requires a positi<strong>on</strong>-dependent <strong>for</strong>ce. For thispurpose a quadrupole magnetic field is added which vanishes at the center of the opticalmolasses and increases linearly in all directi<strong>on</strong>s. The magnetic field lifts the degeneracy ofthe upper level with respect to the Zeeman sublevels resulting in a level splitting proporti<strong>on</strong>alto the distance of the atom to the magnetic zero point, i.e. the trap center. Thisis illustrated in Fig. 1.1(a) in <strong>on</strong>e dimensi<strong>on</strong> <strong>for</strong> the case of a J = 0 ↔ J ′ = 1 transiti<strong>on</strong>.If the red-detuned counter-propagating laser beams have opposite circular polarizati<strong>on</strong>s,a stati<strong>on</strong>ary atom displaced from the zero point is shifted closer into res<strong>on</strong>ance with thatlaser beam which pushes it back to the center. Thus, the atom experiences a restoring<strong>for</strong>ce to the trap center. This principle can be easily generalized to the three-dimensi<strong>on</strong>alcase shown in Fig. 1.1(b). As a result, the radiati<strong>on</strong> <strong>for</strong>ce in the MOT simultaneouslypushes the atom to the center of the MOT and cools it there. Moreover, the multi-levelstructure of the atom can allow sub-Doppler cooling mechanisms resulting in temperatureslower than T D [48, 49], which will be not discussed here.1.2.2 Experimental setupThe first single-atom MOT in our group was c<strong>on</strong>structed in 1996 [36]. The details of themodified MOT setup permitting more flexibility in the observati<strong>on</strong> and manipulati<strong>on</strong> ofthe atoms have been described extensively in previous theses of the group [50, 51]. Here, I<strong>on</strong>ly present the most important comp<strong>on</strong>ents which are relevant <strong>for</strong> this thesis. The basic


6 Chapter 1: Tools <strong>for</strong> single atom c<strong>on</strong>trolxzyvalvelens (f = 300mm)dipole trap laseroptical tableCs ampulesublimati<strong>on</strong>pumpglass cellMOTlaserMOTmagnetic coilsMOTi<strong>on</strong> pumpFigure 1.2: Side view of an optical table with the vacuum system and the optical traps.experimental setup is schematically shown in Figs. 1.2 and 1.4.Vacuum systemOur vacuum system fulfills two main requirements <strong>for</strong> experimenting with trapped atoms:It provides an ultra-high vacuum guaranteeing l<strong>on</strong>g storage times and permits optimaloptical access from outside <strong>for</strong> flexible observati<strong>on</strong> and manipulati<strong>on</strong> of atoms. All ourexperiments are per<strong>for</strong>med inside a compact 3 × 3 × 12.5 cm 3 glass cell with 5 mm thickwalls, see Fig. 1.2. It is c<strong>on</strong>nected to the vacuum steel cube to which the pumps and thecesium reservoir are attached. The c<strong>on</strong>stantly operating i<strong>on</strong> pump provides an ultra-highvacuum in the glass cell with a pressure of less than 10 −10 mbar. An additi<strong>on</strong>al titaniumsublimati<strong>on</strong> pump is mainly used <strong>for</strong> reaching the initial low pressures inside the newlyclosed vacuum chamber, e.g., after installati<strong>on</strong> of a cavity inside the cell, see Sec. 3.3.1.Laser systemIn c<strong>on</strong>trast to the simplified level structure used in Fig. 1.1(a), the D 2 transiti<strong>on</strong> in a cesiumatom used <strong>for</strong> cooling and trapping has a rich spectrum, schematically shown in Fig. 1.3.As the MOT cooling transiti<strong>on</strong> we use the cycling transiti<strong>on</strong> F = 4 → F ′ = 5. Since theatom can be off-res<strong>on</strong>antly excited to the F ′ = 4 level, from where it may sp<strong>on</strong>taneouslydecay into the F = 3 ground state, we apply an additi<strong>on</strong>al repumping laser res<strong>on</strong>ant withthe transiti<strong>on</strong> F = 3 → F ′ = 4 to pump the atom back into the cooling cycle.Both cooling and repumping lasers are diode lasers set up in Littrow c<strong>on</strong>figurati<strong>on</strong> [52].


1.2 A single-atom magneto-optical trap 76 2 P3/2251.0 MHz201.2 MHz151.2 MHzF' = 5F' = 4F' = 3F' = 22-D line852.3 nmMOTcoolingMOTrepumpingF = 46 2 S1/29.193 GHzF = 3Figure 1.3: Hyperfine structure of the cesium D 2 transiti<strong>on</strong>. Shown are the transiti<strong>on</strong>sused <strong>for</strong> cooling and repumping in the MOT.Their frequencies are actively stabilized <strong>on</strong>to atomic res<strong>on</strong>ance frequencies by polarizati<strong>on</strong>spectroscopy [51]. The probe laser is locked to the crossover transiti<strong>on</strong> F = 4 → F ′ = 3/5,which is red-detuned by -225 MHz from the cycling transiti<strong>on</strong>. An acousto-optic modulator(AOM) in double pass c<strong>on</strong>figurati<strong>on</strong> shifts the laser frequency by 2×110 MHz towards theatomic res<strong>on</strong>ance. As a result, the cooling laser is red-detuned from the cooling transiti<strong>on</strong>by approximately Γ, which is required <strong>for</strong> optimal Doppler cooling. In additi<strong>on</strong>, this AOMis also used to c<strong>on</strong>trol the laser power and frequency <strong>for</strong> illuminating atoms in the dipoletrap, see Sec. 1.5. The repumping laser is directly stabilized to the F = 3 → F ′ = 4transiti<strong>on</strong>.Both lasers with their spectroscopies are set up <strong>on</strong> a separate optical table, and weuse optical fibers to transfer the laser light to the main table, shown in Fig. 1.2. Here, thecooling laser is split into three beams, which are shined from three different directi<strong>on</strong> intothe center of the quadrupole field and then retro-reflected. Thus, the six beams intersectin <strong>on</strong>e point in the glass cell providing a 3-D optical molasses. The beam diameter isabout 2 mm with a typical power of 100 − 200 µW per beam. The MOT repumping laseris linearly polarized and is shined into the MOT al<strong>on</strong>g the glass cell.Magnetic coilsThe high field gradient of the MOT is produced by a pair of water cooled magnetic coils,set in anti-Helmholtz c<strong>on</strong>figurati<strong>on</strong> al<strong>on</strong>g the z axis and placed above and below theglass cell, see Fig. 1.2. In all experiments described in this thesis we use a gradient of∂B/∂z = 300 G/cm provided by a current of 15 A, which can be switched <strong>on</strong> and offwithin about 30 ms.Three orthog<strong>on</strong>al pairs of coils placed around the glass cell, see Fig. 1.4, compensate


8 Chapter 1: Tools <strong>for</strong> single atom c<strong>on</strong>troldipole trap laserglasscellvacuum chamberinterferencefilterICCDMOTcoils <strong>for</strong>guiding fieldsobjectivespatialfilterAPDMOT coilsMOT laser beamsxdipole trap laseryzFigure 1.4: Schematic experimental setup of MOT, dipole trap, and imaging system(top view). The MOT c<strong>on</strong>sists of three pairs of counter-propagating laser beams and thecoil pair (al<strong>on</strong>g z) providing the quadrupole field. Both dipole trap lasers are focussedinto the MOT in the center of the glass cell. The fluorescence light from the MOT iscollected and collimated by imaging optics. One part is spatially and spectrally filteredand focussed <strong>on</strong>to an avalanche photodiode (APD), the other part is sent to an intensifiedCCD camera (ICCD) after spectral filtering. Three pairs of orthog<strong>on</strong>al coils are used <strong>for</strong>applying guiding magnetic fields. Beams and coils al<strong>on</strong>g z directi<strong>on</strong> are not shown.DC-magnetic fields in three dimensi<strong>on</strong>s. In additi<strong>on</strong>, they are used <strong>for</strong> applying guidingmagnetic fields of up to several Gauss used in Sec. 4.5.5.1.2.3 Fluorescence detecti<strong>on</strong> of single atomsTo observe the trapped atoms in the MOT we use two detectors, see Fig. 1.4. An avalanchephotodiode (APD) operated in single-phot<strong>on</strong> counting mode allows us to determine theirexact number and an intensified CCD camera (ICCD) provides spatial in<strong>for</strong>mati<strong>on</strong>. Inorder to efficiently collect and collimate extremely low levels of fluorescence light fromsingle trapped atoms we use a home-built diffracti<strong>on</strong>-limited objective with a numericalaperture of NA = 0.29 [53]. The collimated light is then equally divided into two parts bya beam-splitter: the reflected part is imaged <strong>on</strong>to the ICCD with a magnificati<strong>on</strong> of about14.0 (see Sec. 1.5) and the transmitted light is focused <strong>on</strong>to the APD (EG&G, modelSPCM-200). To reduce the stray light background, the collected light is spectrally filtered


1.3 A standing-wave optical dipole trap 9305 atomsPhot<strong>on</strong>s [10 3 counts / 100 ms]25201510500 20 40 60 80 100Time [s]4 atoms3 atoms2 atoms1 atom0 atomsFigure 1.5: APD signal of the MOT fluorescence. Each trapped atom c<strong>on</strong>tributes thesame amount of fluorescence to the signal, allowing us to determine their exact number.When an atom enters or leaves the trap, the count rate instantaneously increases ordecreases accordingly.by an interference filter. In additi<strong>on</strong>, the light sent to the APD is filtered spatially with apinhole. By taking into account the numerical aperture of the objective, the transmissi<strong>on</strong>of the interference filter, and the quantum efficiency of the APD, the overall detecti<strong>on</strong>efficiency with the APD is η = 8 × 10 −3 . Typically, we obtain a count rate of 3 × 10 4 s −1per atom.A typical APD signal of the MOT fluorescence is shown in Fig. 1.5. The step-like<strong>for</strong>m of the count rate indicates the instantaneous capture and loss of the atoms. Sinceeach atom c<strong>on</strong>tributes the same amount of fluorescence to the APD signal, we can use theMOT fluorescence to infer the exact number of trapped atoms within typically 20 ms.1.3 A standing-wave optical dipole trapOptical dipole traps rely <strong>on</strong> the electric dipole interacti<strong>on</strong> of polarizable particles (e.g.,neutral atoms) with far-detuned light. The corresp<strong>on</strong>ding dipole <strong>for</strong>ce arises from theinteracti<strong>on</strong> of the induced dipole moment with the gradient of the light field [54]. Since theoptical excitati<strong>on</strong> in these traps can be kept very low, they provide a nearly c<strong>on</strong>servativetrapping potential favourable <strong>for</strong> experiments <strong>on</strong> quantum state manipulati<strong>on</strong>. We use astanding-wave c<strong>on</strong>figurati<strong>on</strong> of the dipole trap providing tight axial c<strong>on</strong>finement of atomsand allowing their positi<strong>on</strong> c<strong>on</strong>trol.


10 Chapter 1: Tools <strong>for</strong> single atom c<strong>on</strong>trol1.3.1 Dipole potentialThere are several different approaches to treat the dipole <strong>for</strong>ce [38]. Here I recall <strong>on</strong>lythe classical model, which provides a simple descripti<strong>on</strong> of the dipole interacti<strong>on</strong> andits basic properties, namely the dipole potential and the scattering rate, relevant to ourexperiments. The predicti<strong>on</strong>s of this model are good approximati<strong>on</strong>s to the exact soluti<strong>on</strong>given by the quantum-mechanical treatment.In the following we c<strong>on</strong>sider an atom as a charged harm<strong>on</strong>ic oscillator driven by aclassical electromagnetic field E(t) = E 0 exp(−iωt)+c.c. Given α the atomic polarizability,the field induces an electric dipole moment d(t) = αE(t) which evolves according to theequati<strong>on</strong> of moti<strong>on</strong>¨d(t) + Γḋ(t) + ω2 0d(t) = e2m eE(t). (1.2)Here, m e and e are the mass and the charge of the electr<strong>on</strong>, ω 0 is the atomic res<strong>on</strong>ancefrequency, and Γ is the damping rate due to the radiative energy loss given byΓ = e2 ω 26πɛ 0 m e c 3 . (1.3)Integrati<strong>on</strong> of equati<strong>on</strong> (1.2) yields the complex polarizability depending <strong>on</strong> the drivingfrequencyα(ω) = e2 1m e ω0 2 − ω2 − iωΓ . (1.4)The dipole potential is the interacti<strong>on</strong> energy of the induced dipole moment and theelectric field given byU dip (r) = − 1 1〈d · E〉 = − Re(α)I(r) (1.5)2 2ɛ 0 cwhere the angular brackets denote the time average over <strong>on</strong>e oscillati<strong>on</strong> period and the fieldintensity is I(r) = 2cɛ 0 |E 0 | 2 . The dipole trap depth is thus proporti<strong>on</strong>al to the intensityI(r) and to the real part of the polarizability, which describes the in-phase comp<strong>on</strong>ent ofthe atomic dipole moment. The dipole <strong>for</strong>ce results from the gradient of the interacti<strong>on</strong>potential F dip (r) = −∇U dip (r).The scattering rate is determined by the amount of power absorbed from the drivingfield, P abs (r), divided by the energy per phot<strong>on</strong>, ω:R sc (r) = P abs(r)ω=〈ḋ · E〉ω= 1 Im(α)I(r). (1.6)ɛ 0 cIn the regime of large detuning |ω − ω 0 | ≫ Γ equati<strong>on</strong>s (1.5) and (1.6) can be approximatedbyU dip (r) = Γ Γ I(r)8 ∆ ′ , (1.7)I 0R sc (r) = Γ ( ) Γ 2I(r)8 ∆ ′ (1.8)I 0= Γ∆ ′ U dip(r) . (1.9)


1.3 A standing-wave optical dipole trap 11where1 1∆ ′ = + 1 . (1.10)ω − ω 0 ω + ω 0and I 0 is the saturati<strong>on</strong> intensity given byI 0 = πhΓc3λ 3 . (1.11)In the case of a cesium atom, I 0 = 11 W/m 2 . If the light detuning from the atomicres<strong>on</strong>ance, ∆ = ω − ω 0 , is much smaller than the atomic frequency, we can apply therotating-wave approximati<strong>on</strong> and neglect the term with the sum of optical frequenciesω + ω 0 in equati<strong>on</strong> (1.10) leading to ∆ ′ ≈ ∆.If the dipole trap laser is red-detuned from the atomic res<strong>on</strong>ance, i.e. ∆ < 0, the dipolepotential given by equati<strong>on</strong> (1.7) is negative, so that the atom is attracted to the regi<strong>on</strong>sof high light intensities. For the case of a blue-detuned laser (∆ > 0), the interacti<strong>on</strong>energy is positive and thus the atoms are repelled from the light field. The relati<strong>on</strong> (1.9)shows that the scattering rate can be minimized by using large laser detunings whilecompensating the decreasing trap depth by higher light intensities.1.3.2 Multi-level cesium atomThe classical model does not take into account the atomic multi-level structure. Still, wecan independently apply the classical treatment to each individual transiti<strong>on</strong> and add upthe outcomes weighted with the corresp<strong>on</strong>ding transiti<strong>on</strong>’s oscillator strength [51]. As aresult, c<strong>on</strong>sidering both c<strong>on</strong>tributi<strong>on</strong>s from the D 1 and D 2 transiti<strong>on</strong>s of a cesium atom(see Fig. A.1), we can introduce the “effective” laser detuning1∆ eff= 1 3( 1∆ D1+ 2∆ D2)(1.12)and still use expressi<strong>on</strong>s (1.7) and (1.8) <strong>for</strong> calculating the trap depth and the scatteringrate after replacing ∆ ′ by ∆ eff .A full quantum mechanical treatment of a cesium atom in a far-detuned dipole trap,which takes into account atomic fine, hyperfine, and Zeeman states (see Appendix A <strong>for</strong>the relevant states of cesium), can be found in Ref. [55]. It provides ac Stark shifts (alsocalled “light shifts”) <strong>for</strong> each atomic level depending <strong>on</strong> the polarizati<strong>on</strong> of the laser light.1.3.3 Standing-wave trapA widely used geometric c<strong>on</strong>figurati<strong>on</strong> of dipole traps is based <strong>on</strong> a single str<strong>on</strong>gly focusedlaser beam. To achieve a tighter trap c<strong>on</strong>finement, we use a standing-wave dipole trap<strong>for</strong>med by the interference of two counter-propagating laser beams. The beams have equalintensities, equal optical frequencies, and parallel linear polarizati<strong>on</strong>s. Neglecting theGuoy phase and the curvature of the wavefr<strong>on</strong>ts of the Gaussian beams, the resultingdipole potential isw 2 [0U(x, y, z) = −U 0w 2 (x) exp − 2(y2 + z 2 ])w 2 (x)cos 2 ( 2πλ x ), (1.13)


12 Chapter 1: Tools <strong>for</strong> single atom c<strong>on</strong>trolFigure 1.6: Three-dimensi<strong>on</strong>al view of the standing-wave dipole potential <strong>for</strong> the beamwaist of w 0 = 19 µm. In x-directi<strong>on</strong>, the wavelength has been stretched by a factor of400 to show the individual potential wells where atoms can be trapped.with the beam waist w 0 , the beam radius w(x) = w 0 (1 + x 2 /zR 2 )1/2 , and the Rayleighrange z R = πw0 2 /λ. Here we c<strong>on</strong>sider the standing wave to be oriented al<strong>on</strong>g x directi<strong>on</strong>.The maximum trap depth readsU 0 = Γ 8I maxI 0Γ|∆ eff | , (1.14)where the peak intensity isI max = 4Pπw02 (1.15)with P the total optical power of the beams. Figure 1.6 shows the trapping potential <strong>for</strong>w 0 = 19 µm.Being trapped in the dipole trap, an atom oscillates inside its potential wells in bothaxial and radial directi<strong>on</strong>s. If the atomic temperature is much smaller than the trap depth,the trapping potential can be approximated by a harm<strong>on</strong>ic <strong>on</strong>e resulting in the respectiveoscillati<strong>on</strong> frequencies√2U0Ω ax = 2πmλ 2 , (1.16)√4U 0Ω rad = , (1.17)where m denotes the atomic mass of cesium.mw 2 0


1.3 A standing-wave optical dipole trap 131.3.4 Experimental setupNd:YAG laserIn all previous experiments with the standing-wave dipole trap per<strong>for</strong>med in our group[50, 51, 55] as well as in Chapter 2 of this thesis we have used a Nd:YAG laser as a dipoletrap laser. It is an industrial, high power cw laser (Quantr<strong>on</strong>ix, model 116EF-OCW-10)with a wavelength of λ = 1064 nm and a maximum output power of 10 W. The outputlight has linear polarizati<strong>on</strong> and a Gaussian TEM 00 transverse mode profile. The laseroperates <strong>on</strong> 4–5 l<strong>on</strong>gitudinal modes with a spacing of 196 MHz. This corresp<strong>on</strong>ds toa coherence length of about 30 cm ensuring a high-c<strong>on</strong>trast interference pattern of thestanding-wave dipole trap inside the glass cell.The output beam is split into two parts, which are then focused to the MOT centerfrom opposite sites, as shown in Fig. 1.4. Typically we get 2 W of total optical powerinside the glass cell. The beam waist of 19 µm and the trap depth of U 0 /k B = 0.8 mK areinferred by measuring the oscillati<strong>on</strong> frequencies of the trap [55]: Ω ax = 2π ×(265±8) kHz<strong>for</strong> P YAG = 1.56 W and Ω rad = 2π ×(3.6±0.2) kHz <strong>for</strong> P YAG = 1.8 W. The measured U 0 is<strong>on</strong>ly about 40 % of that expected from equati<strong>on</strong> (1.7). This discrepancy can be explainedby aberrati<strong>on</strong>s and diffracti<strong>on</strong> in the dipole trap optics increasing the beam waist anddecreasing the beam power. The scattering rate at the trap minimum of R sc = 9 s −1 isdirectly obtained from the trap depth according to equati<strong>on</strong> (1.9).Yb:YAG laserThe old Nd:YAG laser has served us well during many years. However, our new experimentsset more stringent requirements <strong>on</strong> the dipole trap laser. For instance, we haveobserved that the multi-mode structure of the Nd:YAG laser complicates the illuminati<strong>on</strong>of atoms (see Sec. 1.5). Besides, if using beams with a larger Rayleigh range necessary <strong>for</strong>transporting atoms over 5 mm distance into the cavity (see Sec. 3.4), the reduced beamintensity should be compensated by the increased power of the dipole trap laser in orderto keep the trap depth reas<strong>on</strong>ably large.To get more optical power, we have replaced the Nd:YAG laser by an Yb:YAG disc laser(ELS, model VersaDisk-1030-10-SF) with a wavelength of λ = 1030 nm and a maximumoutput power of about 25 W, which was then used in all experiments in Chapters 3 and 4.However, be<strong>for</strong>e using the laser, we have slightly modified it to provide c<strong>on</strong>tinuous singlemodeoperati<strong>on</strong> and a higher temperature stability. The modificati<strong>on</strong>s have included watercooling of the whole laser head, the utilizati<strong>on</strong> of a scanning Fabry-Perot interferometer <strong>for</strong>m<strong>on</strong>itoring the mode structure of the laser in real time, and the installati<strong>on</strong> of a motorizedholder <strong>for</strong> the outcoupling mirror allowing us to adjust the laser res<strong>on</strong>ator without openingthe laser cover each time when mode hopping is detected. As a result, the laser emits asingle l<strong>on</strong>gitudinal mode with a stable output power of about 20 W, resulting in up to 5 Wlaser power in each DT beam be<strong>for</strong>e the glass cell. The modified dipole trap geometryadapted <strong>for</strong> the cavity experiments is described in Sec. 3.4 in c<strong>on</strong>necti<strong>on</strong> with the cavityproperties.


14 Chapter 1: Tools <strong>for</strong> single atom c<strong>on</strong>trol(a)events806040200 atoms (stray light)1 atom2345678910(b)events806040203 lost2 lost1 lostn = 51 captured(a)00 5 10 15 20 25number of fluorescence counts [103]00 5 10 15 20 25number of fluorescence counts [103]Figure 1.7: Histograms of fluorescence counts from the MOT integrated over 60 ms(bin size: 120 counts). Each peak corresp<strong>on</strong>ds to a specific number of trapped atoms. Inpanel (a), the fluorescence was measured directly after loading the MOT with five atoms<strong>on</strong> average. The actual number of atoms in the MOT follows Poiss<strong>on</strong>ian statistics to agood approximati<strong>on</strong>, leading to the observed distributi<strong>on</strong> of peak areas. For events wherethe software discriminator detected five atoms in the MOT, the atoms were transferredto the DT. These atoms were counted by transferring them back to the MOT where thefluorescence was measured a sec<strong>on</strong>d time, leading to the final histogram in panel (b).1.4 Loading of the DT with atomsTransferring atoms between trapsThe dipole trap is loaded with cold atoms from the MOT. For this purpose, both traps aresimultaneously operated <strong>for</strong> several tens of millisec<strong>on</strong>ds be<strong>for</strong>e we switch off the MOT. Totransfer the atoms back into the MOT at the end of an experimental sequence, e.g., <strong>for</strong>counting them, we use the reverse procedure. Since the presence of the dipole trap shiftsthe atomic res<strong>on</strong>ance, we change the parameters of the MOT cooling beams (power andfrequency) by means of the AOM in order to better cool the atoms into the DT duringtheir transfer. For small numbers of atoms (up to about 5) the transfer efficiency is betterthan 99 %. For larger atom numbers cold collisi<strong>on</strong>s between atoms occasi<strong>on</strong>ally loadedinto the same potential well lead to their loss, thus reducing the transfer efficiency [56].Number-triggered loadingThe loading of the MOT is a statistical Poiss<strong>on</strong>ian process. There<strong>for</strong>e, we cannot predictthe exact number of atoms which will be trapped in the MOT and then transferred intothe DT. However, in some applicati<strong>on</strong>s it is necessary to have a predetermined numberof atoms in the trap. For this purpose, we have developed a computer-c<strong>on</strong>trolled loadingsequence that m<strong>on</strong>itors the atom number via the MOT fluorescence level [56]. If theatom number is the desired <strong>on</strong>e, the atoms are transferred into the DT and the mainexperimental sequence is started. If not, they are ejected from the MOT and the MOT is


1.5 Imaging single atoms 15reloaded. The whole “number-triggered loading” procedure runs in a fully automated wayallowing us to per<strong>for</strong>m the same experimental sequence many times with a preset initialnumber of atoms.In order to correctly determine the number of trapped atoms, we integrate the MOTfluorescence (Fig. 1.5) over 60 ms. The corresp<strong>on</strong>ding histogram of fluorescence countsis shown in Fig. 1.7(a). Here, the MOT parameters are optimized to load <strong>on</strong> average5 atoms. Each peak corresp<strong>on</strong>ds to a specific number of trapped atoms. As an example,Fig. 1.7(b) dem<strong>on</strong>strates the preparati<strong>on</strong> of 5 atoms in the DT. To verify the atom numberactually prepared in the DT, we transfer the atoms back into the MOT and count themagain. The peak corresp<strong>on</strong>ding to five atoms comprises 88 % of all events. We define thispercentage as the preparati<strong>on</strong> efficiency <strong>for</strong> five atoms. In the other events, atoms havebeen lost or an additi<strong>on</strong>al atom has been captured.To avoid the post-selecti<strong>on</strong> of experimental shots started with <strong>on</strong>ly <strong>on</strong>e atom and tospeed up data acquisiti<strong>on</strong>, we per<strong>for</strong>m all experiments <strong>on</strong> transporting single atoms intothe cavity, described in Sec. 4.5, with the number-triggered loading of the DT with <strong>on</strong>eatom (the preparati<strong>on</strong> efficiency of 97(±1) %). More details <strong>on</strong> the presented loadingtechnique are described extensively in Ref. [56].A similar feedback technique has been proposed in Ref. [57] <strong>for</strong> loading the MOT withsingle chromium atoms. Here, the MOT loading is actively suppressed or enhanced depending<strong>on</strong> the actual number of trapped atoms in order to have <strong>on</strong>ly <strong>on</strong>e atom trapped.Another approach <strong>for</strong> preparing <strong>on</strong>e atom in a tightly focused dipole trap is based <strong>on</strong>a “collisi<strong>on</strong>al blockade” mechanism, which locks the average number of loaded atomto 0.5 [58, 59].Storage timeThe measured lifetime of atoms in the dipole trap is about 25 s limited by backgroundgas collisi<strong>on</strong>s. However, <strong>for</strong> the realizati<strong>on</strong> of the moving standing wave (see Sec. 2.4) theboth dipole trap beams are sent through acousto-optic modulators (AOMs) to mutuallydetune their frequency. The phase noise of the dual-frequency generator, which drivesboth AOMs, results in heating and subsequent loss of trapped atoms, limiting the storagetime to about 6 s. For an overview of various heating mechanisms in the DT see Ref. [51].1.5 Imaging single atomsSpatial in<strong>for</strong>mati<strong>on</strong> <strong>on</strong> the trapped atoms, such as their positi<strong>on</strong> and separati<strong>on</strong>s, is essential<strong>for</strong> many experiments with single or few atoms. It was a prerequisite <strong>for</strong> the individualaddressing of the atoms and the realizati<strong>on</strong> of a neutral atom quantum register [18, 60].The first visual evidence of the transport of single atoms with our “optical c<strong>on</strong>veyor belt”(Sec. 2.4) was per<strong>for</strong>med by c<strong>on</strong>tinued imaging of the atoms while shifting the standingwavepattern [45]. In this thesis I use the spatial in<strong>for</strong>mati<strong>on</strong> <strong>for</strong> sub-micrometer positi<strong>on</strong>c<strong>on</strong>trol of single atoms, see Chapter 2 and Ref. [25], allowing deterministic atom-cavitycoupling, see Chapter 4.


16 Chapter 1: Tools <strong>for</strong> single atom c<strong>on</strong>trol(a)10 000counts10 µm0(b)5 40010 µmcounts0(c)1 50010 µmcountsFigure 1.8: (a) CCD image of two atoms in the MOT with an exposure time of 1 s.The rms radius of the MOT inferred from the image is about 5 µm. (b) CCD image oftwo atoms in the DT illuminated by the optical molasses. (c) Five atoms in the DT.0Intensified CCD cameraIn<strong>for</strong>mati<strong>on</strong> <strong>on</strong> the positi<strong>on</strong> and/or distributi<strong>on</strong> of the atoms in the trap is obtainedfrom their fluorescence image recorded by a CCD camera (Roper Scientific, PI-MAX:1K).Combined with the image intensifier (Roper Scientific, GEN III HQ) it has a quantumefficiency of approximately 10 % at 852 nm. One detected phot<strong>on</strong> generates <strong>on</strong> average350 counts <strong>on</strong> the CCD chip (1024 × 1024 pixel large), and <strong>on</strong>e 13 µm × 13 µm CCD pixelcorresp<strong>on</strong>ds to 0.9328(±0.0004) µm in the object plane (see Sec. 2.6). The read-out noiseis about 80 counts per pixel. For more details <strong>on</strong> the ICCD camera see Ref. [61].Illuminati<strong>on</strong> of atomsDue to the low scattering rate in the far-detuned dipole trap, direct imaging of trappedatoms in the DT is not possible. There<strong>for</strong>e, we illuminate them with near-res<strong>on</strong>ant threedimensi<strong>on</strong>aloptical molasses <strong>for</strong>med by the MOT cooling lasers acting like a flash ofa c<strong>on</strong>venti<strong>on</strong>al camera. Besides the illuminati<strong>on</strong>, the molasses provides cooling of thetrapped atoms counteracting heating through phot<strong>on</strong> scattering, resulting in an atomtemperature of about 70 µK. Note that this cooling also counteracts the heating of atoms


1.6 C<strong>on</strong>clusi<strong>on</strong> 17due to phase-noise of the frequency generator driving our “optical c<strong>on</strong>veyor belt” <strong>for</strong>transporting atoms, see Sec. 2.4. Intensity and detuning of the illuminating beams areoptimized to exclude losses of atoms, to prevent them from hopping between differentpotential wells of the DT, and to provide high c<strong>on</strong>trast of the imaged atoms [55]. As aresult, we can n<strong>on</strong>-destructively observe single atoms <strong>for</strong> about half a minute limited <strong>on</strong>lyby atom losses due to background gas collisi<strong>on</strong>s [45].Figure 1.8(b) shows an image of two atoms in the DT with an exposure time of1 s. The horiz<strong>on</strong>tal width of the fluorescence spot of 1.5 µm rms is determined by theresoluti<strong>on</strong> of the imaging optics. It is larger than the periodicity λ/2 ≈ 0.5 µm of thestanding-wave trap and thus does not allow us to resolve atoms trapped in adjacentpotential wells. The vertical width of the fluorescence spot, i.e. perpendicular to the DTaxis, is essentially defined by the spread of the Gaussian thermal wave packet of the atomin the radial directi<strong>on</strong> of the trap due to its radial oscillati<strong>on</strong>. In Sec. 2.2 I describe howwe infer the positi<strong>on</strong> of optically resolved atoms al<strong>on</strong>g the trap axis with sub-micrometerprecisi<strong>on</strong> by analyzing their fluorescence image.1.6 C<strong>on</strong>clusi<strong>on</strong>In this Chapter I have presented a set of tools <strong>for</strong> preparing and c<strong>on</strong>trolling single caesiumatoms required <strong>for</strong> our experiments. A high-gradient magneto-optical trap cools atomsfrom background gas in a vacuum glass cell and c<strong>on</strong>fines them to a volume of about 5 µmradius. The exact number of trapped atoms is then inferred from the MOT fluorescencelight.To provide a c<strong>on</strong>servative trap required <strong>for</strong> any experiments <strong>on</strong> atomic coherence,including cavity <strong>QED</strong> measurements, we use a standing-wave dipole trap allowing us tostore the atoms <strong>for</strong> l<strong>on</strong>g times without significant phot<strong>on</strong> scattering and mixing of quantumstates. Besides, its standing-wave structure provides a tight sub-micrometer c<strong>on</strong>finementof atoms al<strong>on</strong>g the trap axis and enables us to transport the atoms, as will be describedin Chapter 2. The atoms can be transferred between the two trap with almost unityefficiency. Molasses cooling with the MOT beams allows us to image single atoms in thedipole trap using an intensified CCD camera. This gives us spatial in<strong>for</strong>mati<strong>on</strong> <strong>on</strong> theatoms necessary <strong>for</strong> deterministic atom manipulati<strong>on</strong>.To overcome the Poiss<strong>on</strong>ian statistics of the atom loading and to prepare a desirednumber of atoms in the DT <strong>for</strong> a planned experiment, we have implemented a feedbackmechanism <strong>for</strong> the number-triggered loading of the dipole trap. It repeatedly loads theMOT until the desired number of atoms is captured and <strong>on</strong>ly then transfers them into thedipole trap. In Sec. 4.5 we use this method <strong>for</strong> the efficient preparati<strong>on</strong> of <strong>on</strong>e atom inthe dipole trap.


18 Chapter 1: Tools <strong>for</strong> single atom c<strong>on</strong>trol


Chapter 2Submicrometer positi<strong>on</strong> c<strong>on</strong>trolof single atoms2.1 Introducti<strong>on</strong>Precisi<strong>on</strong> positi<strong>on</strong> measurement and localizati<strong>on</strong> of atoms is of great interest <strong>for</strong> numerousapplicati<strong>on</strong>s and has been achieved in and <strong>on</strong> solids using, e.g., scanning tunnellingmicroscopy [43], atomic <strong>for</strong>ce microscopy [62], or electr<strong>on</strong> energy-loss spectroscopy imaging[63]. However, if the applicati<strong>on</strong> requires l<strong>on</strong>g coherence times, as is the case inquantum in<strong>for</strong>mati<strong>on</strong> processing or <strong>for</strong> frequency standards, the atoms should be well isolatedfrom their envir<strong>on</strong>ment. This situati<strong>on</strong> is realized <strong>for</strong> i<strong>on</strong>s in i<strong>on</strong> traps, freely movingneutral atoms, or neutral atoms trapped in optical dipole traps. For the case of i<strong>on</strong>s,positi<strong>on</strong>s [64, 65] and distances [66] have been optically measured and c<strong>on</strong>trolled with asub-optical wavelength precisi<strong>on</strong>. Similar precisi<strong>on</strong> has been reached in an all-optical positi<strong>on</strong>measurement of freely moving atoms [67]. Dipole traps, operated as optical tweezers,have been used to precisely c<strong>on</strong>trol the positi<strong>on</strong> of individual neutral atoms [23, 68]. To myknowledge, however, a nanometric positi<strong>on</strong> and distance measurement has so far not beenachieved in this case. Such a c<strong>on</strong>trol of the relative and absolute positi<strong>on</strong> of single trappedneutral atoms, however, is an important prerequisite <strong>for</strong> cavity quantum electrodynamicsexperiments, allowing us to deterministically place an atom precisely at the point of themaximum atom-phot<strong>on</strong> coupling inside the cavity.In this chapter I describe the methods developed <strong>for</strong> measuring and c<strong>on</strong>trolling thepositi<strong>on</strong> of single neutral atoms stored in a standing-wave optical dipole trap. The mainresults presented here are published in Ref. [25]. The positi<strong>on</strong>s of the atoms are inferredfrom their fluorescence using high resoluti<strong>on</strong> imaging optics in combinati<strong>on</strong> with an intensifiedCCD camera. The absolute positi<strong>on</strong> of individual atoms al<strong>on</strong>g the DT is measuredwith a precisi<strong>on</strong> of 143 nm, see Sec. 2.2. The relative positi<strong>on</strong> of the atoms, i.e. theirseparati<strong>on</strong>, is determined more accurately by averaging over many measurements, yieldingan uncertainty of 21 nm, see Sec. 2.3. Due to this high resoluti<strong>on</strong>, we can resolve thediscreteness of the distributi<strong>on</strong> of interatomic distances in the standing-wave potential,even though our DT is <strong>for</strong>med by a Nd:YAG laser with potential wells separated by <strong>on</strong>ly19


20 Chapter 2: Submicrometer positi<strong>on</strong> c<strong>on</strong>trol of single atomsλ/2 = 532 nm. This resoluti<strong>on</strong> allows us to determine the exact number of potentialwells between simultaneously trapped atoms. Combining an initial positi<strong>on</strong> measurementwith a c<strong>on</strong>trolled transport of single atoms over macroscopic distances by means of our“optical c<strong>on</strong>veyor belt” technique [23, 24], described in Sec. 2.4, we transport individualatoms to a predetermined positi<strong>on</strong> al<strong>on</strong>g the DT axis with an accuracy of 300 nm, therebydem<strong>on</strong>strating a high degree of c<strong>on</strong>trol of the absolute positi<strong>on</strong> of an atom, see Sec. 2.5.2.2 Measurement of the positi<strong>on</strong> of an atomIn order to obtain fluorescence images of the atoms in the DT, we illuminate them witha near-res<strong>on</strong>ant three-dimensi<strong>on</strong>al optical molasses, see Sec. 1.5 and Fig. 2.1(a). AnICCD image of a single atom stored in the DT with an exposure time of 1 s is shownin Fig. 2.1(b). This exposure time is much l<strong>on</strong>ger than the timescale of the thermalpositi<strong>on</strong> fluctuati<strong>on</strong>s of the atom inside the trap. There<strong>for</strong>e, the vertical width of thefluorescence spot, i.e. perpendicular to the DT axis, is essentially defined by the spread ofthe Gaussian thermal wave packet of the atom in the radial directi<strong>on</strong> of the trap, whichdepends <strong>on</strong> the trap depth and the atom temperature. 1 In the axial directi<strong>on</strong> of the DT,the wave packet has a much smaller 1/ √ e-halfwidth of <strong>on</strong>ly ∆x therm = 35–50 nm <strong>for</strong> thetypical depth of the DT of 1–2 mK. In additi<strong>on</strong> to these thermal fluctuati<strong>on</strong>s, the axialpositi<strong>on</strong> of the standing wave itself is fluctuating by σ fluct (1 s) = 42(±13) nm during the1 s exposure time due to drifts and acoustic vibrati<strong>on</strong>s of the optical setup (see below).The horiz<strong>on</strong>tal 1/ √ e-halfwidth of the detected fluorescence peak, w ax = 1.30(±0.15) µm,is much larger and is caused by diffracti<strong>on</strong> within the imaging optics and a slight blurringin the intensificati<strong>on</strong> process of the ICCD [51]. Compared to the point spread functi<strong>on</strong>of our imaging system, ∆x therm and σ fluct have thus a negligible effect <strong>on</strong> w ax . In thefollowing I define the atomic positi<strong>on</strong>s as the center of the Gaussian thermal wave packetsof the atoms.Line spread functi<strong>on</strong>The ICCD image is characterized by its intensity distributi<strong>on</strong> I(˜x, ỹ), where ˜x and ỹare spatial coordinates <strong>on</strong> the CCD ship. It is related to the real object O(x, y), i.e. tothe distributi<strong>on</strong> of an incoherent light source in the object plane of the camera, by thec<strong>on</strong>voluti<strong>on</strong> with the point spread functi<strong>on</strong>, P SF (˜x, ỹ; x, y), of the imaging system asfollows (see e.g. [69, 70])∫∫I(˜x, ỹ) = P SF (˜x, ỹ; x, y) O(x, y) dx dy + N(˜x, ỹ), (2.1)1 In the inverse problem the atom temperature can be determine by direct imaging the atom in the trapand measuring its spatial distributi<strong>on</strong> [51].


2.2 Measurement of the positi<strong>on</strong> of an atom 21(a)dipoletraplaseroptical molassesdipoletraplaser(b)Counts / pixel0 32005 m(c)24Counts/1000120x atomFigure 2.1: Determinati<strong>on</strong> of the positi<strong>on</strong> of a single trapped atom. (a) An atom storedin the standing-wave dipole trap is illuminated with an optical molasses (schematic drawing).(b) ICCD image of <strong>on</strong>e atom stored in the DT with an exposure time of 1 s. Theobserved fluorescence spot corresp<strong>on</strong>ds to about 200 detected phot<strong>on</strong>s. (c) To determinethe positi<strong>on</strong> of the atom al<strong>on</strong>g the DT axis, the pixel counts are binned in the verticaldirecti<strong>on</strong>. In order to reduce background noise, the summati<strong>on</strong> is <strong>on</strong>ly per<strong>for</strong>med overthe narrow vertical regi<strong>on</strong> shown by dashed lines in (b). The thick solid line corresp<strong>on</strong>dsto the line spread functi<strong>on</strong> of our imaging optics and reveals the absolute positi<strong>on</strong> x atomof the atom. The thin lines are two Gaussians composing the fit functi<strong>on</strong>, see text <strong>for</strong>details.where N(˜x, ỹ) denotes additive noise. The trapping potentials as seen from the side, andthus our time-averaged fluorescing atoms, have a slit-like <strong>for</strong>m. 2 Since we are interested<strong>on</strong>ly in the horiz<strong>on</strong>tal positi<strong>on</strong> ˜x atom of the fluorescence peak <strong>on</strong> the ICCD image, we bin2 The extent of the atomic cloud al<strong>on</strong>g the optical axis is within the depth of focus of about ±6 µm ofthe imaging system [51].


22 Chapter 2: Submicrometer positi<strong>on</strong> c<strong>on</strong>trol of single atomsI(˜x i , ỹ j ) in the vertical directi<strong>on</strong> and getI(˜x i ) = ∑ jI(˜x i , ỹ j ), (2.2)with ˜x i and ỹ j denoting the horiz<strong>on</strong>tal and vertical positi<strong>on</strong> of pixel {i, j}, respectively.Then, equati<strong>on</strong> (2.1) can be reduced to the <strong>on</strong>e-dimensi<strong>on</strong>al <strong>on</strong>e:∫I(˜x) = LSF (˜x, x) O(x) dx + N(˜x), (2.3)where the imaging process is characterized by the line spread functi<strong>on</strong> (LSF) of our imagingoptics. For the slit-like object the functi<strong>on</strong> O(x) is a δ-functi<strong>on</strong> at positi<strong>on</strong> x atom . If theimaging system is space invariant, the LSF depends <strong>on</strong>ly <strong>on</strong> the difference (˜x − ˜x atom ).Neglecting noise <strong>for</strong> the moment, this yieldsI(˜x) ∝ LSF (˜x − ˜x atom ) . (2.4)The object coordinate x atom and the image coordinate ˜x atom are c<strong>on</strong>nected by therelati<strong>on</strong>x atom = ˜x atom − Õx , (2.5)Mwhere Õx is the image coordinate of the origin and M is the magnificati<strong>on</strong> of our imagingoptics. In general, Õx and M have to be calibrated from independent measurements. Theprecise calibrati<strong>on</strong> of M is described in Sec. 2.6. The choice of Õx str<strong>on</strong>gly depends <strong>on</strong> theplanned experiment. For instance, in the experiments <strong>on</strong> addressing individual atoms ina magnetic field gradient [18, 60], the origin can be defined as the point of zero magneticfield. In the present case, however, no physical point in space is singled out as an origin,there<strong>for</strong>e we arbitrarily set Õx ≡ 0, i.e. we assign the origin to the left-most CCD pixel.In principle, the LSF could be determined by modelling the imaging process. However,the modelling requires the precise knowledge of the properties of all optical elements inthe optical path as well as of their positi<strong>on</strong>s, which is a n<strong>on</strong>-trivial task if we aim ata satisfactory result. There<strong>for</strong>e, we determine the <strong>for</strong>m of our LSF experimentally byanalyzing about 100 images of single atoms. We have found that the LSF is positi<strong>on</strong>independent,which proves the assumpti<strong>on</strong> of the space-invariant imaging system. It iswell approximated by a sum of two Gaussians with a ratio of 4.4:1 in heights and 1:3.2 inwidths, with a slight horiz<strong>on</strong>tal offset with respect to each other, shown in Fig. 2.1(c). Inthe following, I define ˜x atom as the positi<strong>on</strong> of the maximum of this LSF. In our experiment,it is determined by fitting a simple Gaussian to the fluorescence peak. This procedure hasbeen chosen because it can be carried out in a fast automated way, yielding in<strong>for</strong>mati<strong>on</strong>about the atomic positi<strong>on</strong> during the running experimental sequence. The study <strong>on</strong> theprecise analysis of binned, normally distributed data per<strong>for</strong>med in Ref. [71] shows thatthe simple Gaussian fit gives the undistorted in<strong>for</strong>mati<strong>on</strong> <strong>on</strong> the peak positi<strong>on</strong> as l<strong>on</strong>g asthe peak covers at least three bins, which is the case here. In additi<strong>on</strong>, our simulati<strong>on</strong>staking into account the experimental LSF and the finite bin size show a c<strong>on</strong>stant positi<strong>on</strong>offset of 42 nm of the fitted center of the Gaussian with respect to the maximum of theLSF, due to the slight asymmetry of our LSF. This offset <strong>on</strong>ly leads to a global shift ofÕ x and is thus irrelevant <strong>for</strong> our analysis.


2.2 Measurement of the positi<strong>on</strong> of an atom 23Statistical errorSince our ICCD is a phot<strong>on</strong> counting device, the basic noise source in determining x atomis the Poiss<strong>on</strong>ian fluctuati<strong>on</strong>s of the number of phot<strong>on</strong>s detected per pixel. The statisticalerror of the mean positi<strong>on</strong> of spatially normally distributed counts <strong>on</strong> the CCD image isgiven by∆x stat =w ax√ , (2.6)Nphwhere N ph is the number of phot<strong>on</strong>s detected per atom. Strictly speaking, equati<strong>on</strong> (2.6)is <strong>on</strong>ly valid <strong>for</strong> a spatial Gaussian distributi<strong>on</strong> of detected phot<strong>on</strong>s with an infinitelysmall bin size. In our experiments with the n<strong>on</strong>-Gaussian LSF and the finite bin size wedetermine the positi<strong>on</strong> and the width of the peak by fitting the binned counts with theGaussian functi<strong>on</strong>. To find the statistical error in this case, we have per<strong>for</strong>med numericalsimulati<strong>on</strong> taking into account the experimental LSF and the finite bin size. We havesimulated the arrival of N ph phot<strong>on</strong>s <strong>on</strong>to the CCD chip with the LSF distributi<strong>on</strong>, binnedthem with an arbitrary shift of the bin mask relative to the simulated peak, fitted thebinned counts with the Gaussian functi<strong>on</strong>, and determined the deviati<strong>on</strong> of the fittedpeak center from the real <strong>on</strong>e. The whole procedure was repeated many times to reducestatistical error. The spread of the measured deviati<strong>on</strong>s gave us the statistical error of thepositi<strong>on</strong> detecti<strong>on</strong>, which is 1.44 times larger than that given by expressi<strong>on</strong> (2.6). Thus,our method allows us to determine the fitted peak center x atom with a statistical error of∆x stat = 1.44 w ax√Nph. (2.7)In the experiment the value of N ph depends <strong>on</strong> the illuminati<strong>on</strong> parameters. Here, N ph =200(±30) phot<strong>on</strong>s per sec<strong>on</strong>d per atom, so that ∆x stat = 130(±20) nm.Background noiseIn additi<strong>on</strong> to the statistical error, two further sources influence the precisi<strong>on</strong> of the positi<strong>on</strong>detecti<strong>on</strong>: the background noise of our ICCD image, described in detail in Ref. [61],and the positi<strong>on</strong> fluctuati<strong>on</strong>s of the DT. The background in Fig. 2.1 originates in aboutequal proporti<strong>on</strong>s from stray light and from the read-out process of the ICCD, yielding atotal offset of 2300(±300) counts per bin <strong>for</strong> 1 s exposure time. The numerical analysis ofthe influence of this background <strong>on</strong> the measured atom positi<strong>on</strong> ∆x atom reveals that thenoise of 300 counts per bin introduces an additi<strong>on</strong>al uncertainty of ∆x backgr = 15 nm tothe fitted peak center.Fluctuati<strong>on</strong>s of the trapThe atom positi<strong>on</strong> is subject to positi<strong>on</strong> fluctuati<strong>on</strong>s of the interference pattern of theDT, σ fluct . Since σ fluct cannot be extracted from the ICCD image, we determine it in anindependent measurement. For this purpose, we mutually detune the two trap beams by1.2 MHz and overlap them <strong>on</strong> a fast photodiode as shown in Fig. 2.2(a). From the phase


24 Chapter 2: Submicrometer positi<strong>on</strong> c<strong>on</strong>trol of single atomsAOM2 x 104 MHzbeat signal1.2 MHzfastphoto-diodeLock-in amplifierSR850phasephase fluctuati<strong>on</strong>sglasswedgeAOM2 x 104.6 MHzNd:YAGlaser(a) Experimental setup.150100Phase [deg]500-50-100-1500 1 2 3 4 5Time [s](b) Typical phase variati<strong>on</strong> of the trap.Figure 2.2: Measurement of the phase variati<strong>on</strong>s of the standing-wave dipole trap.of the resulting beat note we infer the phase variati<strong>on</strong>s φ(t) of the standing wave witha 300 kHz bandwidth, see Fig. 2.2(b). The standard deviati<strong>on</strong> of φ(t) during the timeinterval τ, σ φ (τ), is directly related to the positi<strong>on</strong> fluctuati<strong>on</strong>s of the DT during thisperiod by σ fluct (τ) = λ/2 · σ φ (τ)/2π. We have found that σ fluct (1 s) = 42(±13) nm. Thetypical time dependence of φ(t) in Fig. 2.2(b) shows both fast fluctuati<strong>on</strong>s and slow drifts.However, most traces of φ(t) analyzed do not have significant drifts <strong>on</strong> the time scale of


2.3 Measurement of the separati<strong>on</strong> between two atoms 25<strong>on</strong>e sec<strong>on</strong>d. Thus, neglecting linear drifts in the following and using the approximati<strong>on</strong>of Gaussian-distributed positi<strong>on</strong> fluctuati<strong>on</strong>s, which we have checked to be valid to betterthan 1 % in our case, the positi<strong>on</strong> uncertainty immediately after the 1 s exposure time isgiven by∆x 2 atom(1 s) = ∆x 2 stat + ∆x 2 backgr + σ2 fluct (1 s), (2.8)yielding ∆x atom (1 s) = 140(±20) nm.Finally, the read-out of the image and the data analysis take an additi<strong>on</strong>al 0.5 s duringwhich x atom is further subject to positi<strong>on</strong> fluctuati<strong>on</strong>s of the DT. This increases the errorof the positi<strong>on</strong> measurement by √ 2σ fluct (0.5 s) = 29 nm, 3 which was determined similarto σ fluct (1 s). Thus, we can determine the absolute positi<strong>on</strong> of the trapped atom with aprecisi<strong>on</strong> of∆x atom (1.5 s) = 143(±20) nm (2.9)within 1.5 s (1 s exposure time plus 0.5 s read-out and data analysis). Our analysis showsthat this precisi<strong>on</strong> cannot be significantly increased by extending the exposure time,because the benefit of the higher phot<strong>on</strong> statistics <strong>for</strong> l<strong>on</strong>ger times is counteracted by theincreasing influence of the DT drifts.Summarizing, the main sources of noise influencing the precisi<strong>on</strong> of the positi<strong>on</strong> detecti<strong>on</strong>are listed below in Table 2.1.Source of errorValuestatistical error ∆x stat 130(±20) nmCCD background noise ∆x backgr 15 nmDT positi<strong>on</strong> fluctuati<strong>on</strong>s σ fluct (1.5 s) 51(±15) nmTable 2.1: Sources of noise influencing the precisi<strong>on</strong> of the positi<strong>on</strong> detecti<strong>on</strong>.2.3 Measurement of the separati<strong>on</strong> between two atomsWhile <strong>for</strong> some applicati<strong>on</strong>s the absolute positi<strong>on</strong> of the atoms must be known to the highestpossible precisi<strong>on</strong>, other experiments, like, e.g., c<strong>on</strong>trolled cold collisi<strong>on</strong>s between twoatoms [72] possibly with subsequent generati<strong>on</strong> of the entangled cluster states of manyatoms [73, 74], require a precise knowledge of the separati<strong>on</strong> d between atoms. In thefollowing I show that in our experiment this separati<strong>on</strong> can be determined more preciselythan the absolute positi<strong>on</strong> of the individual atoms. The reas<strong>on</strong> is that DT positi<strong>on</strong> fluctuati<strong>on</strong>sequally affect all simultaneously trapped atoms and there<strong>for</strong>e do not change theseparati<strong>on</strong> between them. Thus, this distance can be averaged over many measurements3 We have introduced a factor of √ 2, since we are interesting not in the distributi<strong>on</strong> of data pointsaround the mean value during the c<strong>on</strong>sidered time interval, but in the deviati<strong>on</strong> of the very last point fromthe very first <strong>on</strong>e.


26 Chapter 2: Submicrometer positi<strong>on</strong> c<strong>on</strong>trol of single atoms(a)d = n /2(b)Counts / pixel0 32005 m(c)Counts/100024120dFigure 2.3: Determinati<strong>on</strong> of the distance between atoms. (a) Two atoms in thestanding-wave dipole trap have a separati<strong>on</strong> of nλ/2 with n integer. (b) After loadingthe atoms into the DT, we take camera pictures of them. (c) From each picturewe determine the positi<strong>on</strong>s of the atoms and their separati<strong>on</strong> d. Averaging over manymeasurements of d <strong>for</strong> the same pair of atoms reduces the statistical error and allows usto infer n.providing much higher precisi<strong>on</strong> than <strong>for</strong> the absolute positi<strong>on</strong>. As a result, the obtainedprecisi<strong>on</strong> allows us to resolve the standing-wave structure of our DT and to determine theexact number of potential wells between two optically resolved atoms, a situati<strong>on</strong> that sofar seemingly required much l<strong>on</strong>ger (e.g., CO 2 ) trapping laser wavelengths [75].Distance measurement from a single imageTo study the atomic separati<strong>on</strong>s in our dipole trap, we first load it with two atoms from theMOT, see Fig. 2.3(a). Normally, the atoms transferred from the MOT are distributed overσ MOT = 5 µm al<strong>on</strong>g the DT axis, resulting in a spread of the distances of √ 2σ MOT ≈ 7 µm.In order to increase the mean separati<strong>on</strong> between the atoms we let them freely expandal<strong>on</strong>g the dipole trap by switching off <strong>on</strong>e of the DT laser beams <strong>for</strong> 0.5 ms eliminatingthe standing-wave structure <strong>for</strong> that time interval. Following this expansi<strong>on</strong> the atoms arenow distributed over roughly 50 µm in the standing-wave trap.Next, we take a camera picture of the atom pair, see Fig. 2.3(b). In these experiments,we detect N ph = 270(±30) phot<strong>on</strong>s per sec<strong>on</strong>d per atom. From each picture we determinethe distance d between the atoms, see Fig. 2.3(c). For this purpose, we fit the fluorescence


2.3 Measurement of the separati<strong>on</strong> between two atoms 27300250Number of atom pairs200150100500-0.5 -0.4 -0.2 0.0 0.2 0.4 0.5Atomic separati<strong>on</strong>s modulo λ/2 [λ/2]Figure 2.4: Histogram of the measured distances modulo λ/2 (see equati<strong>on</strong> (2.10). Thethick curve is a fit to the data with a sum of three equal Gaussians at the positi<strong>on</strong>s 0and ±λ/2. The individual Gaussians are shown as thin curves. Their width of ∆d =132(±5) nm determines the precisi<strong>on</strong> of the distance measurement.peaks with the experimentally established line spread functi<strong>on</strong>. For the case of partiallyoverlapping fluorescence spots (d 10 µm), this method yields more precise results <strong>for</strong>the two positi<strong>on</strong>s than fitting simple Gaussians. For d 4 µm the increasing overlapreduces the precisi<strong>on</strong> of the positi<strong>on</strong> determinati<strong>on</strong>. We have there<strong>for</strong>e restricted ourinvestigati<strong>on</strong>s to the case where the atoms are separated by more than 4 µm. After all,we have accumulated more than 3000 images with two atoms, each providing us with ameasured atomic separati<strong>on</strong> d i .Being trapped in the standing-wave potential, the atoms are always separated by adistance equal to an integer multiple of λ/2. Using this fact, we c<strong>on</strong>struct a new data set{d ′ i } by applying the following rule:d ′ i ={di mod λ/2 , if 0 ≤ d i mod λ/2 < λ/4d i mod λ/2 − λ/2, if λ/4 ≤ d i mod λ/2 < λ/2(2.10)Here, in order to get the distributi<strong>on</strong> symmetric around zero, I calculate the data pointsd i modulo λ/2 and shift them by λ/2 if d i mod λ/2 > λ/4. The histogram of the newdistributi<strong>on</strong> {d ′ i } with a bin size of λ/40 is shown in Fig. 2.4. It is centered at 0 and isc<strong>on</strong>fined in a range (−λ/4, λ/4). 4 In the case of a perfect distance measurement, that iswith zero uncertainty, the distributi<strong>on</strong> {d ′ i } should have a δ-functi<strong>on</strong>-like <strong>for</strong>m with all d′ i4 The precise calibrati<strong>on</strong> of the image scale is per<strong>for</strong>med in Sec. 2.6 by centering the histogram of themeasured distances at zero.


28 Chapter 2: Submicrometer positi<strong>on</strong> c<strong>on</strong>trol of single atomsequal zero. On the other hand, if the precisi<strong>on</strong> of the distance measurement is much worsethan the λ/2 periodicity of the trap, the c<strong>on</strong>structed distributi<strong>on</strong> should appear flat.In order to extract the distance uncertainty from {d ′ i }, we fit the histogram of Fig. 2.4with a sum of three equal Gaussians (i.e. of the same height and width) centered atpositi<strong>on</strong>s −λ/2, 0, and λ/2, respectively. The individual Gaussian functi<strong>on</strong>s are shownin Fig. 2.4 as the thin lines, whereas their sum is shown by the thick line. The centralGaussian distributi<strong>on</strong> corresp<strong>on</strong>ds to those measured separati<strong>on</strong>s d measi , from which theactual separati<strong>on</strong>s d actuali between atoms can be found as the nearest multiple of λ/2to d measi . This means that the separati<strong>on</strong> d actuali can be determined correctly <strong>for</strong> thesemeasurement events. The two other Gaussians centered at ±λ/2 denote events <strong>for</strong> whichd actuali will be determined wr<strong>on</strong>g by ±λ/2, respectively. C<strong>on</strong>sequently, the reliability of ourdistance measurement is found by integrating the normalized central Gaussian functi<strong>on</strong>from −λ/4 to λ/4 and thus equals 95.6 %. There<strong>for</strong>e, <strong>on</strong>ly in 4.4 % of cases the determinedseparati<strong>on</strong> is wr<strong>on</strong>g by <strong>on</strong>e potential well.The precisi<strong>on</strong> of our distance measurement is given by the width of the fitted Gaussiansof ∆d meas = 132(±5) nm. Since the DT positi<strong>on</strong> fluctuati<strong>on</strong>s does not affect thedistance measurement, the uncertainty of the separati<strong>on</strong> d between two atoms determinedfrom <strong>on</strong>e picture should be ∆d expect = √ 2 (∆x 2 stat + ∆x 2 backgr )1/2 = 160(±20) nm. Here,∆x stat is calculated from equati<strong>on</strong> (2.7) <strong>for</strong> N ph = 270 phot<strong>on</strong>s. The observed discrepancybetween the measured and expected values may come from smaller DT fluctuati<strong>on</strong>sduring taking data as compared to previous experimental days as well as from differentilluminati<strong>on</strong> parameters resulting in a larger number of detected phot<strong>on</strong>s per atom, thusreducing ∆x stat .Resolving the standing-wave structureThe advantage of the distance measurement compared to the positi<strong>on</strong> detecti<strong>on</strong> is thatthe atomic separati<strong>on</strong>s are not influenced by the fluctuati<strong>on</strong>s of the dipole trap. Thisallows us to reduce the statistical error in the determinati<strong>on</strong> of d by averaging the resultsfrom N pic images of the same atom pair. The reduced statistical error of the mean value¯d then reads∆ ¯d = √ ∆d . (2.11)NpicFor instance, by averaging over <strong>on</strong>ly two images the reliability of the distance measurementcan be increased up to 99.6 %, which is enough <strong>for</strong> most of the feasible experiments withsingle atoms.In the next experiment we determine the atomic separati<strong>on</strong> by averaging over morethan 9 images of the same atoms. In this case, the probability of a wr<strong>on</strong>g determinati<strong>on</strong> ofn should theoretically be less than 2.0×10 −10 . The measurement procedure is the same asbe<strong>for</strong>e, but instead of <strong>on</strong>e picture, we take at least N pic = 10 successive camera pictures ofthe same atom pair be<strong>for</strong>e <strong>on</strong>e of the two atoms leaves the trap. For each pair of atoms wethen calculate the mean value ¯d and the standard deviati<strong>on</strong> ∆d of the measured distances.Since d can now be measured with a precisi<strong>on</strong> ∆ ¯d ≪ λ/2, its distributi<strong>on</strong> should there<strong>for</strong>ereveal the standing-wave structure of the DT. Indeed, the λ/2 period is strikingly apparent


2.3 Measurement of the separati<strong>on</strong> between two atoms 29100Cumulative distributi<strong>on</strong>of atomic separati<strong>on</strong>s8060402008 12 16 20 24 28 32 36 40 44 48 52 56 60 64 68Mean distance d between atoms [l/2]Figure 2.5: Cumulative distributi<strong>on</strong> of separati<strong>on</strong>s between atoms in the dipole trapmeasured with the scheme presented in Fig. 2.3. In order to resolve the periodic structureof the trap, we reduce the statistical error in the measurement of the atomic separati<strong>on</strong> byaveraging over more than 9 distance measurements <strong>for</strong> each atom pair. The discretizati<strong>on</strong>of the distances to nλ/2 is clearly visible in the data. Our resoluti<strong>on</strong> is thus sufficient todetermine the exact number of potential wells between any two optically resolved atoms.in Fig. 2.5 which shows the cumulative distributi<strong>on</strong> of averaged distances ¯d between atomsmeasured with about 100 atom pairs.By analogy with single distance measurements, we can now map the measured { ¯d i } tothe new data set { ¯d ′ i } using equati<strong>on</strong> (2.10). The corresp<strong>on</strong>ding peaked histogram witha bin size of λ/40 is presented in Fig. 2.6. As compared to the distributi<strong>on</strong> in Fig. 2.4c<strong>on</strong>structed from single distance measurements, the averaging has significantly improvedthe resoluti<strong>on</strong> of our distance measurement, which can be directly inferred from the finitewidth of the histogram, yielding ∆ ¯d meas = 21(±1) nm. The expected value ∆ ¯d expect <strong>for</strong>the measurement precisi<strong>on</strong> can be found using equati<strong>on</strong> (2.11). In the current experimentthe mean value of {1/ √ N pic, i } = 0.198, where N pic, i denotes the number of pictures taken<strong>for</strong> the i-th atom pair. Using ∆d meas = 132(±5) nm, this results in ∆ ¯d expect = 26(±1) nm,which agrees reas<strong>on</strong>ably with ∆ ¯d meas .Discussi<strong>on</strong>Due to the thermal positi<strong>on</strong> fluctuati<strong>on</strong>s of the atoms inside the potential wells, theinstantaneous distance d inst between the atoms is known with a lower precisi<strong>on</strong> of∆d inst = (∆ ¯d 2 + 2∆x 2 therm )1/2 . In the axial directi<strong>on</strong> of the DT, the Gaussian thermal


30 Chapter 2: Submicrometer positi<strong>on</strong> c<strong>on</strong>trol of single atoms3025Number of atom pairs20151050-0.5 -0.4 -0.2 0.0 0.2 0.4 0.5Atomic separati<strong>on</strong>s modulo λ/2 [λ/2]Figure 2.6: Histogram of the atomic separati<strong>on</strong>s averaged over more than 9 measurementseach and calculated modulo λ/2. The data analyzed are the <strong>on</strong>e from Fig. 2.5.wave packet of the atom has a 1/ √ e-halfwidth of about ∆x therm = 35–50 nm, depending<strong>on</strong> the depth of the DT and the atomic temperature. In our case ∆ ¯d < √ 2∆x therm , sothat we have reached the fundamental limit <strong>for</strong> determining d inst . This limit is intrinsicto our method because the back-acti<strong>on</strong> of the measurement does result in heating of theatoms to a finite temperature of approximately 70 µK due to phot<strong>on</strong> scattering.Finally, since we can resolve the standing-wave structure of our DT and determinethe number n of potential wells between any optically resolved trapped atoms, we knowthe separati<strong>on</strong> between the centers of their thermal wave packets exactly, even despite offinite measured width ∆ ¯d of the distributi<strong>on</strong> { ¯d ′ i }. Note that it is sufficient to average themeasured atomic separati<strong>on</strong> over two measurements to achieve this absolute precisi<strong>on</strong>.2.4 The optical c<strong>on</strong>veyor beltMany planned experiments, and first of all those <strong>on</strong> accurate placing a single atom into themode of a high-finesse cavity, require a microscopic positi<strong>on</strong> c<strong>on</strong>trol of the trapped atomover macroscopic distances. For this purpose we use an optical c<strong>on</strong>veyor belt technique[23, 24], allowing us to move the atom al<strong>on</strong>g the trap axis over millimeter distances with asub-micrometer precisi<strong>on</strong> [25]. Since the atoms in our experiments are held at the intensitymaxima of the optical light field, created by interfering two counter-propagating laserbeams, their transportati<strong>on</strong> can be achieved by moving the corresp<strong>on</strong>ding standing-waveinterference structure. In our experiment this moti<strong>on</strong> is realized by mutually detuning thefrequencies of the two trapping lasers.


2.4 The optical c<strong>on</strong>veyor belt 31velocitypositi<strong>on</strong> ( detuning) accelerati<strong>on</strong>a0-av max0d0Figure 2.7: Time dependence of the main transportati<strong>on</strong> parameters. In order to transportan atom over the distance d, we expose it to c<strong>on</strong>stant accelerati<strong>on</strong> and decelerati<strong>on</strong> a.The velocity is ramped from 0 to v max = at d /2 and back to 0.t dtimeA moving standing waveOur “optical c<strong>on</strong>veyor belt” c<strong>on</strong>sists of a standing-wave dipole trap made of two counterpropagatinglaser beams. Moti<strong>on</strong> of the interference pattern al<strong>on</strong>g the trap axis is achievedby mutually detuning the laser beams. The time-dependent potential of a standing-wavetrap made of two Gaussian beams of frequencies ν 1 and ν 2 readsw 2 [0U(x, y, z, t) = −U 0w 2 (x) exp − 2(y2 + z 2 )w 2 (x)]cos 2 (π∆νt − 2π λ x )(2.12)with ∆ν = ν 1 − ν 2 the frequency difference between the two beams. If the two frequenciesare equal, ∆ν = 0, we get a stati<strong>on</strong>ary standing-wave dipole trap of equati<strong>on</strong> (1.13).The mutual detuning ∆ν ≠ 0 causes the standing wave to move al<strong>on</strong>g its optical axisat the c<strong>on</strong>stant velocity ofv = λ ∆ν2 . (2.13)To smoothly transport a trapped atom over a distance d, the interference pattern is firstuni<strong>for</strong>mly accelerated al<strong>on</strong>g the first half of d and then uni<strong>for</strong>mly decelerated al<strong>on</strong>g thesec<strong>on</strong>d half with accelerati<strong>on</strong> a, see Fig. 2.7. For this purpose, the frequency difference∆ν is linearly swept from 0 to ∆ν max and then back to 0. Thus, during the transportdurati<strong>on</strong> t d , the velocity changes from 0 to v max = λ∆ν max /2 = at d /2 and back to 0.Experimental realizati<strong>on</strong>In the experiment, the frequencies of the two laser beams c<strong>on</strong>stituting our dipole trapare c<strong>on</strong>trolled by means of acousto-optic modulators (AOMs) installed in each beam as


32 Chapter 2: Submicrometer positi<strong>on</strong> c<strong>on</strong>trol of single atomsM1AOMvacuum chamberICCDdual-frequencygeneratorAPDzxyglasscellM2MOTMOT laser beamsAOMNd:YAGlaserFigure 2.8: Experimental setup of the optical c<strong>on</strong>veyor belt (schematic top view). Thestanding-wave dipole trap is displaced axially by frequency-shifting its laser beams bymeans of acousto-optic modulators (AOMs). Both modulators are set up in double-passc<strong>on</strong>figurati<strong>on</strong> (not shown here) and are driven by a phase-synchr<strong>on</strong>ous dual-frequencyRF generator. By synchr<strong>on</strong>ously tilting mirrors M1 and M2, the dipole trap axis istranslated radially.depicted in Fig. 2.8. The AOMs are set up in double-pass c<strong>on</strong>figurati<strong>on</strong> to avoid beamwalk-off during frequency shifts. The modulators operate at a center frequency of f 0 ≈100 MHz. Since the relative phase of the AOMs is directly translated to the spatial phase ofthe standing wave, the frequency sweep during the transport should be phase-c<strong>on</strong>tinuous.To drive the AOMs, we use a custom built dual-frequency synthesizer with two phasesynchr<strong>on</strong>izedRF outputs (APE Berlin, model DFD 100). For each transportati<strong>on</strong> thegenerator is programmed with a set of the sweep parameters such as the center frequency,f 0 , the maximum frequency detuning, ∆f max , and the sweep durati<strong>on</strong>, τ sweep = t d /2,using an RS232 interface. The reprogramming of the synthesizer takes about 0.5 s, andthe prepared frequency sweep is generated after externally triggering the DFD.Transportati<strong>on</strong> parametersDue to the double-pass c<strong>on</strong>figurati<strong>on</strong> of the AOMs, the frequency difference ∆f betweenthe two outputs of the DFD is translated to an optical detuning of the trapping beams of∆ν = 2 ∆f. The accelerati<strong>on</strong> during the ramped transport thus readsand the standing-wave structure is displaced by the distancea = ± 2λ ∆f maxt d(2.14)d = a t2 d4 . (2.15)


2.5 Active positi<strong>on</strong> c<strong>on</strong>trol 33The c<strong>on</strong>stant accelerati<strong>on</strong> of the standing-wave dipole trap results in a tilted standingwavepotential seen by trapped atoms [24], thus reducing the effective depth of the potentialwells. For a typical trap depth of U 0 = 1 mK, the maximum accelerati<strong>on</strong>, <strong>for</strong> whichthe atoms can be still trapped in the reduced potential, is limited to about 10 5 m/s 2 .There<strong>for</strong>e, <strong>for</strong> the positi<strong>on</strong> c<strong>on</strong>trol experiments in Sec. 2.2 we keep the accelerati<strong>on</strong> fixedat a = ±1000 m/s 2 and vary the sweep durati<strong>on</strong> and the maximum frequency detuningdepending <strong>on</strong> the required transportati<strong>on</strong> distance d:√ √a dd∆f max = and τ sweep =λa . (2.16)In the experiments <strong>on</strong> deterministic coupling of single atoms to the mode of a highfinessecavity, presented in Chapter 4, the arrival time of each atom into the cavity shouldbe known exactly, regardless of its initial positi<strong>on</strong>. There<strong>for</strong>e, to make the time synchr<strong>on</strong>izati<strong>on</strong>of different parts of these experiments easier, they are carried out with a fixedtransport durati<strong>on</strong> of t d = 4 ms. For the mean transportati<strong>on</strong> distance of 4.6 mm (distancebetween the MOT and the cavity mode, see Sec. 3.3.1) this results in the accelerati<strong>on</strong> ofa≈10 3 m/s 2 . The sweep parameters of the frequency synthesizer are then calculated as∆f max = 2 dt d λandτ sweep = t d2 . (2.17)Radial transport of atomsFor some experiments in Chapter 4, we need to precisely c<strong>on</strong>trol the positi<strong>on</strong> of the dipoletrap axis relative to the cavity mode. Technically, it is easier to displace the dipole traprather than to move the massive cavity holder. The displacement of the standing wave intwo radial directi<strong>on</strong>s, i.e. in the horiz<strong>on</strong>tal directi<strong>on</strong> perpendicular to the trap axis andin the vertical directi<strong>on</strong>, is realized by synchr<strong>on</strong>ously tilting the mirrors M1 and M2, seeFig. 2.8, mounted <strong>on</strong> piezo-electric actuators, around the corresp<strong>on</strong>ding axes. For tiltangles below 0.1 mrad, the variati<strong>on</strong> of the interference pattern at the locati<strong>on</strong> of theMOT is small and to a good approximati<strong>on</strong> a pure radial translati<strong>on</strong> is realized. Using thismethod, we can move atoms by a distance of about 40 µm, limited by the dynamic range ofthe actuators, in both radial directi<strong>on</strong>s within typically 50 ms, limited by their bandwidth.2.5 Active positi<strong>on</strong> c<strong>on</strong>trolUsing our scheme to precisely measure the positi<strong>on</strong> of an atom, we can now actively c<strong>on</strong>trolits absolute positi<strong>on</strong> al<strong>on</strong>g the DT axis. This is realized by transporting the atom toa predetermined positi<strong>on</strong> x target by means of our optical c<strong>on</strong>veyor belt. The following experimentdem<strong>on</strong>strates our positi<strong>on</strong> c<strong>on</strong>trol technique. Initially, we determine the positi<strong>on</strong>of the atom and its distance L from x target from an ICCD image, see Fig. 2.9(a). Then,


34 Chapter 2: Submicrometer positi<strong>on</strong> c<strong>on</strong>trol of single atomsTarget positi<strong>on</strong> x target(a)Counts / pixel0 2500Counts/100020100L(b)Counts / pixel0 2500Counts/100020100(c)Number of atoms12010080604020final distributi<strong>on</strong>initial distributi<strong>on</strong>00 10 20 30 40 50Atom positi<strong>on</strong> [ m]Figure 2.9: Absolute positi<strong>on</strong> c<strong>on</strong>trol of single trapped atoms. To transport an atom toa predetermined target positi<strong>on</strong> at x target = 9.5 µm, we first determine its initial positi<strong>on</strong>and calculate the distance L to the target positi<strong>on</strong> (a). Then we transport the atom andtake a sec<strong>on</strong>d camera picture (b). The histogram in (c) shows the accumulative data ofabout 400 experiments carried out with <strong>on</strong>e single atom at a time. Transfer of the atomsfrom the MOT to the DT yields the broad distributi<strong>on</strong> <strong>on</strong> the right (standard deviati<strong>on</strong>5.0 µm). The atoms are transported to the target positi<strong>on</strong> x target with a precisi<strong>on</strong> of300 nm (narrow distributi<strong>on</strong> <strong>on</strong> the left).


2.6 Calibrati<strong>on</strong> of the image scale 35the DFD is programmed with a set of sweep parameters corresp<strong>on</strong>ding to the desiredtransportati<strong>on</strong> distance L with a fixed accelerati<strong>on</strong> of a = 1000 m/s 2 . After programmingthe DFD, it is triggered to start moving the atom. To c<strong>on</strong>firm the successful transport tox target , we take a sec<strong>on</strong>d image of the atom and measure its final positi<strong>on</strong>, see Fig. 2.9(b).The experiment is repeated in a completely automated way many times. Finally we postselectand analyze about 400 experimental shots with a single atom each. 5Because the atoms are randomly loaded from the MOT into the DT, the distributi<strong>on</strong>of their initial positi<strong>on</strong>s, see Fig. 2.9(c), has a standard deviati<strong>on</strong> of 5.0(±0.3) µm, corresp<strong>on</strong>dingto the MOT radius. After the transport, the width of the distributi<strong>on</strong> of the finalpositi<strong>on</strong>s is drastically reduced to σ c<strong>on</strong>trol = 300(±15) nm. This width is limited by theerrors in determining the final and initial positi<strong>on</strong> of the atom, by the DT positi<strong>on</strong> drifts,σ drift , during the typical time of 1.5 s between the two successive exposure intervals, and bythe transportati<strong>on</strong> error, σ transp , resulting from a sort of the discretizati<strong>on</strong> error (round-offerror) of the DFD. From the above DT phase measurement we find σ drift = 140(±20) nm.Assuming that√σ c<strong>on</strong>trol = 2∆x 2 stat + σ2 drift + σ2 transp , (2.18)we calculate σ transp = 190(±25) nm, comparable to the statistical error.In additi<strong>on</strong> to statistical errors, the accuracy of the positi<strong>on</strong> c<strong>on</strong>trol is subject tosystematic errors. The predominant systematic error of this measurement stemmed fromthe calibrati<strong>on</strong> of our length scale. In the present case, a relative calibrati<strong>on</strong> error of0.4 % resulted in a 120 nm shift of the final positi<strong>on</strong>s with respect to the target positi<strong>on</strong>after a transport over L ≈ 30 µm. However, this error can be reduced by improving theaccuracy of the calibrati<strong>on</strong>, see next secti<strong>on</strong>.2.6 Calibrati<strong>on</strong> of the image scaleThe spatial in<strong>for</strong>mati<strong>on</strong> <strong>on</strong> trapped atoms is obtained from CCD images and thus isoriginally expressed in length units of CCD pixels. All experiments presented in thecurrent chapter are per<strong>for</strong>med <strong>on</strong> atoms located in the object plane of the imaging optics.Their analysis would be impossible without precise calibrati<strong>on</strong> of the image scale, that isthe precise corresp<strong>on</strong>dence between the measured dimensi<strong>on</strong> ˜d[pixel] <strong>on</strong> a CCD image andthe real dimensi<strong>on</strong> d[µm] at the locati<strong>on</strong> of the atoms:[ ] µmd[µm] = α ˜d[pixel]. (2.19)pixelThe calibrati<strong>on</strong> parameter α depends <strong>on</strong> the magnificati<strong>on</strong> of the imaging optics, M,and the linear size of a CCD pixel, a pixel , as α = a pixel /M. The pixel size of our ICCDcamera is a pixel = 13 µm and the designed magnificati<strong>on</strong> of the system is M = 14(±1),resulting inα magnif = 0.93(±0.07) µm/pixel. (2.20)5 These experiments have been per<strong>for</strong>med be<strong>for</strong>e the implementati<strong>on</strong> of the c<strong>on</strong>trolled loading of theDT with <strong>on</strong>ly <strong>on</strong>e atom, described in Sec. 1.4.


36 Chapter 2: Submicrometer positi<strong>on</strong> c<strong>on</strong>trol of single atomsThe obtained precisi<strong>on</strong> is not sufficient <strong>for</strong> our data analysis, since <strong>for</strong> a length of100 µm we already get a 7 µm uncertainty. There<strong>for</strong>e, we have developed several othercalibrati<strong>on</strong> methods offering better precisi<strong>on</strong>.Using a c<strong>on</strong>stant transport of atomsThis calibrati<strong>on</strong> method is based <strong>on</strong> a transport of a single atom over a known distanced[µm], while measuring its initial and final positi<strong>on</strong>s from the CCD image and determiningthe corresp<strong>on</strong>ding transportati<strong>on</strong> distance ˜d[pixel]. After transporting 45 single atoms wehave determined a calibrati<strong>on</strong> parameter ofα trans = 0.931(±0.004) µm/pixel. (2.21)The main sources of error are the same as those of the positi<strong>on</strong> c<strong>on</strong>trol experimentin Sec. 2.5, namely a statistical error of two positi<strong>on</strong> measurements and the fluctuati<strong>on</strong>sof the dipole trap between two images. Besides, this calibrati<strong>on</strong> method also relies <strong>on</strong> afaultless operati<strong>on</strong> of the dual-frequency synthesizer driving our c<strong>on</strong>veyor belt. Its internaldata digitalizati<strong>on</strong> can in particular result in a systematical error, which might in generalbe dependent <strong>on</strong> the transportati<strong>on</strong> distance. To eliminate this problem, the calibrati<strong>on</strong>should be per<strong>for</strong>med <strong>for</strong> different transportati<strong>on</strong> distances, making this method morecomplicated.Using the periodicity of the DT potentialAnalyzing the histogram of distancesTo per<strong>for</strong>m a precise scale calibrati<strong>on</strong>, we analyze the measured atomic separati<strong>on</strong>s.Due to the standing-wave structure of our dipole trap, the real atomic separati<strong>on</strong>s exactlyequal integer multiples of λ/2. In Sec. 2.3 we have c<strong>on</strong>structed an auxiliary data set {d ′ i }using equati<strong>on</strong> (2.10) by calculating modulo λ/2 of atomic separati<strong>on</strong>s {d i }. The propertiesof the new set str<strong>on</strong>gly depend <strong>on</strong> the calibrati<strong>on</strong> α, since it was used <strong>for</strong> calculating theparent data set {d i } in units of [µm] from atomic separati<strong>on</strong>s initially measured in pixels.As can be seen from the histogram in Fig. 2.4, the optimal calibrati<strong>on</strong> parameter α 0results in the distributi<strong>on</strong> {d ′ i } centered at zero and having the smallest width. There<strong>for</strong>e,we calculate a mean and a standard deviati<strong>on</strong> of {d ′ i } <strong>for</strong> different α. The corresp<strong>on</strong>dingdependencies, obtained from about 3000 atomic separati<strong>on</strong>s, 6 are presented in Fig. 2.10.The optimal calibrati<strong>on</strong> α 0 is then found at a positi<strong>on</strong> of the mean {d ′ i} = 0 and of thestandard deviati<strong>on</strong> σ ({d ′ i }) = min. The obtained results are listed in a table below.Measureα 0 [µm/pixel]mean value 0.9331 ± 0.0002standard deviati<strong>on</strong> 0.9325 ± 0.00026 This data set is the same <strong>on</strong>e as used in Sec. 2.3 <strong>for</strong> analyzing the periodicity of the dipole trap.


2.6 Calibrati<strong>on</strong> of the image scale 370.040.16Mean value [ ]0.020.00-0.02-0.040.150.140.130.120.110.91 0.92 0.93 0.94 0.95 0.96Calibrati<strong>on</strong> parameter [ m/pixel]Standard deviati<strong>on</strong> [ ]Figure 2.10: Calibrati<strong>on</strong> of the image scale using atomic separati<strong>on</strong>s calculated moduloλ/2. Shown are the mean value and the standard deviati<strong>on</strong> of the atomic separati<strong>on</strong>s {d ′ i }modulo λ/2 (see equati<strong>on</strong> (2.10)) <strong>for</strong> different calibrati<strong>on</strong> parameters α. The optimal α 0corresp<strong>on</strong>ds to zero of the mean value and a minimum of the standard deviati<strong>on</strong>.The statistical error, which is mainly due to the finite size N of the set {d i }, is estimatedby calculating values of α 0 from 20 randomly selected subsets of N/2 distances. If theirstandard deviati<strong>on</strong> is denoted by σ N/2 , the statistical error <strong>for</strong> the whole set of distancesis then given by σ N = σ N/2 / √ 2.We set the final value of α by averaging over two results obtained from different criteria:α modulo = 0.9328(±0.0002) µm/pixel . (2.22)Fourier trans<strong>for</strong>mati<strong>on</strong> of the atomic separati<strong>on</strong>sAn alternative approach to investigate the periodicity in the data set {d i } is based <strong>on</strong>a Fourier analysis of the measured atomic separati<strong>on</strong>s [76]. For this purpose, we c<strong>on</strong>siderthe periodic distributi<strong>on</strong> functi<strong>on</strong>f(x) = 1 NN∑δ(d i − x), (2.23)i=1built by summing over delta functi<strong>on</strong>s at the measured atomic separati<strong>on</strong>s d i [pixel]. Tofind its period, corresp<strong>on</strong>ding to the λ/2 periodicity of the atomic separati<strong>on</strong>s, we calculate


38 Chapter 2: Submicrometer positi<strong>on</strong> c<strong>on</strong>trol of single atoms0.40.2Re(g)0.0-0.2-0.40.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.00.40.2Im(g)0.0-0.2-0.40.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0Spatial frequency k [1/pixel]Figure 2.11: Fourier trans<strong>for</strong>mati<strong>on</strong> of the distributi<strong>on</strong> of atomic separati<strong>on</strong>s.its Fourier trans<strong>for</strong>mg(k) = 1 √2π∫ ∞−∞f(x) e 2πikx dx =1√2π NN ∑i=1e 2πikd i. (2.24)The real, Re(g), and imaginary, Im(g), parts of the Fourier trans<strong>for</strong>m of f(x) areshown in Fig. 2.11. The analyzed atomic separati<strong>on</strong>s are the same as used in the previouscalibrati<strong>on</strong> method. The spatial frequency k 0 , which corresp<strong>on</strong>ds to the periodicity ofthe DT, shows up as a peak in Re(g) and as a zero crossing in Im(g). The calibrati<strong>on</strong>parameter α is then found asα = k 0 λ/2. (2.25)The results of this calibrati<strong>on</strong> method are listed in a table below.Measureα [µm/pixel]real part 0.9326 ± 0.0002imaginary part 0.9331 ± 0.0002The statistical error of α is estimated in the same way as in the previous calibrati<strong>on</strong>method using the same subsets of N/2 distances. After averaging over the two α ′ s, thefinal calibrati<strong>on</strong> parameter readsα Fourier = 0.9329(±0.0002) µm/pixel. (2.26)


2.7 C<strong>on</strong>clusi<strong>on</strong> and discussi<strong>on</strong> 39The two calibrati<strong>on</strong> methods using the histogram analysis and the Fourier trans<strong>for</strong>mare mathematically closely related. Their similarity is indicated by the fact that thecalibrati<strong>on</strong> error as well as the discrepancy between the two measures used <strong>for</strong> determiningα are the same <strong>for</strong> both methods.Summarizing, we have developed several different methods <strong>for</strong> calibrating the CCDimage scale. While providing different precisi<strong>on</strong>, they all result in the same calibrati<strong>on</strong>parameter α within their individual uncertainties. The method based <strong>on</strong> transportingsingle atoms is fast, but it appears to be an order of magnitude less precise than twoother procedures utilizing the discreteness of the atomic separati<strong>on</strong>s. They both result inthe same high precisi<strong>on</strong> <strong>for</strong> determining α of about 0.02, %. However, the more involvedanalysis of atomic separati<strong>on</strong>s require a larger data set as compared to the transportmethod.Note that <strong>for</strong> technical reas<strong>on</strong>s the magnificati<strong>on</strong> of our imaging system was approximatelydoubled be<strong>for</strong>e per<strong>for</strong>ming the set of experiments described in Chapter 4. Thisyields the new calibrati<strong>on</strong> of the CCD image of α Chapter 4 = 0.4967(±0.0002) µm/pixelfound by per<strong>for</strong>ming Fourier analysis of atomic separati<strong>on</strong>s and used in the rest of mythesis. This modificati<strong>on</strong> by no means changes the main results of the present chapter.2.7 C<strong>on</strong>clusi<strong>on</strong> and discussi<strong>on</strong>In this chapter I have presented a detecti<strong>on</strong> scheme <strong>for</strong> the absolute and relative positi<strong>on</strong> ofindividual atoms stored in our standing-wave dipole trap, yielding nanometric resoluti<strong>on</strong>.This scheme allows us to resolve the 532 nm-period standing-wave structure of our dipoletrap and to measure the exact number of potential wells separating simultaneously trappedatoms. We have furthermore used our positi<strong>on</strong> detecti<strong>on</strong> scheme to transport an atom toa predetermined positi<strong>on</strong> with a sub-optical wavelength accuracy of 300 nm.These results represent an important step towards experiments in which the relativeor absolute positi<strong>on</strong> of single atoms has to be c<strong>on</strong>trolled to a high degree. In Chapter 4,this technique is used to transport individual atoms over several millimeter distances andto deterministically couple them to the mode of a high-Q optical res<strong>on</strong>ator.Furthermore, knowing the exact number of potential wells separating the atoms, we cannow c<strong>on</strong>trol this parameter by placing atoms into specific potential wells of our standingwave using additi<strong>on</strong>al optical tweezers [26]. Finally, the dem<strong>on</strong>strated high degree ofc<strong>on</strong>trol allows us to transport two individual atoms into <strong>on</strong>e potential well in order toinduce cold collisi<strong>on</strong>s between them [77].


40 Chapter 2: Submicrometer positi<strong>on</strong> c<strong>on</strong>trol of single atoms


Chapter 3The high-finesse optical res<strong>on</strong>atorThe heart of any cavity <strong>QED</strong> experiment is a res<strong>on</strong>ator <strong>for</strong> c<strong>on</strong>fining and storing light. Inthis chapter, I first describe the high-reflectivity mirrors, which were specially designedand manufactured to meet the requirements <strong>for</strong> our planned cavity <strong>QED</strong> experimentswith optically trapped cesium atoms. Several techniques <strong>for</strong> their characterizati<strong>on</strong> andanalysis as well as the corresp<strong>on</strong>ding results are presented. The sec<strong>on</strong>d part of thechapter is dedicated to assembling, locking, and studying the high-finesse optical cavity,the particular properties of which render it compatible with the current atom-trappingexperimental setup. This cavity is used in the experiments <strong>on</strong> coupling single atoms to acavity mode as described in Chapter 4.3.1 An optical res<strong>on</strong>ator <strong>for</strong> cavity <strong>QED</strong> experimentsStr<strong>on</strong>g coupling in cavity <strong>QED</strong>For exploring coherent atom-phot<strong>on</strong> interacti<strong>on</strong> the coupling between atom and cavityfield should be str<strong>on</strong>g compared to all losses (dissipati<strong>on</strong>s) in the system, see Fig. 3.1. Thecoupling rate, g, <strong>for</strong> a given atomic transiti<strong>on</strong> with the dipole moment, d, is proporti<strong>on</strong>alto the electric field strength, E, <strong>for</strong> a single phot<strong>on</strong>, resulting in [78]g = d E = d √ω02ɛ 0 V , (3.1)see also Sec. 4.2.1. The atom-phot<strong>on</strong> coupling can be there<strong>for</strong>e increased by decreasingthe cavity mode volume, V , providing better c<strong>on</strong>finement <strong>for</strong> the light field.The main sources of energy dissipati<strong>on</strong> in an atom-cavity system are the atomic dipoledecay rate, γ, and the cavity field decay rate, κ. The first occurs due to sp<strong>on</strong>taneousemissi<strong>on</strong> of an atom into modes of the electromagnetic field other than the cavity mode.The other source of dissipati<strong>on</strong> is based <strong>on</strong> leakage and absorpti<strong>on</strong> of cavity phot<strong>on</strong>s dueto n<strong>on</strong>-perfect, lossy mirrors. This cavity loss rate can be kept small by utilizing mirrorswith the highest possible reflectivity.41


42 Chapter 3: The high-finesse optical res<strong>on</strong>atorgFigure 3.1: Basic elements and rates in a cavity <strong>QED</strong> system. An atom is coupled toan electromagnetic mode of a cavity, built of i.e. two mirrors. In the str<strong>on</strong>g couplingregime, the atom-phot<strong>on</strong> coupling rate, g, is larger than the atomic dipole decay rate, γ,and the cavity field decay rate, κ.If the c<strong>on</strong>diti<strong>on</strong> g ≫ (γ, κ) is satisfied, we get into the so-called str<strong>on</strong>g coupling regime,which allows us to study coherent energy evoluti<strong>on</strong> in the coupled atom-cavity systemand to explore cavity <strong>QED</strong> effects while neglecting all sources of energy dissipati<strong>on</strong>. Asa measure of a coupling strength compared to dissipati<strong>on</strong> we can use the single-atomcooperativity parameter, defined asIn the str<strong>on</strong>g coupling regime, C 1 ≫ 1.Mode volumeC 1 = g22κγ . (3.2)I c<strong>on</strong>sider in the following a symmetric Fabry-Perot res<strong>on</strong>ator built of two opposing c<strong>on</strong>cavemirrors, as schematically depicted in Fig. 3.1. The electromagnetic mode inside such acavity has a Gaussian standing-wave profile described by a spatial functi<strong>on</strong>ψ(⃗r) = exp[− x2 + y 2 ] ( ) 2πzsin , (3.3)λw 2 0where w 0 is the mode waist. Here, the divergence of the Gaussian mode <strong>for</strong> the shortcavity (i.e. if the cavity length is smaller than the Rayleigh range of the Gaussian mode)has been neglected. The mode volume V of the electromagnetic field is obtained byspatially integrating |ψ(⃗r)| 2 , yielding <strong>for</strong> the fundamental TEM 00 modeV = πw2 04 L , (3.4)where L denotes the cavity length. If the cavity is made of equal c<strong>on</strong>cave mirrors with aradius of curvature R, its waist equals [79]√√λ L(w 0 = R − L ). (3.5)π 2 2


3.1 An optical res<strong>on</strong>ator <strong>for</strong> cavity <strong>QED</strong> experiments 43Combining equati<strong>on</strong>s (3.4) and (3.5), we getV = λ 8√L 3( 2R − L ) . (3.6)For R ≫ L, which is justified <strong>for</strong> our short cavities (see Sec. 3.3), equati<strong>on</strong> (3.6) leads toV ∝ R 1/2 L 3/2 . (3.7)Thus, the mode volume can be reduced either by shortening a cavity length or by usingmirrors with larger surface curvature 1/R.Radius of curvatureTaking into account equati<strong>on</strong>s (3.1) and (3.7), the atom-cavity coupling scales as g ∝R −1/4 . Since κ is inverse proporti<strong>on</strong>al to the cavity finesse F, see equati<strong>on</strong> (3.12), thedependence of the cooperativity parameter <strong>on</strong> F and the mirror radius of curvature readsC 1 ∝ F/ √ R. Thus, the natural way to increase C 1 is to use mirrors with smaller radiusof curvature.However, technical difficulties in polishing mirror substrates with small radii prevent<strong>on</strong>e to achieve small scattering losses and thus a high mirror reflectivity. For instance, theF = 10 5 finesse of our first high-reflectivity mirrors [80, 81] having a radius of curvature ofR = 1 cm was limited mainly by their surface quality. Thus, after discussi<strong>on</strong>s with mirrormanufacturers <strong>on</strong> possible trade-offs between the small R and the large F, we decided<strong>on</strong> using mirrors with R = 5 cm, expecting to get better surface quality of the glasssubstrates and finally to achieve a finesse of about <strong>on</strong>e milli<strong>on</strong>. In this case, by changingfrom a cavity with R = 1 cm and F = 10 5 to a cavity with R = 5 cm and F = 10 6 , weexpect an increase of the cooperativity parameter by more than a factor of 4.<strong>Cavity</strong> length and mirror shapeAnother way to increase the coupling g is to reduce the cavity length. However, the planned<strong>QED</strong> experiments require the laser beams <strong>for</strong>ming the dipole trap to pass between thecavity mirrors and cross the cavity mode. This sets the lower limit <strong>on</strong> the cavity clearance(the distance between mirror edges) of about 150 µm <strong>for</strong> a waist of the trapping beamsof w 0, DT = 35–40 µm, see Sec. 1.3. In this case the dipole trap beams are clipped by themirrors at 2w 0, DT .Since the mirrors are spherical, there is an indentati<strong>on</strong> from the mirror edges to theircenter. Thus, the gap between edges, as seen by the dipole trap beams, is less than the realcavity length, L, occupied by the cavity mode. The reducti<strong>on</strong> of the cavity clearance is∆L = 2R − √ 4R 2 − D 2 (3.8)with R and D denoting the radius of curvature and the mirror diameter, respectively. Forinstance, <strong>for</strong> R = 5 cm and D = 3 mm, ∆L = 45 µm, thus significantly reducing thespacing between the mirror edges <strong>for</strong> a desired cavity length of L = 150 µm.


44 Chapter 3: The high-finesse optical res<strong>on</strong>ator4 mm1 mm 3 mmback flat surfaceAR coatedfr<strong>on</strong>t c<strong>on</strong>cave surfacesuper HR coatedradius of curvatureR = 5 cm 1 mmFigure 3.2: High-reflectivity mirror <strong>for</strong> cavity <strong>QED</strong> (side view).The clearance of the cavity can be increased by using smaller mirrors. However, smallsubstrates are difficult to handle. A soluti<strong>on</strong> is to turn relatively large substrates down tothe desired diameter at the mirror face. For a c<strong>on</strong>ed mirror with D = 1 mm, ∆L = 5 µm,much smaller than L. Besides, another advantage of a c<strong>on</strong>ed substrate shape is thepossibility to place an assembled cavity closer to the MOT without blocking the MOTbeams, as will be discussed in Sec. 3.3.1.Note that if a mirror substrate is not symmetrically curved relative to its rotati<strong>on</strong>axis, ∆L will be further increased. For the mirrors investigated in Sec. 3.2, the maximalobserved offset of the center of the mirror curvature with respect to the substrate axis is100 µm, resulting in additi<strong>on</strong>al decrease of the cavity clearance of about 1 µm.Taking into account all presented arguments and drawing <strong>on</strong> experience of otherresearch groups, we decided <strong>on</strong> the <strong>for</strong>m of mirror substrates presented in Fig. 3.2. Here,the 3 mm substrates are turned down to a D = 1 mm diameter at the tip. The radiusof curvature of the c<strong>on</strong>cave surface is 5 cm. For a cavity length of, e.g. 150 µm, the sizeof the TEM 00 mode at the mirror surface is w ≈ 23 µm ≪ D, resulting in negligiblediffracti<strong>on</strong> losses due to the finite mirror aperture seen by the light. 1 For the cavityfinesse of <strong>on</strong>e milli<strong>on</strong>, the expected cooperativity parameter is C expect1 = 134 ≫ 1.3.2 High-reflectivity mirrors3.2.1 Mirror designA high-finesse cavity exhibits low phot<strong>on</strong> losses, i.e. the mirror transmissi<strong>on</strong>, absorpti<strong>on</strong>,and scattering are kept at very low level. There<strong>for</strong>e, the two most important issues tobe c<strong>on</strong>cerned while manufacturing low-loss mirrors are a small roughness of the mirrorsurface and a special dielectric coating of ultra-high reflectivity.1 Diffracti<strong>on</strong> losses reach 1 ppm <strong>for</strong> a mirror radius of 2.63 w, where w is the beam radius at the mirrorsurface.


3.2 High-reflectivity mirrors 45Currently, the <strong>on</strong>ly company able to fabricate, coat, and c<strong>on</strong>e high-reflectivity opticalmirrors suitable <strong>for</strong> cavity <strong>QED</strong> experiments is the Research Electro-Optics Inc. (REO)in Boulder, Colorado. The mirror manufacturing process c<strong>on</strong>sists of the following steps:First, the substrates of BK7 glass are prepared in a cylindrical <strong>for</strong>m with a diameterof 7.75 mm and a thickness of 4 mm. One side of the substrates is flat and the otherside is superpolished to a c<strong>on</strong>cave surface with a radius of curvature of 5 cm. Then,the substrates are coated. A standard anti-reflecti<strong>on</strong> coating is applied to their flatback surface with measured reflectivity below 0.03 % at 852 nm and below 0.06 % at836 nm (wavelengths of the probe and the lock laser, respectively, see Sec. 3.3.2). Thehigh-reflecting coating of the curved face should provide a reflectivity of 99.99997 %(corresp<strong>on</strong>ding to a finesse of 10 6 ) and thus is more complex. It c<strong>on</strong>sists of 45 alternatingdielectric layers of Ta 2 O 5 (n high = 2.041) and SiO 2 (n low = 1.455), and each layer has anoptical thickness of λ/4 at 875 nm. 2 Note that the coating is not centered at 852 nm(probe laser) in order to get a higher transmissi<strong>on</strong> at 836 nm (lock laser) making thecavity stabilizati<strong>on</strong> easier, see Sec. 3.3.2. The total physical thickness of the coating isabout 5640 nm. In the last producti<strong>on</strong> stage, the coated substrates are turned down to adiameter of 3 mm, and then c<strong>on</strong>ed to 1 mm at the mirror tip.3.2.2 Portable cavity ring-down setupCollaborati<strong>on</strong> with REOHigh-finesse c<strong>on</strong>ed c<strong>on</strong>cave mirrors, used in cavity <strong>QED</strong> experiments, c<strong>on</strong>stitute the currenttechnological state of the art that can be reached using special super-polishing andsuper-coating techniques. Since mirrors with R = 5 cm and F ≈ 10 6 were never producedand investigated be<strong>for</strong>e, REO could not guarantee a finesse of <strong>on</strong>e milli<strong>on</strong> <strong>for</strong> the c<strong>on</strong>cavemirrors with the specified radius of curvature. For getting the highest achievable finesse,the mirror substrates should be tested after each intermediate manufacturing process. Thiswould help REO to identify sources of errors and imperfecti<strong>on</strong>s and to find out possibleways <strong>for</strong> optimizati<strong>on</strong> and improvement of different producti<strong>on</strong> steps. Besides, the finalfinesse measurement could be used to select the best mirrors with presumably F 10 6 .A straight-<strong>for</strong>ward way to determine the mirror reflectivity is to measure the fracti<strong>on</strong>of light reflected off the mirror. However, achieving a high precisi<strong>on</strong> using this methodbecomes a n<strong>on</strong>-trivial task due to a reduced signal-to-noise ratio if going to smaller mirrorlosses. Alternatively, the task of measuring the mirror reflectivity is equivalent to measuringthe phot<strong>on</strong> lifetime in a cavity composed of the mirrors under study. By using thecavity ring-down method (CRD, see below), the direct measurement of the cavity decaytime <strong>for</strong> a cavity of a known length allows us to deduce the cavity losses and, there<strong>for</strong>e,its finesse.We reached an agreement with REO, where I had to assist them in testing the mirrorsubstrates by measuring the mirror reflectivity after each manufacturing step. This impliedthe verificati<strong>on</strong> of the actual finesse level achieved utilizing our CRD test setup furnished2 For a model of multi-layer dielectric coatings see [82, 83].


46 Chapter 3: The high-finesse optical res<strong>on</strong>atorin REO’s metrology lab. The CRD measurements at the specified wavelength had to beper<strong>for</strong>med <strong>on</strong> all coated mirrors prior to the c<strong>on</strong>ing process as well as <strong>on</strong> the final reshapedmirrors. Moreover, up<strong>on</strong> delivery of the mirrors and the CRD setup from REO back toB<strong>on</strong>n, I had to retest the mirrors and verify that the per<strong>for</strong>mance had not changed dueto transport-induced damage. On its part, REO was resp<strong>on</strong>sible to fully characterize andanalyze the mirror substrates prior to coating as well as to use the results of my finessemeasurements <strong>for</strong> optimizing the coating, c<strong>on</strong>ing, and cleaning processes in order to getthe best possible finesse. Besides, the manufacturer warranted that his packaging wouldprotect the substrates and coatings such that the finesse would not be measurably reducedas a result of the shipment to Germany.<strong>Cavity</strong> ring-downThe cavity finesse is defined as the ratio of the free spectral range (FSR) to the cavity fulllinewidth at half maximum (FWHM),F =The phot<strong>on</strong> lifetime is c<strong>on</strong>nected to the cavity linewidth asτ =ω FSRω FWHM. (3.9)1ω FWHM. (3.10)Combining (3.9) and (3.10) and using the definiti<strong>on</strong> of ω FSR , we get the basic <strong>for</strong>mula ofthe CRD measurement:F = cπ L τ (3.11)with the cavity length L. Since κ = ω FWHM /2, the cavity field decay rate can then befound asκ = 1 or κ = cπ2τ2LF . (3.12)The phot<strong>on</strong> lifetime can be directly inferred from the energy decay of the intracavitylight field <strong>for</strong> a cavity initially filled with res<strong>on</strong>ant light. The exp<strong>on</strong>ential decay is measuredby observing the light lacking through the cavity mirror and is known as cavity ring-down.The first experiment <strong>on</strong> measuring decay time to determine cavity losses is presentedin Ref. [84]. Measurement of ultra-low losses <strong>on</strong> mirrors, similar to those used in ourexperiment, was first per<strong>for</strong>med in the group of Prof. Kimble [85].A typical experimental procedure of the CRD measurement is the following: By scanningthe cavity length, the cavity res<strong>on</strong>ance frequency is tuned towards the laser frequency,until the cavity transmissi<strong>on</strong> starts to increase. If the transmitted intensity reaches a presetthreshold level, an auxiliary electr<strong>on</strong>ics triggers an acousto-optic switch, which turnsoff the incident laser. The subsequent decay of the cavity output is detected by a fastphotodetector and recorded <strong>on</strong> a digital storage oscilloscope.


3.2 High-reflectivity mirrors 47Portable cavity ring-down setupThe portable and self-c<strong>on</strong>tained cavity ring-down setup enabling us to measure the transmissi<strong>on</strong>of both c<strong>on</strong>ed and unc<strong>on</strong>ed HR coated substrates at REO is based <strong>on</strong> that developedin our group by W. Rosenfeld [81] <strong>for</strong> testing the first REO mirrors (R = 10 cm,F ≤ 10 5 ), but extended to c<strong>on</strong>tain a diode laser and an AOM <strong>for</strong> fast switching a probelaser. In additi<strong>on</strong>, the modified setup is compatible with testing large unc<strong>on</strong>ed mirrorsubstrates and is supplemented with required electr<strong>on</strong>ics and power supplies.A photo of the breadboard (size of 30 cm×60 cm), c<strong>on</strong>taining the optical setup, as wellas the corresp<strong>on</strong>ding schematic drawing is presented in Fig. 3.3. The key element of thesetup is a mirror holder, developed and described in detail in Ref. [81]. It allows us to scanthe cavity length without gluing or clamping the c<strong>on</strong>stituent mirrors. In the holder, themirrors are placed face-to-face in a V-groove providing good coaxial alignment. In order toscan their distance, the two parts of the holder, each carrying <strong>on</strong>e mirror, are c<strong>on</strong>nected bya piezo-electric transducer. By applying a voltage to it, the two holder halves are pushedapart, thus changing the cavity length. Besides small c<strong>on</strong>ed mirrors, the developed holderis also suitable <strong>for</strong> testing large mirror substrates of 7.5 mm in diameter be<strong>for</strong>e turningthem down to 3 mm. This was necessary <strong>for</strong> comparing the mirror reflectivity be<strong>for</strong>e andafter the c<strong>on</strong>ing process.As a laser source <strong>for</strong> probing the cavity we use a home-made temperature-stabilizeddiode laser in Littrow c<strong>on</strong>figurati<strong>on</strong> emitting at 852 nm. The short term frequencystability (at 1 ms) is expected to be 100 kHz. An optical isolator (OI) prevents opticalfeed-back to the laser diode and, thus, insures its stable operati<strong>on</strong>. An acousto-opticmodulator (AOM), operated in first order, is used <strong>for</strong> fast switching the incident laseroff and <strong>on</strong>. By means of both a half-wave (λ/2) and a quarter-wave (λ/4) plate wecan choose between different polarizati<strong>on</strong> modes of the (in general) birefringent cavity,see Sec. 3.2.4. The aperture (A), placed after the cavity holder, blocks stray light notcoupled to the cavity mode. Next, the cavity output light is split <strong>on</strong> a beam splitter (BS)and send to two optical detectors: a CCD camera is used to m<strong>on</strong>itor and to distinguishbetween different transversal modes of the cavity and, finally, an avalanche photodiode(APD) measures the cavity output light.3.2.3 Reflectivity measurementTo determine the mirror quality, we per<strong>for</strong>m a CRD measurement pairwise <strong>on</strong> all producedmirrors using our CRD setup. For this purpose, two mirrors are put into the mirror holderwith a separati<strong>on</strong> of 10 mm (estimated uncertainty is 5 %). The mode coupling of theincoming laser beam is optimized <strong>for</strong> the cavity TEM 00 mode. For better signal-to-noiseratio, the cavity decay signal is averaged over about 20 single traces. A typical measurementresult is shown in Fig. 3.4. Here, the measured decay time of 12.4 µs corresp<strong>on</strong>dsto a finesse of F = 1.17 × 10 6 and thus to cavity losses of (1 − R) = π/F = 2.7 ppm.The estimated error is about 5 % resulting from the uncertainty of the cavity length. Theswitching time of the incident laser as well as the rise- and fall times of the APD have


48 Chapter 3: The high-finesse optical res<strong>on</strong>atorCCDlaserAPD852 nmBSAOMOImirrorholderl/2Al/4voltagescanstoragescopetriggerAOMdriver(a) Scheme of the setup.(b) Photo of the setup.Figure 3.3: Portable cavity ring-down setup. Laser: diode laser at 852 nm, OI: opticalisolator, AOM: acousto-optic modulator, λ/2 and λ/4: half- and quarter-wave plate,respectively, A: aperture, BS: beam-splitter, APD: avalanche photodiode, CCD: CCDcamera. For better guidance, the arrangement of optical elements in (a) and (b) isthe same. The white line in (b) follows the laser beam path. The size of the opticalbreadboard is 30 × 60 cm 2 .


3.2 High-reflectivity mirrors 490.7<strong>Cavity</strong> output intensity [V]0.60.50.40.30.20.10.00 10 20 30 40 50Time [µs]Figure 3.4: <strong>Cavity</strong> ring-down signal. At time 0 µs the cavity transmissi<strong>on</strong> exceeds atrigger threshold and after a delay of about a 5 µs the incident laser is turned off. Afterthat the intracavity light field starts to leak out. The thick curve is an average overabout 20 single traces detected by the APD. The thin line is an exp<strong>on</strong>ential fit to thedata (starting from 5 µs), yielding a decay time of τ = 12.4 µs and, thus, a finesse ofF = 1.17 × 10 6 <strong>for</strong> a 1 cm-l<strong>on</strong>g cavity.been measured independently and are found to be in the sub-microsec<strong>on</strong>d range, thus noteffecting the measured phot<strong>on</strong> lifetime.The measured cavity decay time al<strong>on</strong>e does not provide in<strong>for</strong>mati<strong>on</strong> about individualmirrors, but <strong>on</strong>ly about the cavity as a whole. Moreover, the measured losses characterizenot the entire mirror surface, but <strong>on</strong>ly the small spot probed by the cavity mode. Incase of inhomogeneous cavity losses (due to, e.g., varying roughness of the surface orlocal damages of the coating) the measured finesse may str<strong>on</strong>gly vary from point to point.Thus, the finesse of a finally assembled cavity may be in general different from the <strong>on</strong>emeasured be<strong>for</strong>e.In order to eliminate these difficulties, we have tested the mirrors after differentcoating, c<strong>on</strong>ing, or cleaning runs in the following way. For each mirror pair we per<strong>for</strong>medseveral CRD measurements with different relative orientati<strong>on</strong>s of the mirrors in order toinvestigate the spatial inhomogeneity of the cavity losses. Each mirror was investigatedseveral times in different pairs enabling us to determine its individual properties.As a result of the successful cooperati<strong>on</strong> with REO, which included CRD measurements<strong>on</strong> all mirrors after each intermediate step in their manufacturing, we got a set of about 10high-reflectivity mirrors, any two of each can c<strong>on</strong>stitute a cavity with a finesse of about10 6 . Further investigati<strong>on</strong> of the mirror properties was c<strong>on</strong>tinued in our laboratory afterthe selected mirrors had been delivered to B<strong>on</strong>n.


50 Chapter 3: The high-finesse optical res<strong>on</strong>ator3.2.4 Mirror birefringenceOrigin of the birefringenceIn general, all optical elements exhibit some degree of birefringence, even if their materialsdo not show inherent birefringent properties. This effect may come from slight deviati<strong>on</strong>of the optical system from a perfect circularly symmetric <strong>on</strong>e, <strong>for</strong> instance, because of itsgeometrical <strong>for</strong>m or stress tensi<strong>on</strong> in a bulk material.The high-reflecting coating might also possess some birefringence depending <strong>on</strong> thecoating process, e.g., <strong>on</strong> the orientati<strong>on</strong> of a mirror substrate in a coating chamber whichmight result in a slight asymmetry of the coating. Usually, the birefringent phase shiftin this case is negligibly small and cannot be directly observed or measured. But in ahigh-finesse cavity the light field probes the coating many times resulting in a dramaticenhancement of any small phase shift, making the cavity birefringence significant. If amirror is glued <strong>on</strong>to a holder, the induced mechanical stress in its substrate transmits tothe coated surface resulting in an additi<strong>on</strong>al stress-induced birefringence [82], which canbe much larger than the inherent birefringence of the coating. This effect, essential <strong>for</strong>assembled cavities, will be investigated in Sec. 3.3.4.The cavity birefringence can, in principle, also be caused by the mirror substrates.However, the res<strong>on</strong>ant intra-cavity light field penetrates <strong>on</strong>ly through a few top-mostcoating layers and thus cannot effectively probe the substrate material. The light res<strong>on</strong>antlytransmitted by the cavity passes through both glass substrates <strong>on</strong>ly <strong>on</strong>ce and thusdoes not accumulate enough phase shift to be detected.Mode splittingThe relative phase shift δ between two orthog<strong>on</strong>al polarizati<strong>on</strong> modes (λ a , λ b ), whichis acquired <strong>on</strong> each round trip in the cavity, defines a frequency splitting between thesemodes, ν split = ν b − ν a [86]. Thus, the birefringence appears as a splitting of the cavitytransmissi<strong>on</strong> line. Since the round-trip phase shift <strong>for</strong> a mode λ i is given bythe mode splitting is proporti<strong>on</strong>al to δ and readsδ i = 2π 2Lλ i, (3.13)ν split = δ c2π 2L . (3.14)Here, L denotes the cavity length.The relevance of this mode splitting can be determined by comparing it to the cavitylinewidth. Since ν FWHM = ν FSR /F = c/(2LF), the ratio of the mode splitting to thecavity linewidth readsν split= δν FWHM 2π F . (3.15)It becomes clear, that <strong>for</strong> a high finesse of F = 10 6 even a tiny phase shift of δ = 10 −5can be readily observed.


3.2 High-reflectivity mirrors 510.7<strong>Cavity</strong> output intensity [V]0.60.50.40.30.20.10.00 5 10 15 20 25 30 35 40 45Time [µs]Figure 3.5: Birefringence of high-reflectivity mirrors measured using the CRD setup.The two polarizati<strong>on</strong> modes interfere while leaking out of the cavity. The oscillati<strong>on</strong>frequency is equal to the mode splitting. The thin line is a fit, corresp<strong>on</strong>ding to equati<strong>on</strong>(3.16), with fit parameters listed in Table 3.1. The graph represents a single traceresult.Measuring the birefringenceThe most general way to measure the cavity birefringence is to directly measure thepeak splitting ν split by scanning the length of the cavity and recording its double-peaktransmissi<strong>on</strong>. This method is simple if ν split is larger than the cavity linewidth, ν FWHM ,and the laser linewidth, ∆ν laser . But, if the splitting is small and the peaks are not wellresolved, the accuracy of this measurement might not be satisfactory. Several differentways to measure the small birefringence of high-finesse cavities have been extensivelystudied in the group of J. Kimble, see e.g. [86]. A general technique is based <strong>on</strong> scanningthe cavity length over the entire cavity res<strong>on</strong>ance profile, injecting light with a well-definedpolarizati<strong>on</strong>, and detecting the polarizati<strong>on</strong> of the transmitted light <strong>on</strong> a rotatable linearpolarizer.We use an alternative method to measure the birefringence by employing our currentCRD setup. Here, if both polarizati<strong>on</strong> modes of the cavity are filled with light, they startto leak out of the cavity simultaneously after switching off an incident laser source. If thenthese, initially orthog<strong>on</strong>al, polarizati<strong>on</strong> modes with different optical frequencies are mixed<strong>on</strong> some optical elements (e.g., <strong>on</strong> a mirror, a beam splitter, or a detector active surface),they start to interfere. The resulting beat signal can be detected <strong>on</strong> top of the overallCRD decay. Moreover, if both modes have the same decay rate, we get the exp<strong>on</strong>entialdecay modulated at their beat frequency with a c<strong>on</strong>stant c<strong>on</strong>trast. A typical signal of suchinterference is shown in Fig. 3.5.


52 Chapter 3: The high-finesse optical res<strong>on</strong>atorThe fit functi<strong>on</strong> to the oscillating decay has the following <strong>for</strong>m:()T (t) = A exp(−t/τ) 1 + C sin(2πνt + φ) , (3.16)where C is the c<strong>on</strong>trast of the oscillati<strong>on</strong>s, ν and φ are the frequency and the phase ofthe oscillati<strong>on</strong>s, respectively. The resulting fit parameters are listed in Table 3.1. Themeasured oscillati<strong>on</strong> frequency gives the frequency difference between two polarizati<strong>on</strong>modes of the cavity.amplitude A 0.535 ± 0.001 Vdecay time τ 11.02 ± 0.01 µsc<strong>on</strong>trast C 37.5 ± 0.1 %oscillati<strong>on</strong> frequency ν 195.1 ± 0.1 kHzphase φ 70 ± 3 mradTable 3.1: Fit results <strong>for</strong> the oscillating cavity ring-down signal of Fig. 3.5, accordingto equati<strong>on</strong> (3.16).Analysis and experimental observati<strong>on</strong>sIf two mirrors composing a cavity have almost the same birefringent phase shift, the overallphase shift can vanish <strong>for</strong> a specific relative orientati<strong>on</strong> of the mirrors, thus reducing theoscillati<strong>on</strong> frequency to zero. On the other hand, if the mirrors show a very different degreeof birefringence, the overall phase shift cannot vanish and oscillati<strong>on</strong>s are present <strong>for</strong> anymirror orientati<strong>on</strong>. This effect of an orientati<strong>on</strong>-dependent beat frequency was observed<strong>for</strong> different pairs of mirrors by rotating <strong>on</strong>e of them with respect to the other <strong>on</strong>e.The c<strong>on</strong>trast of the oscillati<strong>on</strong>s depends both <strong>on</strong> the relative populati<strong>on</strong> of the modesand <strong>on</strong> the degree of mixing of both polarizati<strong>on</strong>s after the cavity. The oscillati<strong>on</strong> canvanish, if the input laser polarizati<strong>on</strong> exactly matches <strong>on</strong>e of the cavity polarizati<strong>on</strong> modes.The input polarizati<strong>on</strong> can be set by rotating a λ/2-plate placed be<strong>for</strong>e the cavity holder.We have observed that the c<strong>on</strong>trast of the oscillati<strong>on</strong>s <strong>for</strong> each measurement staysc<strong>on</strong>stant during the decay, as in Fig. 3.5. This means, that the cavity decay rate is thesame <strong>for</strong> both polarizati<strong>on</strong> modes and, c<strong>on</strong>sequently, the mirrors do not show noticeablepolarizati<strong>on</strong>-dependent losses, such as absorpti<strong>on</strong>, scattering, or transmissi<strong>on</strong>.The measured birefringence varies str<strong>on</strong>gly within a charge of manufactured highreflectivitymirrors. The maximal observed mode splitting was about 200 kHz, corresp<strong>on</strong>dingto a phase shift of δ = 0.8 × 10 −4 rad. At the same time some pairs showed nodetectable splitting. Thus, the best candidates to c<strong>on</strong>stitute a high-finesse cavity <strong>for</strong> theplanned experiments are mirrors, which have the largest reflectivity (the l<strong>on</strong>gest phot<strong>on</strong>lifetime), while possessing negligible birefringence.Limits of the methodIf the birefringence is large, i.e. if the mode splitting is larger than both the cavity and laserlinewidth, the laser may be switched off after triggering <strong>on</strong> the transmissi<strong>on</strong> of the first


3.2 High-reflectivity mirrors 53cavity mode be<strong>for</strong>e the sec<strong>on</strong>d mode becomes c<strong>on</strong>siderably populated. Of course, in thiscase, no interference is possible and the CRD method fails to measure the birefringence.The smallest measurable birefringence corresp<strong>on</strong>ds to the slowest beat signal betweenthe polarizati<strong>on</strong> modes which can be detected with the current method. We have found,that the slowest oscillati<strong>on</strong>s, which still can be identified <strong>on</strong> top of the cavity decay signal,have a period of T osc ≈ 3.6 τ. Since ν split = 1/T osc and ν FWHM = 1/(2πτ), the lowest limit<strong>on</strong> measuring a mode splitting is thus ν split ≈ 1.8 ν FWHM .If the mode splitting is even smaller, the induced oscillati<strong>on</strong>s cannot be observed <strong>on</strong> topof the relative fast cavity decay. Although oscillati<strong>on</strong>s are not detectible in this case, theystill can lead to a small deviati<strong>on</strong> of the decay curve from the ideal exp<strong>on</strong>ential <strong>on</strong>e. Thisthen unavoidably results in a decay time, obtained from an exp<strong>on</strong>ential fit, significantlysmaller (up to 20 %) than the real cavity decay time τ.The c<strong>on</strong>venti<strong>on</strong>al limit <strong>for</strong> the spectral resoluti<strong>on</strong> of two close lines from a spectrum isν split ≈ ν FWHM , thus being lower than that of the presented method based <strong>on</strong> CRD effect.However, the main advantages of our approach compared to the direct mode splittingmeasurement are that it does not require a cavity scan to be linear, does not need itsfrequency calibrati<strong>on</strong>, and does not require a laser linewidth to be smaller than that ofthe cavity.Birefringence in cavity <strong>QED</strong> experimentsIn case of str<strong>on</strong>g birefringence (ν split >ν FWHM ) <strong>on</strong>ly linear polarized light can be injectedinto the cavity. Thus, the birefringence makes it difficult <strong>for</strong> the cavity field to drive <strong>on</strong>lyσ + or σ − transiti<strong>on</strong>s in an atom coupled to it. On the other hand, if the mode splittingis comparable to the cavity linewidth, an atom can be simultaneously coupled to bothcavity modes, which then can be treated as two different cavities. This significantlycomplicates the treatment of the atom-cavity system and should be avoided.The investigati<strong>on</strong> of the birefringence in our set of high-finesse mirrors allowed us toselect several mirrors showing the best per<strong>for</strong>mance (that is the smallest birefringenceif tested together with other mirrors), and thus being the most applicable <strong>for</strong> cavityc<strong>on</strong>structi<strong>on</strong>. Later <strong>on</strong>, the birefringence of the assembled cavity after gluing and backingthe mirrors is studied by scanning through the cavity res<strong>on</strong>ance and directly measuringmode splitting, as will be presented and discussed in Sec. 3.3.4.3.2.5 <strong>Cavity</strong> ringingAn alternative method to measure cavity finesse in based <strong>on</strong> the so-called “cavity ringing”phenomen<strong>on</strong>, see Ref. [87] and references therein. The transmissi<strong>on</strong> of a scanning Fabry-Perot interferometer is given by the well-known Airy functi<strong>on</strong>, and <strong>for</strong> a finesse of F 10the transmissi<strong>on</strong> peak approximately has a Lorentzian profile. However, if the cavityis swept over the laser frequency faster than the cavity decay time, the decaying cavityoutput field shows an amplitude modulati<strong>on</strong> with increasing frequency, known as cavity


54 Chapter 3: The high-finesse optical res<strong>on</strong>ator0.7<strong>Cavity</strong> transmissi<strong>on</strong> [V]0.60.50.40.30.20.10.00 20 40 60 80 100Scan time [µs]Figure 3.6: <strong>Cavity</strong> ringing. If the cavity scan over the laser frequency is faster than thecavity decay time, the recorded transmissi<strong>on</strong> signal str<strong>on</strong>gly deviates from that given bythe Airy functi<strong>on</strong> and shows speeding up oscillati<strong>on</strong>s <strong>on</strong> the tail of the transmissi<strong>on</strong> peak.ringing, see Fig. 3.6. The physical insight into this behavior can be obtained by c<strong>on</strong>sideringthe interference between the probe laser and the intracavity field, the frequency differenceof which c<strong>on</strong>tinuously increases, since the probe laser frequency is c<strong>on</strong>stant, while thefrequency of the intracavity field is c<strong>on</strong>tinuously shifted due to the cavity scan.From a classical point of view, the cavity in this problem may be seen as a damped harm<strong>on</strong>icoscillator with a time-dependent res<strong>on</strong>ance frequency driven by an external periodic<strong>for</strong>ce (i.e. by a laser field). The differential equati<strong>on</strong> describing the oscillator’s dynamicsis solved in Appendix B. Figure 3.7 shows the time dependence of the cavity output intensity<strong>for</strong> realistic scan parameters. This dependence reproduces well the experimentallymeasured cavity ringing, shown in Fig. 3.6.A more complicated decay signal can be observed if both the cavity ringing and thebeat of the polarizati<strong>on</strong> modes (see Sec. 3.2.4) come into play, as can be seen in Fig. 3.8.From this time dependence <strong>on</strong>e can in principle simultaneously get in<strong>for</strong>mati<strong>on</strong> aboutboth these effects. However, the analytical analysis of the curve <strong>for</strong>m becomes even morecomplex.<strong>Cavity</strong> ringing and finesse measurementSince the <strong>for</strong>m of the ringing str<strong>on</strong>gly depends <strong>on</strong> the relati<strong>on</strong> between the sweep timeand the cavity decay time, by analyzing the cavity ringing <strong>on</strong>e can determine the cavityfinesse [87]. The advantage of such a measurement is that it does not require switching anincident laser field <strong>on</strong> and off. However, a linear and c<strong>on</strong>tinuous cavity scan is essential.If it is not the case, the analytical signal <strong>for</strong>m of equati<strong>on</strong> (B.7) is not valid anymore.Un<strong>for</strong>tunately, our bulky cavity holder with clamped PZT, initially designed <strong>for</strong> CRDmeasurements, does not support a linear cavity scan. As seen in Fig. 3.6, there are severalbends <strong>on</strong> the signal tail. Besides, the presence of several close res<strong>on</strong>ance lines, e.g., of two


3.2 High-reflectivity mirrors 551.0Normalized cavity output intensity0.80.60.40.20.00 20 40 60 80 100Scan time [a.u.]Figure 3.7: Theoretical cavity ringing signal as given by equati<strong>on</strong> (B.7) <strong>for</strong> realisticexperimental parameters (see Appendix B <strong>for</strong> details).1.2<strong>Cavity</strong> transmissi<strong>on</strong> [V]1.00.80.60.40.20.00 20 40 60 80 100Scan time [µs]Figure 3.8: <strong>Cavity</strong> ringing and birefringence as seen from a single shot scan over thecavity res<strong>on</strong>ance. The slow c<strong>on</strong>stant oscillati<strong>on</strong> is the beat of the two cavity polarizati<strong>on</strong>modes, see Sec. 3.2.4, while the faster oscillati<strong>on</strong>s with increasing frequency representthe cavity ringing.polarizati<strong>on</strong> modes in our case (see Sec. 3.2.4), badly distorts the ringing signal and alsoprevents a robust finesse measurement compared to the CRD approach.By very rapidly scanning the cavity, i.e. if the scan over the cavity res<strong>on</strong>ance is muchfaster than the cavity decay, the cavity output does not show significant ringing, insteadit changes into an exp<strong>on</strong>entially decaying signal [88]. This effect can be used to realizea cavity ring-down experiment that does not require turning a probe laser <strong>on</strong> and off.Besides, it does not suffer from a scan n<strong>on</strong>-linearity as in the case of the measuring finesse


56 Chapter 3: The high-finesse optical res<strong>on</strong>ator(a) Side view(b) Top viewcavity holdershear PZTc<strong>on</strong>veyor beltcavitymirrorcavitymodeMOTlasers( z beams)MOT150 µm4.6 mmMOT lasers(in x-y plane)Figure 3.9: Schematic view of the high-finesse cavity integrated into our atom-trappingexperiment. All elements and laser beams are approximately drawn to scale. Not showndimensi<strong>on</strong>s are the waist of the MOT beams (1 mm), the MOT diameter (10 µm), thewaist of the dipole trap beams (35 µm), and the cavity mode waist (23 µm).from a ringing decay. However, because of the fast scan, the probe laser will not haveenough time to significantly excite the cavity mode resulting in a reduced signal amplitudeof such a measurement.The cavity ringing is a promising method of measuring the finesse of low-loss res<strong>on</strong>ators.However, since our optical setup is better suitable <strong>for</strong> CRD-based measurements, we usethe ring-down approach <strong>for</strong> determining reflectivity of our mirrors.3.3 High-finesse cavityOur cavity c<strong>on</strong>sists of two spherical mirrors with a radius of curvature of 5 cm <strong>on</strong> superpolishedsubstrates with diameter of 3 mm c<strong>on</strong>ed to 1 mm at their tips. The cavitylength is 156 µm, the high-reflectivity dielectric coating results in a finesse of about 10 6at 852 nm. The cavity is assembled <strong>on</strong> a specially designed cavity holder, which allows usto integrate the cavity into our main atom-trapping experimental setup in a most efficientway and to align it <strong>on</strong>to the dipole trap, see Sec. 3.3.1. The cavity length is stabilizedby means of a “lock chain”, which is developed <strong>for</strong> transferring the stability of <strong>on</strong>e of thecesium transiti<strong>on</strong>s to the high-finesse cavity, see Sec. 3.3.2. Finally, the cavity is fullycharacterized in a series of measurements, see Secs. 3.3.4–3.3.5.


3.3 High-finesse cavity 57vacuum chamberxzywindowdipole trapglasscellMOTMOTbeamscavity holdercavitycardan jointxyz-positi<strong>on</strong>er5 cmFigure 3.10: <strong>Cavity</strong> holder in the vacuum setup (top view).3.3.1 AssemblyIntegrati<strong>on</strong> into the atom-trapping experimentWe aim to use our dipole trap to transport <strong>on</strong>e or few single atoms, initially cooled,trapped, and prepared at the positi<strong>on</strong> of the magneto-optical trap (MOT), into the modeof the high-finesse cavity. A schematic view of the integrati<strong>on</strong> of the high-finesse cavity intothe atom-trapping experiment is shown in Fig. 3.9. To provide lossless atom transport,the cavity should be placed close to the MOT. At the same time, the relatively largemirror substrates should not block or disturb the laser beams composing the MOT. Thedistance of about 5 mm between the MOT and the cavity mode allows us both to efficientlytransport atoms and not to interfere with the laser beams. Besides, the c<strong>on</strong>ical substrateshape leaves more space <strong>for</strong> the MOT laser beams in the x-y plane.The cavity clearance should allow the high-power laser beams of the dipole trap to passbetween the two mirrors without being scattered, absorbed, or significantly diffracted <strong>on</strong>their edges. For the waist of the dipole trap beams of w 0 ≈ 35 − 40 µm the chosen cavitylength of 150 µm should be sufficient <strong>for</strong> this purpose.<strong>Cavity</strong> holderThe main functi<strong>on</strong> of our cavity holder is to hold and locate the cavity mirrors insideour narrow glass cell about 5 mm away from the MOT, which is operating rather deepinside the cell, see Figs. 3.10 and 3.11. Two mirrors selected <strong>for</strong> maximum reflectivity andminimum birefringence are glued <strong>on</strong>to shear piezoelectric transducer (PZTs) attached to


58 Chapter 3: The high-finesse optical res<strong>on</strong>atorFigure 3.11: Assembled cavity integrated into the vacuum glass cell.the aluminium holder. Details of the gluing procedure, including details <strong>on</strong> aligning themirrors and c<strong>on</strong>trolling their separati<strong>on</strong>, are described in Ref. [81].After installing the cavity inside the glass cell and pumping out the vacuum chamberwe have found that <strong>on</strong>ly <strong>on</strong>e PZT has an electric c<strong>on</strong>necti<strong>on</strong> to the outside of the chamberand thus can be used. Still, by driving <strong>on</strong>e of the PZTs, the cavity length can be scannedover about <strong>on</strong>e and a half of its free spectral range.The l<strong>on</strong>g and rigid holder leaves enough space <strong>for</strong> the dipole trap beams and guidesthe electric wires from the PZTs. The positi<strong>on</strong> of the dipole trap is fixed by the MOTpositi<strong>on</strong> and by a set of apertures <strong>for</strong> the DT beams placed outside the vacuum chamber.Thus, the cavity holder is made adjustable to permit the alignment of the cavity mode<strong>on</strong>to the dipole trap by moving the cavity in two dimensi<strong>on</strong>s perpendicular to thetrap axis. For this purpose, the holder rests inside the glass cell <strong>on</strong> a short piece ofbellows, acting both as a pivot and as a spring <strong>for</strong> crude vibrati<strong>on</strong> isolati<strong>on</strong> of the cavityholder, and is c<strong>on</strong>nected via a cardan joint to a linear XYZ-positi<strong>on</strong>er, c<strong>on</strong>sisting of aXY -manipulator and Z-feedthrough, as shown in Fig. 3.10. This combinati<strong>on</strong> allows usto adjust the cavity positi<strong>on</strong> with micrometer precisi<strong>on</strong> relative to the trap axis. Movingthe cavity al<strong>on</strong>g the dipole trap, which is achieved by sliding the bellow <strong>on</strong> its supportbase, sets the desired distance to the MOT.3.3.2 Frequency stabilizati<strong>on</strong><strong>Cavity</strong> <strong>QED</strong> experiments require a precise c<strong>on</strong>trol of the res<strong>on</strong>ance frequency of the cavityrelative to the atomic transiti<strong>on</strong> frequency. The corresp<strong>on</strong>ding precisi<strong>on</strong> should be muchbetter than the cavity linewidth. However, any instability of the cavity length, caused by,e.g., thermal drifts and mechanical (acoustical) vibrati<strong>on</strong>s, inevitable impairs the cavityfrequency stability, since δω/ω = −δL/L. Thus, if we would like to keep the cavity


3.3 High-finesse cavity 59probelaser852 nmAOM 1EOMPDtransfer cavityhigh finessecavityservoamplifierAOM 3res<strong>on</strong>antAPDLO 120 MHzAOM 2LO 280 MHzlocklaser836 nmPSservoamplifierservoamplifierFigure 3.12: Frequency stabilizati<strong>on</strong> of the high-finesse cavity. AOM: acousto-opticmodulator, EOM: electro-optic modulator, PD: photodiode, APD: avalanche photodiode,LO: local oscillator, PS: RF power splitter. The probe laser is stabilized to cesiumpolarizati<strong>on</strong> spectroscopy. The three servo amplifiers are based <strong>on</strong> the PDH method. AllAOMs are set up in the double-pass c<strong>on</strong>figurati<strong>on</strong>.res<strong>on</strong>ance stabile to within a fracti<strong>on</strong> A of its linewidth, the stability of the cavity lengthshould be better thanδL = Aλ2F . (3.17)For a high-finesse cavity with F = 10 6 and a required relative stability of, e.g., A = 0.1we need a stability of δL = 43 fm.The cavity’s res<strong>on</strong>ance frequency is locked to a reference “lock” laser using the Pound-Drever-Hall (PDH) method [89]. This method utilizes a phase-modulated lock laser andderives an error signal <strong>for</strong> an electr<strong>on</strong>ic servo loop by demodulating the laser power reflectedoff the cavity and detected by a fast photodetector. Our locking scheme <strong>for</strong> the stabilizati<strong>on</strong>of the high-finesse cavity is similar to the <strong>on</strong>e used in the group of J. Kimble [90].The comp<strong>on</strong>ents and per<strong>for</strong>mance of our setup are described in detail in Ref. [80, 81] andsummarized in Ref. [51]. Here, I present a short overview of the main lock elements,schematically depicted in Fig. 3.12.The lock laser is a diode laser in Littrow c<strong>on</strong>figurati<strong>on</strong> emitting at 836 nm. Thechosen laser frequency is far blue-detuned with respect to the D 2 transiti<strong>on</strong> line of cesium,which insures a small scattering rate of an atom in the cavity, as it is required <strong>for</strong> cavity<strong>QED</strong> experiments (<strong>for</strong> more details <strong>on</strong> the influence of the lock laser <strong>on</strong> trapped atomssee Sec. 4.5.1). Yet, this frequency still lies within the window of high reflectivity of themirror coating, providing the high finesse necessary <strong>for</strong> cavity stabilizati<strong>on</strong>. Because of theabsence of easily accessible atomic frequency standards at this wavelength, the lock laseris locked to an auxiliary cavity, which transfers the frequency stability from the stable


60 Chapter 3: The high-finesse optical res<strong>on</strong>atorprobe laser to the lock laser. Both servo loops <strong>for</strong> locking the transfer cavity and the locklaser are based <strong>on</strong> PDH method.The probe laser is derived from the cooling laser used to drive our magneto-opticaltrap, see Sec. 1.2. Its frequency is locked to the crossover signal of the F = 4 → F ′ = 3and the F =4 → F ′ =5 transiti<strong>on</strong>s using a Doppler-free polarizati<strong>on</strong> spectroscopy [91, 92],such that its frequency lies about 225 MHz below the F = 4 → F ′ = 5 transiti<strong>on</strong>. Thislaser is used both as a reference laser <strong>for</strong> locking the transfer cavity and as a probe laser<strong>for</strong> the cavity <strong>QED</strong> experiments. The part of the beam which is used <strong>for</strong> locking is phasemodulatedby an EOM res<strong>on</strong>antly driven at 20 MHz. The reflecti<strong>on</strong> off the transfer cavityis detected with a fast photodiode. A servo amplifier extracts and amplifies an error signal,which is then sent to a piezoelectric actuator attached to the transfer cavity, compensating<strong>for</strong> cavity frequency drifts.The frequency modulati<strong>on</strong> of the lock laser is realized by direct modulati<strong>on</strong> of the diodecurrent at 80 MHz. The reflecti<strong>on</strong> of the lock laser beam off the transfer cavity is detectedwith the same photodiode used <strong>for</strong> detecting the probe laser. After being demodulatedand amplified, the PDH error signal is fed back to the grating and to the current of thelock laser.Since in a high-finesse cavity the incident power of a res<strong>on</strong>ant laser is drastically enhancedby a factor <strong>on</strong> the order of F/π, we have to use the smallest possible lock laserpower <strong>for</strong> locking our <strong>QED</strong> cavity in order not to affect a trapped atom. For instance,170 nW of lock laser power coupled into the high-finesse cavity results in a scattering rateof 10 phot<strong>on</strong>s per sec<strong>on</strong>d, see Sec. 4.5.1. For detecting a weak reflected lock beam we utilizea home-built avalanche photodiode followed by a res<strong>on</strong>ant amplifier. After demodulati<strong>on</strong>,the PDH error signal is processed by a proporti<strong>on</strong>al-integral servo amplifier, amplified bya fast low-noise high-voltage amplifier [51], and finally sent to the shear PZT supportingthe cavity mirrors.The acousto-optic modulator AOM 1 (see Fig. 3.12) is used to preset the frequency ofthe probe laser. AOM 2 is used to scan the locked high-finesse cavity with respect to theprobe laser. The probe laser frequency can be tuned over the cavity res<strong>on</strong>ance by means ofAOM 3. In this way, both frequencies can be varied around the atomic res<strong>on</strong>ance: the locklaser over about ±100 MHz and the probe laser over about ±200 MHz (see also Sec. 4.5.1).Summarizing, the described stabilizati<strong>on</strong> scheme is sufficient to stabilize the res<strong>on</strong>ancefrequency of the high-finesse cavity to an atomic res<strong>on</strong>ance with about 100 nW ofcoupled lock laser power. The residual frequency fluctuati<strong>on</strong>s are about 0.2 ν FWHM rms.The cavity remains locked during executi<strong>on</strong> of different experimental sequences (seeChapter 4), even if they use optical shutters mounted to the optical table and producingmechanical disturbance to the cavity length. More importantly, the cavity lock can followrapid thermal expansi<strong>on</strong> and c<strong>on</strong>tracti<strong>on</strong> of the mirror substrates caused by switching thestr<strong>on</strong>g dipole laser beams <strong>on</strong> and off within 100 ms.


3.3 High-finesse cavity 613.3.3 Detecti<strong>on</strong> of the cavity transmissi<strong>on</strong>The main in<strong>for</strong>mati<strong>on</strong> obtained in experiments, where a coupled atom-cavity system isprobed by an external laser, is c<strong>on</strong>tained in the transmissi<strong>on</strong> of this laser through thecavity, see Sec. 4.2. Since the probe laser field must be very weak, we must be able todetect the transmitted probe power at the level of single phot<strong>on</strong>s. Below I present ouroptical setup allowing us to fulfil this task.Coupling light into and out of the cavityThe optical setup around the cavity c<strong>on</strong>sists basically of two parts: <strong>on</strong>e <strong>for</strong> stabilizing thecavity length and the other <strong>for</strong> detecting the transmissi<strong>on</strong> of the probe laser through thecavity. In the laboratory the whole setup is distributed over two optical tables c<strong>on</strong>nectedby optical fibers. The first table c<strong>on</strong>tains the lock chain, except <strong>for</strong> the res<strong>on</strong>ant APD andthe high-finesse cavity, as well as the elements of the frequency and power c<strong>on</strong>trol of thelock and probe lasers, see Fig. 3.12. The optical system <strong>for</strong> detecting the cavity reflecti<strong>on</strong>and transmissi<strong>on</strong> is located <strong>on</strong> the sec<strong>on</strong>d table c<strong>on</strong>taining the vacuum chamber with thehigh-finesse cavity and the optical traps. Its main comp<strong>on</strong>ents are schematically shown inFig. 3.13. They include the optics <strong>for</strong> coupling the probe and lock lasers into the cavitymode, detecting the reflecti<strong>on</strong> of the lock laser <strong>for</strong> locking the cavity, and separati<strong>on</strong> andindependent detecti<strong>on</strong> of the transmitted lasers.Both lock and probe lasers are guided from the first table to the sec<strong>on</strong>d <strong>on</strong>e by meansof <strong>on</strong>e optical fiber. It provides perfect geometrical overlap of the beams, their high spatialstability, and nearly Gaussian spatial mode profile be<strong>for</strong>e the cavity. The combinati<strong>on</strong> ofthe two zero-order wave plates, λ/4 and λ/2, allows us to arbitrary set the polarizati<strong>on</strong> ofthe incoming lasers, <strong>for</strong> instance, <strong>for</strong> populating <strong>on</strong>ly <strong>on</strong>e polarizati<strong>on</strong> mode of the cavity,see Sec. 3.3.4. Since the probe and the lock laser are coupled into the fiber in orthog<strong>on</strong>allinear polarizati<strong>on</strong>s, they also stay orthog<strong>on</strong>al be<strong>for</strong>e the cavity. The lasers are coupledinto the cavity mode by using a mode matching lens system and a pair of mirrors. Thereflecti<strong>on</strong> off the cavity is deflected by a n<strong>on</strong>-polarizing 50/50 beam splitter (BS) <strong>on</strong>tothe res<strong>on</strong>ant APD, which is used <strong>for</strong> locking the cavity <strong>on</strong>to the lock laser, see Sec. 3.3.2.Being designed to detect and amplify <strong>on</strong>ly a signal modulated at 80 MHz, the res<strong>on</strong>antAPD is not sensitive to the probe laser light.After passing through the cavity and being again collimated, a part of the transmittedlight is reflected by a removable BS <strong>on</strong>to a camera CCD 1. It provides in<strong>for</strong>mati<strong>on</strong> <strong>on</strong> thetransversal mode structure of the two beams. In additi<strong>on</strong> this camera allows us to observethe scattering of the DT laser beams off the cavity mirrors, which is very useful during theDT alignment, see Sec. 3.4. After the DT is aligned and the desired transversal mode ofthe probe laser is chosen, the BS is removed to avoid additi<strong>on</strong>al losses of the transmittedlight.Separating the two lasersThe lock and probe lasers with a wavelength difference of about 16 nm are spatially separated<strong>on</strong> a diffracti<strong>on</strong> grating (Thorlabs Inc., model GR25-1210) ruled with 1200 lines/mm


62 Chapter 3: The high-finesse optical res<strong>on</strong>atormode matching and lock laser reflecti<strong>on</strong>res<strong>on</strong>antAPDcavityfiberoutcoupler l/4 l/2BSmode matchingclass cellcavity transmissi<strong>on</strong> and separati<strong>on</strong> of the two beamsCCD 1removableBSgratingprobe laser transmissi<strong>on</strong>lock lasertransmissi<strong>on</strong>APD 1BSpinholeIFfibercouplerflip-mirrorSPCMCCD 2APD 2Figure 3.13: Schematic optical setup <strong>for</strong> coupling the lock and probe lasers into thecavity and detecting their transmissi<strong>on</strong>. BS: n<strong>on</strong>-polarizing beam splitter, λ/4 and λ/2:quater- and a half-wave plate, respectively, CCD: infrared CCD camera, IF: interferencefilter, APD: avalanche photo-diode, SPCM: single phot<strong>on</strong> counting module.and a blaze wavelength of 1000 nm. The grating is set up in Littrow c<strong>on</strong>figurati<strong>on</strong>,i.e. the incident angle is close to the blaze angle of 37 0 . With the designed dispersi<strong>on</strong> of0.67 nm/mrad the beams are separated by about 10 mm after a distance of 40 cm from thegrating. Having a waist of 1 mm they are then easily separated by mirrors. To increasethe reflectivity of the grating surface, it is coated with gold. The measured diffracti<strong>on</strong>


3.3 High-finesse cavity 63efficiency, η grating , i.e. the relative light power in the first diffracti<strong>on</strong> order, is 76(±2) % at836 nm and 71(±2) % at 852 nm. The efficiency is polarizati<strong>on</strong> independent within themeasurement uncertainty.Additi<strong>on</strong>ally, the transmissi<strong>on</strong> of the probe laser is filtered by an interference filter(company “Dr. Hugo Anders”). It has a transmissi<strong>on</strong> of η filter = 0.77 at 852 nm and about10 −4 <strong>for</strong> the 836 nm lock laser light.Spacial filteringThe probe beam is spatially filtered with a 70 µm pinhole, installed at the focus betweentwo c<strong>on</strong>vex lenses with 50 mm focal length. It helps to block stray light not originatingfrom the cavity mode as well as parasitic scattering and diffracti<strong>on</strong> off the ruled gratingdue to its spacing errors (“ghost” stray light).If the cavity transmissi<strong>on</strong> is detected by the single phot<strong>on</strong> counting module (SPCM,see below), additi<strong>on</strong>al mode cleaning is per<strong>for</strong>med by coupling the transmitted light into asingle-mode optical fiber directly attached to the detector via FC c<strong>on</strong>nector. The fiber isshielded from ambient room light by a furcati<strong>on</strong> tubing. The typically achieved collecti<strong>on</strong>efficiency, η fiber , of the fiber is about 60 % (fiber collimator from Schäfter+Kirchhoff, model60FC-4-A18-02 with FC plug c<strong>on</strong>necti<strong>on</strong>).Detecti<strong>on</strong> of the transmissi<strong>on</strong>In order to eliminate measurements when the cavity is not properly locked, we detect thetransmissi<strong>on</strong> of the lock laser by an analog APD during all measurements. In the caseof an unlocked cavity, the probe transmissi<strong>on</strong> drops to zero. Besides, the transmissi<strong>on</strong> ispermanently m<strong>on</strong>itored by the camera CCD 2 <strong>for</strong> visual inspecti<strong>on</strong> of the cavity stability.The transmitted probe laser light is detected either by an analog APD or by an SPCM,see below. The choice of the detector is realized by a flip-mirror.Analog APDThe first detector measuring the probe transmissi<strong>on</strong> is an avalanche photodiode(PerkinElmer, Si-APD, model C30902S) with a home-built transimpedance amplifier. Theminimal measurable light power is about 1 pW, resulting in 4.9 mV of the APD outputvoltage with about 20 % uncertainty. The output of the APD saturates at about 4 V,while the APD’s dark current causes a c<strong>on</strong>stant background of about 2 V. The high transimpedanceof 100 MΩ limits the detector’s bandwidth to 9 kHz.The analog APD detector can be used with higher laser power than comm<strong>on</strong> phot<strong>on</strong>countingmodules. This makes it an ideal detector <strong>for</strong> aligning purposes and in themeasurements testing the overall per<strong>for</strong>mance of the experimental setup where an overexposureof the detector is not improbable. Besides, it is used to measure the transmissi<strong>on</strong>spectrum of the probe laser, its transversal mode spectrum, and the cavity polarizati<strong>on</strong>modes. All measurements in this chapter <strong>on</strong> cavity characterizati<strong>on</strong> as well as the firsttest experiments of Sec. 4.4 with transporting many atoms into the cavity have been per<strong>for</strong>medwith this detector.


64 Chapter 3: The high-finesse optical res<strong>on</strong>ator<str<strong>on</strong>g>Single</str<strong>on</strong>g> phot<strong>on</strong> counting moduleTo detect low laser intensities we use a single phot<strong>on</strong> counting module (PerkinElmer,model SPCM-AQR-12-FC). Its maximum count rate is 10 Mc/s and the dark count rateis less than 500 c/s. The SPCM has an FC c<strong>on</strong>nector <strong>for</strong> attaching an optical fiber close toits active area. The measured phot<strong>on</strong> detecti<strong>on</strong> efficiency, η SPCM , of the SPCM is 0.30 at852 nm and c<strong>on</strong>sists of the detector’s quantum efficiency and of the coupling efficiency ofthe fiber light into the detector’s active area. The total detecti<strong>on</strong> efficiency of the phot<strong>on</strong>stransmitted through the cavity is thenη total = η optic · η grating · η filter · η fiber · η SPCM ≈ 0.089 , (3.18)where η optic = 0.9 denotes transmissi<strong>on</strong> of the light through all other optical elements <strong>on</strong>the beam path from the cavity to the detector (if the removable beam splitter is removed,see Fig. 3.13). The SPCM detector was used in experiments presented in Sec. 4.5 typicallydetecting 10–50×10 3 counts/s.<strong>Cavity</strong> c<strong>on</strong>trol softwareDuring cavity <strong>QED</strong> measurements, the detectors’ signals resulting from the detecti<strong>on</strong> ofthe cavity transmissi<strong>on</strong> <strong>for</strong> both lock and probe lasers are read out and stored <strong>on</strong> a computer<strong>for</strong> later data processing. For this purpose, the output voltages of the analog APDsare sent to two analog inputs of a high-speed data acquisiti<strong>on</strong> card (Nati<strong>on</strong>al Instruments,model PCI-6251). The output of the SPCM, which is put out as TTL pulses, are countedby a counter/timer card (Nati<strong>on</strong>al Instruments, model PCI-6601). Both cards are installedinto a standard pers<strong>on</strong>al computer. During a high TTL gate input, which is sentby an experiment c<strong>on</strong>trol software running the whole experiment [50], either the analogcard digitalizes the APD signal of the probe transmissi<strong>on</strong> with a preset sampling rate orthe counter card bins the phot<strong>on</strong> counts detected by the SPCM with a preset bin time,depending <strong>on</strong> the detector in use. At the same time the analog card processes the signalfrom the lock laser APD. Both the sampling rate and the bin time are preset by the cavityc<strong>on</strong>trol software and are typically 0.1–1 ms. In additi<strong>on</strong>, the same software reads out thecorresp<strong>on</strong>ding cards, plots traces of cavity transmissi<strong>on</strong> <strong>for</strong> both lasers, and saves them<strong>for</strong> their later analysis. Moreover, this software is used to detect the thermal drifts of thelocked cavity relative to the probe laser and to compensate <strong>for</strong> them by c<strong>on</strong>trolling thefrequency of the lock laser, <strong>for</strong> details see Sec. 3.5.3.3.4 Characterizati<strong>on</strong> of the cavityLength measurementOne of the most important cavity parameters is its length. It defines the free spectral range,the waist of the cavity mode, its volume, and c<strong>on</strong>sequently the atom-phot<strong>on</strong> couplingstrength. Be<strong>for</strong>e gluing the mirror substrates we tried to manually set their separati<strong>on</strong>to about 150 µm while observing the mirror spacing under a microscope. However, the


3.3 High-finesse cavity 65gluing and baking processes can change the set separati<strong>on</strong>. There<strong>for</strong>e, the cavity lengthshould be precisely measured again after the cavity is assembled.A widely used method to determine the cavity length is based <strong>on</strong> a measurement ofthe transverse mode spacing, ∆ν transv . For a symmetric res<strong>on</strong>ator∆ν transv =arccos( 1 − L/R )πc2L , (3.19)where R is the mirror’s radius of curvature (see e.g. [79]). In approximati<strong>on</strong> of a shortcavity, that is <strong>for</strong> L ≪ R, equati<strong>on</strong> (3.19) can be rewritten asL =2c 2(2π∆ν transv ) 2 R . (3.20)Despite of its c<strong>on</strong>ceptual simplicity, a precise measurement of ∆ν transv <strong>for</strong> finding L usuallyrequires a highly linear cavity scan, see e.g. [80, 86], which is difficult to realize. Indeed,the measured length of our cavity of 135 µm has a relatively large error of ±20 µm, whichresults from the scan n<strong>on</strong>linearity and uncertainty of the scan calibrati<strong>on</strong>.To avoid scanning the cavity <strong>for</strong> measuring its length, we tune the laser through a fullfree spectral range of the cavity. The two corresp<strong>on</strong>ding wavelengths λ 1 and λ 2 satisfy therelati<strong>on</strong> L = nλ 1 /2 = (n − 1)λ 2 /2. Thus, the cavity length is given byL =λ 1 λ 22(λ 2 − λ 1 ) . (3.21)In this way, we have measured the cavity length of L = 156.52(±0.04) µm. The freespectral range is ν FSR = ν 1 − ν 2 = 957.7(±0.3) GHz, corresp<strong>on</strong>ding to 2.3 nm at a852 nm wavelength. The mode waist of the probe laser field, given by equati<strong>on</strong> (3.5), isw 0 = 23.15 µm, and the mode volume, given by equati<strong>on</strong> (3.6), is V = 65897 µm 3 .Note that the measured L is an effective cavity length, L eff . It c<strong>on</strong>sists of the physicaldistance between the mirror surfaces, L phys , which is an integer number of λ/2, and thepenetrati<strong>on</strong> of the cavity electro-magnetic field into the multi-layer dielectric coating [86].Thus, in general, n is not an integer and in our case n 852 ≈ 367.4 and n 836 ≈ 374.4.Since the penetrati<strong>on</strong> depth depends <strong>on</strong> the wavelength, L eff <strong>for</strong> the probe and locklasers may differ. This issue becomes important <strong>for</strong> very short cavities, e.g., of L ≈ 10 µmused in Ref. [82]. In our case, the discrepancy between different L eff is small and doesnot significantly influence the calculati<strong>on</strong> of the cavity parameters important <strong>for</strong> us, e.g.,such as a mode waist and an atom-phot<strong>on</strong> coupling. However, in Sec. 3.5 we will see thatthe probe laser and the lock laser fields indeed have different penetrati<strong>on</strong> depths into thecoating stack and, thus, see slightly different cavity lengths.Finesse measurementGenerally speaking, the finesse of the assembled cavity differs from that measured withn<strong>on</strong>-glued mirrors in the CRD setup in Sec. 3.2.3. If <strong>on</strong>e is not careful enough, themirror coating may be damaged and/or mirror surfaces may be polluted during the cavity


66 Chapter 3: The high-finesse optical res<strong>on</strong>ator1.0(a) Probe laser 852 nm; F=1 120 000(b) Switching probe laser offwithout cavityNormalized light intensity0.80.60.40.2(d) (c) (b)(c) Lock laser 836 nm; F= 548 000(d) Switching lock laser offwithout cavity(a)0.00.0 0.5 1.0 1.5 2.0Time [µs]Figure 3.14: <strong>Cavity</strong> ring-down of the high-finesse cavity. The cavity decay time ismeasured <strong>for</strong> both the probe (a) and lock laser (c) wavelengths (852 nm and 836 nm,respectively). The influence of the detector bandwidth is investigated by recording thelight intensity while switching off the both lasers without the cavity, see curves (b)and (d). Each trace shown is an average over about 20 single shots. All traces arenormalized to 1 <strong>for</strong> better visibility.c<strong>on</strong>structi<strong>on</strong>, thus deteriorating their quality. In additi<strong>on</strong>, the mode of an assembled cavityprobes points at the mirror surface which are in general different from those tested withnot glued mirrors.The finesse of the assembled cavity is measured in situ, that is in the vacuum system,after the holder with the cavity is placed inside the glass cell. The method used is the cavityring-down, introduced in Sec. 3.2.2. The measured cavity decay at both the probe andlock laser wavelengths, i.e. at 852 nm and 836 nm, respectively, is shown in Fig. 3.14. Thedifference in switching-off times <strong>for</strong> the two lasers comes from the difference in switchingelectr<strong>on</strong>ics and optics (including different AOMs) used <strong>for</strong> the probe and lock lasers. Eachpresented trace is an average over about 20 single measurements, normalized to 1 <strong>for</strong> easiercomparis<strong>on</strong> of different curves. From the measured decay times we get the cavity finesseat both wavelengths: F 852 nm = 1.12 × 10 6 and F 835 nm = 0.55 × 10 6 as well as the cavitylinewidth at 852 nm of ω FWHM = 2π × 0.85 MHz.Compared to the CRD measurements with a cavity of 1 cm length, presented inSec. 3.2.2, the lifetime of phot<strong>on</strong>s inside the assembled 156 µm-l<strong>on</strong>g cavity is much shorter.This sets more stringent requirements <strong>on</strong> the bandwidth of the employed photodetector.For the current measurement we have used an avalanche photodiode with 9 MHz bandwidth.To investigate its influence <strong>on</strong> the <strong>for</strong>m of the recorded CRD signal, we recordthe signal when the probe or lock laser are switch off without passing through the cavity.The observed drop-off times of the detected signal, see traces (b) and (d) in Fig. 3.14,


3.3 High-finesse cavity 671.80<strong>Cavity</strong> transmissi<strong>on</strong> [a.u.]1.751.701.651.601.55112 116 120 124 128 132 136Frequency shift of the lock laser [MHz]Figure 3.15: Birefringent modes of the cavity. The length of the locked cavity is scannedby sweeping the frequency of the lock laser. Two transmitted peaks corresp<strong>on</strong>d to twopolarizati<strong>on</strong> modes of the cavity. The thick line is a fit to the data with a sum of twoLorentzian lines, resulting in a mode splitting of 3.86(±0.01) MHz. The peaks have afull width at half maximum of about 1.4 MHz.are significantly shorter than the cavity decay time, thus justifying the utilizati<strong>on</strong> of thedetector.Birefringence measurementDue to the gluing and backing processes, the stress-induced birefringence of an assembledcavity can be much larger than the inherent birefringence of the mirror coating,investigated in Sec. 3.2.4, and thus must be studied separately. In c<strong>on</strong>trast to n<strong>on</strong>-gluedmirrors, we measure the birefringence of the assembled cavity from the splitting of thetransmissi<strong>on</strong> peak by scanning the cavity. In order to eliminate drifts and fluctuati<strong>on</strong> ofthe cavity length during the measurement, we lock the cavity to the lock laser. By slowlyscanning the lock laser frequency using an acousto-optic modulator (in Fig. 3.12 denotedas AOM 2), the cavity length can be tuned while keeping the cavity locked. The cavitytransmissi<strong>on</strong> recorded in a single shot is presented in Fig. 3.15. Here, we have scannedthe AOM frequency over 14 MHz within 200 ms. For the better signal-to-noise, the inputlinear polarizati<strong>on</strong> of the probe light is set such that both polarizati<strong>on</strong> modes are approximatelyequally excited. This corresp<strong>on</strong>ds to input polarizati<strong>on</strong> rotated by about 45 o degreerelative to the birefringence axis of the cavity. The transmissi<strong>on</strong> spectrum is then fitted bya sum of two Lorentzian functi<strong>on</strong>s, resulting in a mode splitting of 3.86(±0.01) MHz. Themeasured linewidth at half maximum of 1.4 MHz is larger than ω FWHM /2π = 0.85 MHzexpected from the CRD measurement because of the finite linewidth of the lock laser.


68 Chapter 3: The high-finesse optical res<strong>on</strong>atorThe measured mode splitting is larger than that typically measured with ungluedcavities in Sec. 3.2.4. According to equati<strong>on</strong> (3.14), this difference is caused first of all byusing here the shorter cavity length. From relati<strong>on</strong> (3.15) we find the round-trip phaseshift between the polarizati<strong>on</strong> modes of δ ≈ 1.5 × 10 −5 . Since we have not observed thebirefringence with the unglued mirrors, which now compose our cavity, the intrinsic phaseshift between polarizati<strong>on</strong> modes, if any, was smaller than 10 −5 . The acquired birefringenceresults from the mechanical stress induced in the mirror substrates after gluing them andtransmitted to the coated surface.In general, the polarizati<strong>on</strong> directi<strong>on</strong> of the birefringent modes depends <strong>on</strong> thesymmetry of the induced mechanical strain. Since the glued surfaces of the mirrorsubstrates are perpendicular to the cavity holder, the resulted strain, and thus the modepolarizati<strong>on</strong>s, should be symmetrical with respect to plane c<strong>on</strong>taining the axes of boththe cavity and the holder. As expected, the measured polarizati<strong>on</strong>s of the two modes (seeFig. 3.15) are parallel and perpendicular to the holder axis, respectively. By setting thepolarizati<strong>on</strong> of the incident light, we choose which cavity mode to excite. All experimentspresented in Chapter 4 were per<strong>for</strong>med with the probe laser polarized parallel to thecavity holder axis and, thus, to the dipole trap axis.3.3.5 Mirror lossesSince high-finesse cavities str<strong>on</strong>gly enhance the intra-cavity light, <strong>on</strong>e should use verylow light intensities <strong>for</strong> most cavity <strong>QED</strong> experiments in combinati<strong>on</strong> with ultra sensitivelight detectors such as single phot<strong>on</strong> counting modules. In this case any cavity losses otherthan mirror transmissi<strong>on</strong> reduce the valuable transmitted light power and thus significantlydecrease the detecti<strong>on</strong> sensitivity. Below, I identify and measure the main losses of lightin our high-finesse cavity.A mirror can be characterized by three parameters, R, T , and A, the intensity reflecti<strong>on</strong>,transmissi<strong>on</strong>, and loss coefficients, respectively. Here, A includes both absorpti<strong>on</strong> andscattering losses. The coefficients are related through energy c<strong>on</strong>servati<strong>on</strong> by R+T +A = 1and can be calculated by analyzing the cavity transmissi<strong>on</strong> and reflecti<strong>on</strong> of res<strong>on</strong>ant laserlight [82, 83]. Since the method presented below does not allow us to distinguish betweenthe mirrors, in the following we suppose they are equivalent and thus have equal losses.For a m<strong>on</strong>ochromatic light of a power P in and a frequency ω incident <strong>on</strong> a Fabry-Perotres<strong>on</strong>ator, the reflected power, P r , and the transmitted power, P t , are given byP rP in= R(1 − R − T )2 + 4R(T + R) sin 2 φ 2(1 − R) 2 + 4R sin 2 φ 2P tP in=T 2(1 − R) 2 + 4R sin 2 φ 2, (3.22a). (3.22b)where φ = 2πω/ω FSR is the round-trip phase of the light field in the res<strong>on</strong>ator and ω FSRis its free spectral range. In case of res<strong>on</strong>ance, φ = 0 and equati<strong>on</strong>s (3.22b) and (3.22a)


3.3 High-finesse cavity 69are simplified toP rP in=P tP in=(1 − T − A)A2(T + A) 2 ≈A 2(T + A) 2 (<strong>for</strong> R ≈ 1) , (3.23a)T 2(T + A) 2 . (3.23b)For a res<strong>on</strong>ator without absorpti<strong>on</strong> losses (A = 0), as usually c<strong>on</strong>sidered in general textbooks<strong>on</strong> optics, these equati<strong>on</strong>s result in a total transmissi<strong>on</strong> of res<strong>on</strong>ant light. However,if A cannot be neglected, we can calculate the mirror losses, T and A, by measuring powersP in , P r , and P t . For this purpose, we make use of an additi<strong>on</strong>al equati<strong>on</strong> c<strong>on</strong>necting thecavity losses, which is based <strong>on</strong> a cavity finesse and readsF = π√ R1 − R ≈ πT + A(<strong>for</strong> R ≈ 1). (3.24)Un<strong>for</strong>tunately, in real experiments not all of the incident power P in is coupled into a cavityTEM 00 mode. If taking into account a mode matching factor, ε, the useful, coupled inputpower is reduced to εP in , whereas a power (1−ε)P in is directly reflected off the first cavitymirror without carrying any in<strong>for</strong>mati<strong>on</strong> <strong>on</strong> cavity properties. C<strong>on</strong>sequently, the extendedequati<strong>on</strong>s (3.23a) and (3.23b) should now readP r − (1 − ε)P inεP in=P tεP in=A 2(T + A) 2 , (3.25a)T 2(T + A) 2 . (3.25b)Finally, after making rearrangements in equati<strong>on</strong>s (3.23a)–(3.24), we getT =2α1 + αA = 1 − α1 + απF ,πF ,(3.26a)(3.26b)ε = P tP in(T + A) 2T 2 , (3.26c)with a subsidiary parameter α = P t /(P in − P r ).In our cavity setup, we measure the incident laser power P in using a powermeter, andthe powers P r and P t using calibrated APDs. For this purpose we scan the cavity lengthand determine the minimum cavity reflecti<strong>on</strong> and the maximum cavity transmissi<strong>on</strong>, respectively.Additi<strong>on</strong>al power losses <strong>on</strong> intermediate optical elements, e.g., <strong>on</strong> the glasscell, are approximately taken into account. The results of this measurement are summarizedin Table 3.2. The estimated uncertainty of measured losses is about 10 %. A highermirror absorpti<strong>on</strong> measured <strong>for</strong> the lock laser results from its deeper penetrati<strong>on</strong> into thedielectric coating compared to the probe laser.


70 Chapter 3: The high-finesse optical res<strong>on</strong>atorParameter 852 nm 836 nmphot<strong>on</strong> lifetime τ 187 ns 91 nscavity finesse F 1 120 000 548 000mirror transmissi<strong>on</strong> T 1.3 ppm 3.6 ppmmirror absorpti<strong>on</strong> A 1.8 ppm 2.7 ppmmode coupling ε 36 % 41 %Table 3.2: Measured mirror losses and mode coupling efficiency <strong>for</strong> the wavelengths ofthe probe and lock laser. Phot<strong>on</strong> lifetimes and finesse are those measured with the CRDmethod in Sec. 3.3.4.Because of the difference in Gaussian beam geometry <strong>for</strong> the probe and lock laserwavelengths, the cavity mode coupling <strong>for</strong> these beams slightly differs. Note that inexperiments in Chapter 4 the mode coupling efficiency was about 55 %. Alternatively,the efficiency of the mode coupling has been determined by comparing the height of thefundamental TEM 00 mode measured from the transmissi<strong>on</strong> spectrum of the cavity to thesum of the heights of all transversal modes which can be observed. This method hasresulted in a coupling efficiency of about 80 % <strong>for</strong> the probe laser. The origin of thediscrepancy between this value and that in Table 3.2 is unclear <strong>for</strong> us.Intra-cavity intensityBesides their importance <strong>for</strong> general mirror characterizati<strong>on</strong>, the mirror losses are necessary<strong>for</strong> calculating the intra-cavity light intensity affecting an atom coupled to the cavitymode. In analogy to equati<strong>on</strong>s (3.25a) and (3.25b), the intra-cavity power circulatinginside a res<strong>on</strong>ant Fabry-Perot res<strong>on</strong>ator is given byP intraεP in=T(T + A) 2 . (3.27)With (3.25b) this allows us to determine P intra by measuring the transmitted laser powerusing the obvious relati<strong>on</strong>P intra = 1 T P t . (3.28)Finally, from P intra , we calculate the intra-cavity light intensity at an antinode of thestanding wave of the cavity field asI max = 8P intraπw 2 0. (3.29)


3.3 High-finesse cavity 713.3.6 C<strong>on</strong>clusi<strong>on</strong>In the previous secti<strong>on</strong>s I have presented the c<strong>on</strong>structi<strong>on</strong> and characterizati<strong>on</strong> of the highfinesseoptical cavity applicable <strong>for</strong> the planned cavity <strong>QED</strong> experiments with trappedcesium atoms. The successful cooperati<strong>on</strong> with a mirror fabricator during manufacturingthe high-reflectivity mirrors allowed us to achieve a cavity finesse of about 10 6 . Afterthe additi<strong>on</strong>al investigati<strong>on</strong> of their properties, we have selected the best mirrors <strong>for</strong> thecavity.The specially designed cavity holder provides the efficient integrati<strong>on</strong> of the cavityinto our atom-trapping experimental setup. Here, the cavity mode can be placed at 5 mmdistance from our magneto-optical trap and, using the adjustable holder, can be preciselyaligned <strong>on</strong>to the dipole trap. Our stabilizati<strong>on</strong> scheme is able to stabilize the res<strong>on</strong>ancefrequency of the high-finesse cavity to an atomic res<strong>on</strong>ance with about 100 nW of coupledlock laser power. The residual frequency fluctuati<strong>on</strong>s are about 0.2 ν FWHM rms.Finally, the assembled high-finesse cavity was characterized in a series of measurements.The basic properties of the cavity are summarized in Table 3.3.Parameter Value CommentsGeometrycavity length L 156.52 µm effective, measured at 839 nmradius of curvature R 5 cm specified by manufacturermirror diameter D 1 mm specified by manufacturercavity clearance L c ≈ 150 µmTEM 00 mode waist w 0 23.15 µm <strong>for</strong> light at 852 nmmode size at the surface w 23.17 µmmode volume V 65 900 µm 3Spectroscopic propertiesfree spectral range ν FSR 957.7 THzfinesse F 1 120 000 from CRD at 852 nmF 836 nm 550 000 from CRD at 836 nmlinewidth ν FWHM 0.85 MHz from CRD at 852 nmpolarizati<strong>on</strong> mode splitting ∆ν split 3.86 MHz birefr. axis al<strong>on</strong>g the holder<strong>Cavity</strong> <strong>QED</strong> parametersatom-phot<strong>on</strong> coupling rate g/2π 18.0 MHz <strong>for</strong> F =4 → F ′ =5with σ + -polarized cavity modecavity field decay rate κ/2π 0.43 MHz from CRD at 852 nmatomic dipole decay rate γ/2π 2.61 MHz decay rate <strong>for</strong> cesium atomcooperativity parameter C 1 146.7Table 3.3: Properties of the high-finesse optical cavity. For mirror losses see Table 3.2.


72 Chapter 3: The high-finesse optical res<strong>on</strong>ator3.4 Interplay of the atom trap setup and the cavityBeing a massive object in the glass cell, the cavity with its holder can in principle influenceour atom trap setup, even though it is placed far from the MOT center and the mirrorspacing is larger than the size of the DT beams. Of course the atom traps can also affectthe cavity per<strong>for</strong>mance. In the following I investigate the possible cross-influence of theatom trap setup and the cavity.The MOTThe presence of the cavity does not noticeably prevent the normal operati<strong>on</strong> of our singleatom MOT: The cavity mirrors are located far enough from the MOT beams and thusdo not noticeably block them, and the n<strong>on</strong>-magnetic cavity holder does not influence themagnetic field of the MOT. However, the small scattering of the MOT lasers off the cavitymirrors increases the stray light seen by the APD detecting the MOT fluorescence rate(Sec. 1.2). In the absence of the cavity, the typical stray light-induced background countrate of the MOT APD is 2 × 10 4 counts/s. The presence of the cavity increases thisnumber to about 7 × 10 4 counts/s. Of course, the fluorescence rate of a single atom staysthe same, namely 3 × 10 4 counts/s. The increased background level results in an increasedstatistical noise of the detected fluorescence of the running MOT, thus decreasing themaximal faultlessly countable number of trapped atoms from 19 (see Ref. [56]) to about 10.However, since in the planned cavity <strong>QED</strong> experiments we aim <strong>on</strong> using few or even singleatoms, the scattering of the MOT beams off the cavity mirrors does not c<strong>on</strong>strain us.Due to the spatial filtering of the probe laser beam transmitted through the cavity,this scattering is not visible <strong>on</strong> the APD detecting the cavity transmissi<strong>on</strong> and thus alsodoes not influence our cavity transmissi<strong>on</strong> measurements.Modificati<strong>on</strong> of the dipole trap geometryThe main demand <strong>on</strong> the geometry of the DT beams is a reas<strong>on</strong>ably small beam size at theMOT positi<strong>on</strong> (to provide a deep trapping potential <strong>for</strong> efficient loading of the DT) andinside the cavity mode (to not clip the DT beams <strong>on</strong> the mirror edges). These two pointsare separated by 4.6 mm. The DT beams with a waist of w old = 25.6 µm, which were usedin the experiments described in Chapter 2, had a Rayleigh range of 2.0 mm. This smallrange would have resulted in a beam diameter of 128 µm inside the cavity, comparableto the clearance of 156 µm. There<strong>for</strong>e, <strong>for</strong> the cavity experiments we have increased thebeam waist to about w new = 36 µm, extending the Rayleigh range to 4.0 mm, and haveshifted the waist positi<strong>on</strong> midway between the MOT and the cavity mode. As a result,from now <strong>on</strong> the beam radius at these two positi<strong>on</strong>s is about 42 µm.Due to the increased size of the DT beams at the MOT positi<strong>on</strong> the depth of thedipole potential <strong>for</strong> the same YAG laser power is smaller by a factor of wnew/w 2 old 2 ≈ 2.7than was, e.g., in Chapter 2. For atoms trapped in a shallower DT, the optical molassesilluminati<strong>on</strong> is more critical. For instance, a small imbalance in the intensities of thecounter propagating illuminati<strong>on</strong> beams can push atoms out of a shallow trap, while in adeep trap such imbalance in light pressure is outweighted by a large dipole <strong>for</strong>ce holding


3.4 Interplay of the atom trap setup and the cavity 73the atoms. Reducing the intensity of the optical molasses beams helps to keep the atomstrapped. However, it also reduces their fluorescence and thus the signal of the atoms seen<strong>on</strong> a CCD image. To overcome this difficulty we have per<strong>for</strong>med an accurate balancing ofthe intensities of the MOT laser beams used <strong>for</strong> illuminating trapped atoms and increasedthe power of the trapping beams.Note that <strong>for</strong> technical reas<strong>on</strong>s described in Sec. 1.3 we have replaced the Nd:YAG laser(1064 nm) used in experiments of Chapter 2 by an Yb:YAG laser (1030 nm) providing upto 20 W output power in a single mode. Normally, in experiments in Chapter 4 involvingimaging of atoms, we use about 3 W of the DT laser power per beam (in c<strong>on</strong>trast to 1 Wused in Chapter 2). According to equati<strong>on</strong> (1.7), the resulting DT depth is about 1.4 mK.The new beam waist of 36 µm was measured without glass cell by deflecting the beamunder investigati<strong>on</strong> sidewards. Its measured M 2 -parameter is close to 1.0 indicating thatthe beam is close to a perfect Gaussian <strong>on</strong>e [79]. After passing through the class cell andrefocusing the DT beam by similar optics we measure a waist of 41 µm and M 2 -parameterof about 1.2, although the divergence of the beam is the same as measured be<strong>for</strong>e the cell.There<strong>for</strong>e we c<strong>on</strong>clude that the vacuum system introduces additi<strong>on</strong>al aberrati<strong>on</strong>s to thebeam. This can result in decreased DT laser power and increased beam width inside thecavity leading to a dipole trap depth smaller than the 1.4 mK calculated above.Heating the cavity mirrors with the DT beamsWe have observed, that some small fracti<strong>on</strong> of the DT beams is clipped and absorbed bythe cavity mirrors heating them even when the DT passes through the center of the cavity.Besides, due to the small cavity clearance of <strong>on</strong>ly 150 µm, it is difficult to avoid hittingthe cavity and its holder with the powerful YAG beams while aligning the DT <strong>on</strong>to theMOT and then the cavity <strong>on</strong>to the DT. To estimate the possible damages to our cavitycaused in this situati<strong>on</strong>, we have investigated the thermal influence of the YAG beam <strong>on</strong>different cavity elements.Several minutes of exposing a mirror substrate in air to the YAG beams with a powerof 2 W, shined in <strong>on</strong> the substrate from a side, result in heating the mirror up to about40–50 centigrade which is not harmful. However, because of the lower heat c<strong>on</strong>ductanceof mirrors placed in vacuum, the mirror temperature in the glass cell can be higher andis difficult to estimate. There<strong>for</strong>e, we try to avoid l<strong>on</strong>g direct exposure of the mirrorsubstrates to the YAG beams.We have tested that the same YAG laser power focused in air down to about 30 µmdirectly <strong>on</strong> a mirror surface (what is very unlikely to happen in the experiment) does notnoticeably damage the mirror coating. Moreover, the damage threshold of the coating,as specified by the manufacturer, lies about 10 times higher. Furthermore, no damage isdetected if focusing the laser <strong>on</strong>to the glue between mirrors and PZTs.The most vulnerable element of the cavity assembly is found to be the gold-coatedsurface of the shear PZTs. If during beam walking the DT beam slides across the PZTs’surface, the laser immediately burns holes and traces <strong>on</strong> it. Of course, this should bestrictly avoided by carefully per<strong>for</strong>ming the DT prealignment with low power.Although we have seen that the DT beams going through the cavity spacing do not


74 Chapter 3: The high-finesse optical res<strong>on</strong>atordamage the cavity, heating of the locked cavity can significantly shift its res<strong>on</strong>ance frequencyrelative to the probe laser frequency and thus hinders our experiments. Thisproblem and its soluti<strong>on</strong> are discussed in the next Sec. 3.5.Aligning the cavity <strong>on</strong>to the DTDue to the small size of our atom traps and their high sensitivity <strong>on</strong> all instabilities andthermal drifts of optical elements <strong>for</strong>ming them, both the MOT and the DT beams shouldbe accurately aligned at least <strong>on</strong>ce a day be<strong>for</strong>e per<strong>for</strong>ming any experiment with atoms.The presence of the cavity inside the glass cell complicates the alignment procedure ofthe DT beams. To avoid str<strong>on</strong>g heating of the cavity and damaging the PZTs with theYAG laser beams while aligning them <strong>on</strong>to the MOT, their transmissi<strong>on</strong> through thecavity spacing is permanently observed. If we detect clipping of the beams, the cavityholder is displaced appropriately to let them pass between the cavity mirrors and the DTalignment is c<strong>on</strong>tinued. Besides, an infrared CCD camera watching the cavity from theside m<strong>on</strong>itors the YAG scattering off the cavity mirrors, thus giving us good qualitativein<strong>for</strong>mati<strong>on</strong> about the positi<strong>on</strong> of the cavity with respect to the DT beams.After the DT is aligned <strong>on</strong>to the MOT and its positi<strong>on</strong> is fixed, we start to alignthe cavity relative to the DT. First, we set the right vertical positi<strong>on</strong> of the cavity bymeasuring the transmissi<strong>on</strong> of <strong>on</strong>e of the DT beams through the cell while moving thecavity up and down. For large cavity displacements in both directi<strong>on</strong>s the transmissi<strong>on</strong>starts to decrease due to clipping of the beams. The optimal positi<strong>on</strong> is in the middlebetween these two regi<strong>on</strong>s. The estimated precisi<strong>on</strong> of this method is about 20 µm.The alignment of the cavity in the horiz<strong>on</strong>tal plane transversely to the DT beams ismore demanding, since the DT axis and the cavity mode should overlap precisely. Thisalignement is per<strong>for</strong>med by using transported atoms as a probe <strong>for</strong> measuring the positi<strong>on</strong>of the cavity mode with micrometer precisi<strong>on</strong>. Several alternative experimental procedures<strong>for</strong> locating the mode are presented and discussed in Sec. 4.3.Local vacuum between the mirrorsIn general, the local vacuum between the two closely spaced cavity mirrors can beworse than the vacuum in the rest of the glass cell [93] due to outgassing of the mirrorsubstrates. In order to make sure that the local vacuum here does not affect the trappedatoms, we measure the lifetime of atoms in the DT transported between the cavitymirrors. For this purpose, we move several (<strong>on</strong> average 5) atoms into the cavity withthe probe and lock lasers switched off. After a variable waiting time we transport theremaining atoms back and count them by transferring the atoms back into the MOT.The experimental sequence is repeated 30 times. As a reference, we per<strong>for</strong>m the samelifetime measurement without transporting atoms. The result is shown in Fig. 3.16. Inboth measurements the measured lifetime is about 10 s, mainly limited by phase noise ofthe AOMs used <strong>for</strong> shifting the standing-wave trap, see also Sec. 1.3. C<strong>on</strong>sequently, thelocal vacuum between the cavity mirrors should not c<strong>on</strong>strain our experiments.


3.5 Compensati<strong>on</strong> of thermal drifts of the locked cavity 75100Survival probability [%]80604020without transportbetween the mirrors00.01 0.1 1 10 100Time in the cavity [s]Figure 3.16: Lifetime of trapped atoms placed between the cavity mirrors (circles).For comparis<strong>on</strong> the atom survival is also measured without transporting atoms into thecavity (squares).3.5 Compensati<strong>on</strong> of thermal drifts of the locked cavityThe laser beams of the dipole trap, while passing between the mirrors, are partially clippedand absorbed by the mirror edges. Since the index of refracti<strong>on</strong> of the mirror coating isin general temperature dependent, the resulting heating of the coating changes the phaseshift of the cavity mode light penetrating into it. If the cavity is locked, this phase shift iscompensated by changing the cavity length, thus keeping the cavity always res<strong>on</strong>ant withthe lock laser.Due to the difference in wavelength, the probe and lock lasers have different penetrati<strong>on</strong>depths into the mirror dielectric coating (the lock laser penetrates deeper) and thus their“thermal” phase shifts are different. As l<strong>on</strong>g as the mirror temperature stays c<strong>on</strong>stant,the c<strong>on</strong>stant differential phase shift can be corrected by slightly changing the frequencyof the lock laser in order to bring the cavity in res<strong>on</strong>ance with the probe laser.However, during any experimental run c<strong>on</strong>sisting of many repetiti<strong>on</strong>s of some experimentalsequence we repeatedly switch the DT beams <strong>on</strong> and off (e.g., <strong>for</strong> loading andcounting atoms in the MOT). The rapidly alternating heating and cooling of the mirrorsdetermine a new thermal equilibrium <strong>for</strong> the cavity which is approached after the experimenthas been started. This equilibrium depends <strong>on</strong> the average DT laser power affectingthe cavity during <strong>on</strong>e repetiti<strong>on</strong> of the experiment and is thus different from that be<strong>for</strong>ethe experiment was started. As a result, the cavity, while remaining stabilized to the locklaser, drifts away from the probe frequency. Hence, the probe laser transmissi<strong>on</strong> throughthe cavity drops during the successive executi<strong>on</strong> of several experimental shots, as can beclearly seen in Fig. 3.17. In this measurement the durati<strong>on</strong> of <strong>on</strong>e shot is about 1 s with


76 Chapter 3: The high-finesse optical res<strong>on</strong>ator18Probe laser transmissi<strong>on</strong> [mV]161412108642no compensati<strong>on</strong>with compensati<strong>on</strong>00 2 4 6 8 10 12 14 16 18 20Shot numberFigure 3.17: Probe transmissi<strong>on</strong> with (diam<strong>on</strong>ds) and without (circles) compensati<strong>on</strong>of the cavity thermal drifts. The transmissi<strong>on</strong> is measured <strong>for</strong> 20 successive experimentalshots within <strong>on</strong>e experimental run. The compensati<strong>on</strong> method used is the center-of-massmethod. A slight increase of the cavity transmissi<strong>on</strong> originates from a drift of the incidentpower of the probe laser.the DT switched <strong>on</strong> about 60 % of the time. Be<strong>for</strong>e the experiment has been started, theDT laser beams were <strong>on</strong> c<strong>on</strong>tinuously and the cavity was set in res<strong>on</strong>ance to the probelaser. One sees that already after a few experimental shots the transmissi<strong>on</strong> is significantlydecreased, making further executi<strong>on</strong> of the experiment useless.The differential thermal phase shift between the lock and the probe laser can be decreasedby reducing the amount of the absorbed DT laser power. However, in our experimentsthis cannot be achieved by simple means. For instance, decreasing the DTlaser power, what is the first natural candidate in reducing the cavity heating, will notallow us to take images of the trapped atoms without loosing them, see Sec. 3.4. Next,increasing the spacing between cavity mirrors will result in an undesirable decrease of theatom-phot<strong>on</strong> coupling, see Sec. 3.1. Finally, decreasing the diameter of the DT beams willinevitably reduce their Rayleigh range, complicating the transport of atoms over severalmillimeter distances. There<strong>for</strong>e, having no means to reduce the cavity heating, we cantake it into account and try to actively compensate it.Active compensati<strong>on</strong> of the differential thermal phase shiftTo actively compensate <strong>for</strong> the described thermal drifts of the locked cavity, we adjust thefrequency of the lock laser during the experimental run in such a way that the cavity, beingstabilized to the lock laser, stays res<strong>on</strong>ant with the probe laser. The lock laser frequencyis tuned by means of the AOM 2 shown in Fig. 3.12. The frequency of the AOM driver is


3.5 Compensati<strong>on</strong> of thermal drifts of the locked cavity 77(a)A B C D E FAOM c<strong>on</strong>trol voltageU 0+ UU 0U0-U(b)2.7732.7722.771APD signal [V]2.7702.7692.7682.7672.7662.7650 10 20 30 40 50 60 70 80 90 100 110 120 130Time [ms]Figure 3.18: Compensating <strong>for</strong> the cavity thermal drifts resulting from switching the DTbeams <strong>on</strong> and off. The time interval A is reserved <strong>for</strong> per<strong>for</strong>ming cavity <strong>QED</strong> experimentsand can have an arbitrary length depending <strong>on</strong> the planned experiments. The intervalsB to F are reserved <strong>for</strong> determining the optimal c<strong>on</strong>trol voltage of the lock laser AOMwhich provides the maximum transmissi<strong>on</strong> of the probe laser. (a) Time dependence ofthe c<strong>on</strong>trol voltage sent to the lock laser AOM. The initial optimal voltage U 0 is foundfrom the previous experimental shot. (b) The corresp<strong>on</strong>ding cavity transmissi<strong>on</strong> of theprobe laser as detected by the APD.c<strong>on</strong>trolled via its VCO input by the PC analog output card.The compensati<strong>on</strong> procedure works as follows. After the cavity <strong>QED</strong> measurementhas been per<strong>for</strong>med and the atoms are moved out of the cavity mode we scan the cavityres<strong>on</strong>ance over the probe laser frequency in a way depending <strong>on</strong> the specific compensati<strong>on</strong>method, see below, and measure the cavity transmissi<strong>on</strong> of the probe laser. Then, wedetermine the AOM c<strong>on</strong>trol voltage U max , which corresp<strong>on</strong>ds to the maximum transmissi<strong>on</strong>of the probe laser, and use its value in the next experimental shot.


78 Chapter 3: The high-finesse optical res<strong>on</strong>atorCenter-of-mass methodThere are several algorithms to determine U max . The easiest <strong>on</strong>e is to linearly scan overthe whole transmissi<strong>on</strong> spectrum of the probe laser and to find the center of mass of thetransmissi<strong>on</strong> peak. The realizati<strong>on</strong> of this method is illustrated in Fig. 3.18. Part (a)shows a time sweep of the AOM c<strong>on</strong>trol voltage prepared <strong>for</strong> scanning the cavity anddetermining U max . Part (b) presents a typical cavity transmissi<strong>on</strong> signal corresp<strong>on</strong>ding tothe cavity scan of (a). In this example the transmissi<strong>on</strong> is detected by an analog APD.However, the compensati<strong>on</strong> methods described in this chapter work as well with a singlephot<strong>on</strong> counting detector.The time interval A is reserved <strong>for</strong> the cavity <strong>QED</strong> part of the experiment and itslength can be set arbitrary. The AOM voltage U 0 , which is supposed to corresp<strong>on</strong>d tothe maximum transmissi<strong>on</strong> T max of the probe laser, equals U max , which was calculatedin the previous shot. During the intervals B to F we determine a new optimal voltageU max which will be automatically used <strong>for</strong> the next experimental shot. The probe lasertransmissi<strong>on</strong> line is recorded in the time interval D. During the time intervals C and Ewe measure the background level.After the whole voltage sweep A–F is executed and the corresp<strong>on</strong>ding cavity transmissi<strong>on</strong>is recorded, the cavity c<strong>on</strong>trol software calculates the positi<strong>on</strong> t c of the centerof mass of the background-free cavity transmissi<strong>on</strong> and the corresp<strong>on</strong>ding AOM voltageU max . Then, it reprograms the PC analog output card with a new AOM voltage sweepsimilar to Fig. 3.18(a) with U 0 given by the newly found U max .Two-point methodOur sec<strong>on</strong>d method of finding the center positi<strong>on</strong> of the transmissi<strong>on</strong> peak, which hasa Lorentzian <strong>for</strong>m, is based <strong>on</strong> precisely measuring the transmissi<strong>on</strong> level at two pointslocated <strong>on</strong> the opposite slopes of the transmissi<strong>on</strong> peak. Using these two values togetherwith the maximum transmissi<strong>on</strong> and the background level, which are in general c<strong>on</strong>stantand thus are measured independently be<strong>for</strong>e the main experiment, we then determine thepositi<strong>on</strong> of the peak maximum. To realize this method, we apply an AOM voltage scansimilar to that of Fig. 3.18(a). The two transmissi<strong>on</strong> values are measured during thetime intervals C and E, respectively. For this purpose, we use a small voltage span ∆V ,corresp<strong>on</strong>ding to a half width of the transmissi<strong>on</strong> peak.ResultsThe efficiency of the center-of-mass method (CMM) <strong>for</strong> compensating cavity thermal driftsis dem<strong>on</strong>strated in Fig. 3.17. We are able to keep the cavity transmissi<strong>on</strong> high in eachexperimental shot, making our experiments insensitive to the thermal drifts of the cavity.The cavity transmissi<strong>on</strong> compensated with the two-point method (TPM) looks very similar(within statistical errors) to that of CMM and thus is not shown here.Although in this dem<strong>on</strong>strati<strong>on</strong> the results of the two compensati<strong>on</strong> methods looksimilar, the methods differ significantly by their capture range ∆U. It defines the maximalthermal drift between two successive experimental shots, which can be compensated. In


3.5 Compensati<strong>on</strong> of thermal drifts of the locked cavity 797.740.0Optimal AOM voltage U max[V]7.727.707.687.66-0.2-0.4-0.6-0.8Frequency shift [MHz]7.64-1.00 20 40 60 80 100Shot numberFigure 3.19: Re-thermalizati<strong>on</strong> of the locked cavity during an experimental run. Oneshot lasts 1.06 s. The straight line is an exp<strong>on</strong>ential fit with a decay time of 27.6(±0.6) s.c<strong>on</strong>trast to TPM having a small ∆U, the capture range of CMM can be set arbitrarylarge, depending <strong>on</strong> the experimental needs. However, being more sensitive to smalltransmissi<strong>on</strong> shifts, TPM gives a more precise peak positi<strong>on</strong> compared to CMM <strong>for</strong> thesmall thermal drifts, since it measures the signal at the most sensitive points (the slopes)of the transmissi<strong>on</strong> peak.All experiments described in Sec. 4.5 have been per<strong>for</strong>med with the robust center-ofmassmethod, although in the future, after improving the stability of the experimentalsetup and passively reducing the thermal drifts, see Outlook, we plan to use the moreprecise two-point method.Re-thermalizati<strong>on</strong> of the cavityThe cavity re-thermalizati<strong>on</strong> during successive executi<strong>on</strong>s of many experimental shots canbe explored by recording values of U max determined in individual shots. Since the cavityis actively kept res<strong>on</strong>ant with the probe laser, the change of U max corresp<strong>on</strong>ds to thedifferential thermal phase shift in the cavity between the lock and probe laser. Figure 3.19presents the re-thermalizati<strong>on</strong> in an experiment c<strong>on</strong>sisting of 100 shots. The graph showsthe optimal AOM voltage U max found in each experimental shot. The right vertical axisshows the corresp<strong>on</strong>ding frequency shift applied to the lock laser <strong>for</strong> compensating thethermal changes in the mirror coating. The durati<strong>on</strong> of <strong>on</strong>e experimental sequence togetherwith the time interval between two successive shots is 1.060 s. The overall run durati<strong>on</strong>is thus 106 s. The DT beams are switched <strong>on</strong> during 59 % of this time. The c<strong>on</strong>tinuousline is an exp<strong>on</strong>ential fit with a decay time of 27.6(±0.6) s. One sees that already after 20


80 Chapter 3: The high-finesse optical res<strong>on</strong>atorshots the thermal drift of the locked cavity relative to the probe laser frequency is about0.5 MHz, which is about a half-width of the cavity res<strong>on</strong>ance line. This result is c<strong>on</strong>sistentwith the decrease of the cavity transmissi<strong>on</strong> observed in Fig. 3.17 with no compensati<strong>on</strong>applied.Knowing the frequency shift, ν shift , the differential phase shift can be found fromthe same <strong>for</strong>mula as used <strong>for</strong> calculating the frequency splitting of birefringent modesdepending <strong>on</strong> a round trip phase shift, see equati<strong>on</strong> (3.14). This results in the followingrelati<strong>on</strong>δ therm = 2πν shift2Lc . (3.30)Thus, the frequency shift of, e.g., ν shift = 1 MHz results from the phase shift between theprobe and lock laser wavelengths in the cavity of δ therm = 0.65 × 10 −5 rad.C<strong>on</strong>clusi<strong>on</strong> and discussi<strong>on</strong>All experiments described in Sec. 4.5 have been per<strong>for</strong>med with the center-of-mass stabilizati<strong>on</strong>method. In this way the probe laser transmissi<strong>on</strong> was kept maximal even inexperiments with almost 10 sec<strong>on</strong>ds l<strong>on</strong>g shots, see e.g. measurements of Fig. 4.19. Moreover,both methods have been successfully applied in experiments with very low probelaser power having a signal-to-noise (S/N) ratio of about 1.5. For comparis<strong>on</strong>, S/N inFig. 3.18(b) is about 6.5. If experiments require an even smaller probe laser power, thedurati<strong>on</strong>s of the intervals C, D, and E can be increased in order to overcome the lowerS/N. Alternatively, we can increase the probe laser power <strong>on</strong>ly during the compensati<strong>on</strong>phase B–F, when the atoms are not affected by the cavity field anymore.Although both stabilizati<strong>on</strong> schemes serve reliably, we still try to keep thermal driftsof the cavity as small as possible by leaving the DT beams switched <strong>on</strong> all the time, except<strong>for</strong> loading the MOT with atoms and collecting the MOT fluorescence. Furthermore, thebeams are kept <strong>on</strong> even between executi<strong>on</strong> of different experimental runs and betweendifferent experimental shots within individual runs. This helps to reduce temperaturechanges of the cavity mirrors.In some experiments we might require the cavity not to be res<strong>on</strong>ant with the probelaser, but instead to have a well-defined detuning relative to it. Still, we can use thedescribed stabilizati<strong>on</strong> methods in order to find the exact positi<strong>on</strong> of the maximum of thelaser transmissi<strong>on</strong> and <strong>on</strong>ly then shift the cavity by the required frequency relative to theprobe laser by means of the lock laser AOM.Summarizing, the developed stabilizati<strong>on</strong> schemes allow us to precisely c<strong>on</strong>trol thecavity res<strong>on</strong>ance frequency relative to the frequency of the probe laser in the presence ofthermal heating and cooling of the cavity mirrors due to switching the DT laser beams<strong>on</strong> and off. This thermal changes result in different thermal expansi<strong>on</strong> of the cavity seenby the lock laser and the probe laser. In this case, the cavity res<strong>on</strong>ance, stabilized bythe lock laser, moves away from the probe laser frequency due to thermal expansi<strong>on</strong> orc<strong>on</strong>tracti<strong>on</strong>.


Chapter 4Deterministic atom-cavitycoupling4.1 Introducti<strong>on</strong>For several years all experimental progress in our research group has been achieved with theobjective of realizing deterministic atom-cavity coupling. The developed optical c<strong>on</strong>veyorbelt allows us to transport a single atom or any small number of atoms over macroscopicdistances [23, 24]. Besides, this transport has been dem<strong>on</strong>strated to preserve the coherenceof the prepared atomic quantum state [19]. The realizati<strong>on</strong> of a quantum register <strong>on</strong> astring of trapped atoms [18] gives us the possibility to prepare and manipulate quantumstates of individual atoms. Thus, in combinati<strong>on</strong> with the coherent transport, this shouldallow us to let atoms coherently interact at a locati<strong>on</strong> different from the state preparati<strong>on</strong>and read-out, e.g., inside a remote high-finesse cavity.The first versi<strong>on</strong> of the required cavity setup was developed in our group about 5 yearsago [80]. During my thesis the assembly and test of the improved high-finesse opticalres<strong>on</strong>ator have been per<strong>for</strong>med. Besides, several special techniques, such as the numberlockedloading of the dipole trap [56] and the precise positi<strong>on</strong> c<strong>on</strong>trol of single atoms [25],necessary <strong>for</strong> the deterministic manipulati<strong>on</strong> of single atoms, have been developed. S<strong>on</strong>ow, after coming a l<strong>on</strong>g way, we are ready to tackle the l<strong>on</strong>g-expected goal of our group– the deterministic atom-cavity coupling.This chapter starts with the basic theoretical backgrounds necessary <strong>for</strong> understandingthe main properties of a coupled atom-cavity system (Sec. 4.2). Next, I describe severalmethods <strong>for</strong> precisely locating the cavity mode, which is necessary <strong>for</strong> placing atoms at thefield maximum and thus <strong>for</strong> achieving the highest possible atom-cavity coupling (Sec. 4.3).In Sec. 4.4 I present the transportati<strong>on</strong> of bunches of several atoms into the cavity andtheir coupling to the cavity mode. These experiments with many atoms helped us to betterunderstand our new system as well as to test the compatibility of different comp<strong>on</strong>entsand techniques. Finally, the main result of my thesis – the deterministic transport andcoupling of single atoms to the mode of a high-finesse optical res<strong>on</strong>ator – is presented inSec. 4.5.81


82 Chapter 4: Deterministic atom-cavity coupling4.2 Basics of cavity quantum electrodynamicsIn this secti<strong>on</strong> I recall a basic theoretical approach how to treat a coupled atom-cavitysystem. It is based <strong>on</strong> the so-called Jaynes-Cummings model of the spin-spring interacti<strong>on</strong>– coupling of a two-level system (atom) with a quantum harm<strong>on</strong>ic oscillator (cavitymode) [2]. The Jaynes-Cummings Hamilt<strong>on</strong>ian gives us the eigenstates and eigenvaluesof the coupled system depending <strong>on</strong> the coupling strength and the respective frequenciesof the system comp<strong>on</strong>ents. The c<strong>on</strong>tributi<strong>on</strong> of dissipati<strong>on</strong> present in any open physicallyrealizable system is treated by using the density matrix <strong>for</strong>malism and solving the masterequati<strong>on</strong> [94]. This approach allows us to find the expectati<strong>on</strong> values of the relevant operatorsdescribing the system as well as to calculate the modified spectrum in the presenceof the coupling. A more detailed introducti<strong>on</strong> into the cavity quantum electrodynamicssystem can be found, e.g., in Refs. [78, 95]. C<strong>on</strong>cerning atom-phot<strong>on</strong> interacti<strong>on</strong> I refer to[96].4.2.1 Atom-cavity coupling rateThe interacti<strong>on</strong> between atom and cavity is based <strong>on</strong> the electric dipole interacti<strong>on</strong> ofthe atomic dipole moment, d, with the electric field of the cavity mode, E(⃗r). Thecorresp<strong>on</strong>ding coherent coupling rate g can be introduced asg(⃗r) = d · E(⃗r). (4.1)Due to the spatial structure of the cavity mode field, ψ(⃗r), the atom-cavity coupling rateis positi<strong>on</strong> dependent and can be expressed asg(⃗r) = g 0 ψ(⃗r), (4.2)where g 0 is the coupling rate at the maximum of the mode. C<strong>on</strong>sidering the fundamentalTEM 00 mode of the cavity and neglecting the divergence of the Gaussian mode <strong>for</strong> theshort cavity, we getψ(⃗r) = exp[− x2 + y 2 ] ( ) 2πzsin . (4.3)λw 2 0Here, the oscillatory term indicates the standing-wave structure of the cavity mode witha wavelength λ al<strong>on</strong>g its axis (z directi<strong>on</strong>) and the Gaussian term is the transversal modeprofile of the cavity fundamental mode with a waist ω 0 (x and y directi<strong>on</strong>s).The r.m.s. electric field amplitude of the vacuum in the cavity mode of a volume Vand of an angular frequency ω c is given by [97]√ωcE =(4.4)2ɛ 0 Vwith ɛ 0 denoting the permittivity of free space. Thus, using (4.1), the maximum couplingrate reads√ωcg 0 = d2 ɛ 0 V . (4.5)


4.2 Basics of cavity quantum electrodynamics 83.........atomcavityuncoupled coupledatom-cavity system(a) (b) (c) (d)Figure 4.1: Eigenstates of the res<strong>on</strong>ant atom-cavity system. For the res<strong>on</strong>ant system,the frequency spacing ω coincides with ω a and ω c .4.2.2 Jaynes-Cummings modelThe Hamilt<strong>on</strong>ian of a coupled atom-cavity system is generally expressed as a sum of threeterms – atomic, cavity, and interacti<strong>on</strong>:H = H atom + H cavity + H int . (4.6)Let us c<strong>on</strong>sider a two-level atom with the ground and excited levels |g〉 and |e〉, respectively.The levels are separated by the energy difference ω a with ω a the angular res<strong>on</strong>ancefrequency of the atomic transiti<strong>on</strong> as shown in Fig. 4.1(a). The Hamilt<strong>on</strong>ian <strong>for</strong> the atomthen readsH atom = ∑ E i |i〉〈i| (4.7)iwith E i the energy of the atomic state |i〉. For c<strong>on</strong>venience, we set the energy E g of theground state to zero.In the following we introduce the atomic raising and lowering operators, σ † and σ,defined as σ † = |e〉〈g| and σ = |g〉〈e|. They describe the excitati<strong>on</strong> and deexcitati<strong>on</strong> ofthe atom, respectively, and their product σ † σ = |e〉〈e| gives the projecti<strong>on</strong> operator <strong>on</strong>tothe atomic excited state. The atomic Hamilt<strong>on</strong>ian can now be expressed asH atom = ω a σ † σ . (4.8)


84 Chapter 4: Deterministic atom-cavity couplingThe Hamilt<strong>on</strong>ian <strong>for</strong> the cavity mode with a frequency ω c is given by(H cavity = ω c a † a + 1 ), (4.9)2where a † and a are the field creati<strong>on</strong> and annihilati<strong>on</strong> operators, respectively. Their productcorresp<strong>on</strong>ds to the phot<strong>on</strong> number operator, ̂N, defining the number of the excitati<strong>on</strong>sin the cavity. The energy spectrum of the cavity c<strong>on</strong>sists of equidistant phot<strong>on</strong> numberstates |i〉 separated by the cavity res<strong>on</strong>ance frequency ω c , see Fig. 4.1(b). The groundstate |0〉 is referred to as the vacuum state.The interacti<strong>on</strong> Hamilt<strong>on</strong>ian based <strong>on</strong> the electric dipole coupling between the atomand the cavity readsH int = ̂d · Ê (4.10)= d(σ † e iωat + σ e −iωat ) · E(a † e iωct + a e −iωct ). (4.11)Here, the dipole moment operator, ̂d, and the electric field operator, Ê, are expressed viathe atomic and cavity operators in the Heisenberg interacti<strong>on</strong> picture. In the approximati<strong>on</strong>|ω a − ω c | ≪ (ω a + ω c ) the Hamilt<strong>on</strong>ian (4.11) reduces toH int = g(σ † a + σa † ) (4.12)with g the coherent atom-field coupling rate, introduced above.The total Hamilt<strong>on</strong>ian of the atom-cavity system now reads(H = ω a σ † σ + ω c a † a + 1 )+ g(σ † a + σa † ) (4.13)2and corresp<strong>on</strong>ds to the Jaynes-Cummings model [2] <strong>for</strong> a single two-level atom in anelectromagnetic field in the absence of dissipati<strong>on</strong>.Without interacti<strong>on</strong>, e.g., <strong>for</strong> g = 0, the levels |g〉|n〉 and |e〉|n−1〉 of the res<strong>on</strong>antuncoupled atom-cavity system are degenerate, having energy nω with ω = ω a = ω c , seeFig. 4.1(c). The coupling lifts this degeneracy and the new eigenstates of the coupledsystem are found by diag<strong>on</strong>alizing the Hamilt<strong>on</strong>ian (4.13), yielding|± n 〉 = (|g〉|n〉 ± |e〉|n−1〉)/ √ 2 , (4.14)where n is the number of excitati<strong>on</strong>s in the system. The corresp<strong>on</strong>ding energy eigenvaluesreadE ±n = nω ± √ n g (4.15)and their spectrum is schematically shown in Fig. 4.1(d). The eigenstates with n<strong>on</strong>-zeroenergy are symmetrically split by 2 √ n g. The frequency splitting of the first excited stateof 2g is called vacuum Rabi splitting. This frequency is also known as the single-phot<strong>on</strong>Rabi frequency since it determines the energy exchange rate between the atomic dipoleand the cavity electric field c<strong>on</strong>taining <strong>on</strong>e phot<strong>on</strong>. It is worthwhile menti<strong>on</strong>ing that if Nidentical atoms are simultaneously coupled to the cavity mode, the vacuum Rabi splittingis increased by a factor of √ N and reads [78]Ω Rabi = 2 √ N g . (4.16)


4.2 Basics of cavity quantum electrodynamics 854.2.3 Density matrix and master equati<strong>on</strong>The Jaynes-Cummings Hamilt<strong>on</strong>ian describes a system in the absence of dissipati<strong>on</strong> anddriving, i.e. a closed system isolated from the envir<strong>on</strong>ment. However, our system is subjectto dissipati<strong>on</strong>: The atom sp<strong>on</strong>taneously decays into modes other than the cavity <strong>on</strong>e witha dipole decay rate γ and the cavity phot<strong>on</strong> leaks out of the cavity mode with a leakagerate κ due to finite transmissi<strong>on</strong>, absorpti<strong>on</strong>, and/or scattering <strong>on</strong> cavity mirrors. Inadditi<strong>on</strong>, the coupled system can be probed by an external classical driving (laser) field.Thus, our experimental atom-cavity coupling can not be completely described by theJaynes-Cummings Hamilt<strong>on</strong>ian (4.13) al<strong>on</strong>e.In the presence of dissipati<strong>on</strong> the atom-cavity system cannot be treated independentlyfrom the envir<strong>on</strong>ment. The joint evoluti<strong>on</strong> of the atom-cavity-envir<strong>on</strong>ment system canbe treated using the Schrödinger equati<strong>on</strong> with the joint Hamilt<strong>on</strong>ian. However, since weare not interested in the envir<strong>on</strong>mental states, which are in general not known, we cantrace over them if using the density matrix <strong>for</strong>malism. The time evoluti<strong>on</strong> of the openatom-cavity system is then described by the master equati<strong>on</strong> [94, 98]dρdt= Lρ (4.17)with the atom-cavity density operator ρ and the Liouvillian L, which is defined asLρ = − i [H, ρ ] + Ĉρ Ĉ† − 1 2 Ĉ† Ĉρ − 1 2 ρ Ĉ† Ĉ . (4.18)The Jaynes-Cummings Hamilt<strong>on</strong>ian H now includes coherent driving of the cavity bya probe laser field ε(t) = ε 0 e iωpt and can be expressed in a frame rotating at the probefrequency ω p asH = (ω a − ω p )σ † σ + (ω c − ω p )a † a + g(a † σ + a σ † ) + ε 0 (a + a † ). (4.19)Dissipati<strong>on</strong> is <strong>for</strong>mally introduced in (4.18) by the collapse operator Ĉ in the <strong>for</strong>m [94, 99]Ĉ = √ 2γ σ + √ 2κ a . (4.20)The two terms corresp<strong>on</strong>d to sp<strong>on</strong>taneous emissi<strong>on</strong> from the atom and leakage from thecavity, respectively.The master equati<strong>on</strong> provides a valid descripti<strong>on</strong> of the atom-cavity system in anyrange of parameters (g 0 , κ, γ). In some limits it can be solved analytically, see e.g. [81] <strong>for</strong>the case of res<strong>on</strong>ance between the atom and the cavity driven by a weak external field.In general we solve the master equati<strong>on</strong> numerically by seeking <strong>for</strong> a steady-state soluti<strong>on</strong><strong>for</strong> ρ such that Lρ = 0. Since the expectati<strong>on</strong> value of any operator Ô can be found as〈Ô〉 = Tr(ρ Ô), (4.21)the density matrix provides us the complete in<strong>for</strong>mati<strong>on</strong> <strong>on</strong> the system. For instance,the intra-cavity field intensity is given by the expectati<strong>on</strong> value of the phot<strong>on</strong> numberoperator, 〈 ̂N〉, and thus the cavity output rate is 2κ〈 ̂N〉 = 2κ Tr(ρ a † a).


86 Chapter 4: Deterministic atom-cavity couplingFigure 4.2: Vacuum Rabi splitting from a steady-state soluti<strong>on</strong> of the master equati<strong>on</strong><strong>for</strong> our experimental dissipati<strong>on</strong> rates (κ, γ) = 2π × (0.43, 2.6) MHz. The system isprobed with a weak external laser with a drive strength of n = 10 −3 phot<strong>on</strong>s (i.e., theprobe laser fills the res<strong>on</strong>ant cavity mode with <strong>on</strong> average 10 −3 phot<strong>on</strong>s).4.2.4 Vacuum Rabi splittingBe<strong>for</strong>e per<strong>for</strong>ming a numerical analysis of the master equati<strong>on</strong>, we have to choose appropriatevalues of the main experimental parameters to use, namely the parameter set(g 0 , κ, γ). Given by equati<strong>on</strong> (4.5), the coupling rate depends <strong>on</strong> the atomic dipole momentand the mode volume. We c<strong>on</strong>sider the atomic cycling transiti<strong>on</strong> |F = 4, m F =±4〉 → |F ′ = 5, m ′ F= ±5〉 in cesium. Its electric dipole moment readsd =√3 ɛ0 λ 3 Γ8π 3 = 3.167 ea 0 (4.22)with λ and Γ denoting the wavelength and the natural line width of this transiti<strong>on</strong>, respectively.The cavity geometry gives us the mode volume of V = 6.6 ×10 4 µm 3 , see Sec. 3.3.4,resulting in the coupling g 0 /2π = 18 MHz.According to equati<strong>on</strong> (3.12), the decay rate of the cavity field is determined by thecavity finesse and <strong>for</strong> F = 1.12 × 10 6 equals κ = 2π × 0.43 MHz. For a cesium atomγ = 2π × 2.6 MHz. C<strong>on</strong>sequently, the expected parameters of our atom-cavity system are(g 0 , κ, γ) = 2π × (18, 0.43, 2.6) MHz . (4.23)


4.2 Basics of cavity quantum electrodynamics 8760Probe laser detuning ∆ P/2π [MHz]40200-20-40-60-60 -40 -20 0 20 40 60<strong>Cavity</strong> detuning ∆ C/2π [MHz]Figure 4.3: Energy levels of the coupled atom-cavity system <strong>for</strong> (g 0 , κ, γ) = 2π ×(18, 0.43, 2.6) MHz. The grey color gradient from white to black corresp<strong>on</strong>ds to thecavity transmissi<strong>on</strong> from 0 to 1. Two straight lines indicate the positi<strong>on</strong> of the uncoupledenergy states.Figure 4.2 shows the coupled atom-cavity transmissi<strong>on</strong> spectrum found via a steadystatenumerical soluti<strong>on</strong> of the master equati<strong>on</strong> (4.17) <strong>for</strong> the parameters (4.23). In thiscalculati<strong>on</strong> the cavity is res<strong>on</strong>ant with the atom and is driven by a weak probe laser.The transmissi<strong>on</strong> is calculated as a functi<strong>on</strong> of the coupling rate g and of the probe laserdetuning from the atomic res<strong>on</strong>ance ∆ p = ω p − ω a .Without interacti<strong>on</strong> between the atom and the cavity (g = 0), the transmissi<strong>on</strong> hasa Lorenzian line profile of an empty cavity with a full width at half maximum of 2κ. Asthe coupling increases the peak splits into two lines reflecting the vacuum Rabi splitting(or normal-mode splitting). As predicted by equati<strong>on</strong> (4.15), the splitting equals twicethe coupling rate, i.e. ∆ split = 2g. The line width of the Rabi sidebands is given by twicethe sum of all dissipati<strong>on</strong> rates in the system, i.e. ∆ FWHM = 2(κ + γ) [81]. The reducti<strong>on</strong>of the height of the Rabi sidebands, T Rabi , relative to the height of the Lorentzian peak,T Lorentz , is given by( )T Rabi κ 2= . (4.24)T Lorentz κ + γThis relati<strong>on</strong> is valid in the str<strong>on</strong>g coupling regime when g ≫ (κ, γ) and does not depend <strong>on</strong>g. In our case of (κ, γ) = 2π×(0.43, 2.6) MHz, the reduced height of the sidebands is 2.0 %of the maximal. Note that a dependence similar to (4.24) has been observed in classicaltransmissi<strong>on</strong> of light through a Fabry-Perot res<strong>on</strong>ator. According to equati<strong>on</strong> (3.23b), the


88 Chapter 4: Deterministic atom-cavity couplingpresence of mirror absorpti<strong>on</strong> and scattering losses A reduces the cavity transmissi<strong>on</strong> bya factor of T 2 /(T + A) 2 . In the coupled atom-cavity system the atom with a decay rate γplays a role of a phot<strong>on</strong> scatterer, similar to A, leading to the same transmissi<strong>on</strong> law.As seen from Fig. 4.2, the Rabi splitting cannot be observed if the coupling rate gis smaller than the system’s linewidth, i.e. if g ≪ (κ + γ). In other words, in the weakcouplingregime the dissipati<strong>on</strong> processes, represented by κ and γ, overwhelm the atomfieldinteracti<strong>on</strong>, g. In c<strong>on</strong>trast, if g ≫ (κ, γ), we get into the str<strong>on</strong>g-coupling regime whenthe atom-cavity system has time to couple coherently be<strong>for</strong>e the energy in the systemdissipates. This regime is of most interest <strong>for</strong> us since it allows us to explore coherentevoluti<strong>on</strong> in the atom-cavity system. The quality of atom-cavity system is often measuredby the cooperativity parameter, defined asC N = Ng22κγ(4.25)<strong>for</strong> N atoms simultaneously coupled to the cavity mode. It compares the atom-cavitycoupling rate g with the cavity phot<strong>on</strong> decay rate κ and the atom decay rate γ. Thus, inthe str<strong>on</strong>g coupling regime <strong>for</strong> <strong>on</strong>e atom in the cavity C 1 ≫ 1.The transmissi<strong>on</strong> of the probe laser through the cavity detuned with respect to theatomic res<strong>on</strong>ance is shown in Fig. 4.3. The two dark stripes of high cavity transmissi<strong>on</strong>represent the behavior of the energy states of the coupled system. The straight linesindicate the eigenstates of the uncoupled atom-cavity system: the horiz<strong>on</strong>tal and diag<strong>on</strong>allines show the atom and the cavity res<strong>on</strong>ances, respectively. In the res<strong>on</strong>ance case of∆ c = 0, the spectrum dem<strong>on</strong>strates the vacuum Rabi splitting discussed above. The n<strong>on</strong>res<strong>on</strong>antatom-cavity coupling (i.e. in the so-called dispersive regime) of |∆ c | = |ω c −ω a | ≫g results in a shift of the cavity res<strong>on</strong>ance with respect to the empty-cavity by [78]4.2.5 Atom oscillati<strong>on</strong>s in the dipole trap∆ shift = g2∆ c. (4.26)Due to the finite atomic temperature an atom trapped in a potential well oscillates al<strong>on</strong>gits axial as well as two radial directi<strong>on</strong>s. Since the cavity field has a complex spatialstructure, given by equati<strong>on</strong> (4.3), an atom, which is placed into the cavity mode andsimultaneously oscillates in the DT, experiences a time-dependent atom-cavity couplingrate, which is smaller than the maximum coupling g 0 given by (4.5). The axial oscillati<strong>on</strong>(al<strong>on</strong>g the DT axis and the cavity holder) has a small amplitude due to the tight axialc<strong>on</strong>finement in the standing-wave trap of better than λ DT /2 = 0.515 µm. Thus thisoscillati<strong>on</strong> does not significantly change the atom-cavity coupling.In c<strong>on</strong>trast, the radial atom c<strong>on</strong>finement is weaker due to the larger radial extensi<strong>on</strong> ofthe trapping potential given by the width of the DT beams, w DT ≫ λ DT /2. In harm<strong>on</strong>icapproximati<strong>on</strong> the radius of the atomic spatial distributi<strong>on</strong> can be found as [51, Eq. (2.62)]√k B Tw rad =m Ω 2 (4.27)rad


4.2 Basics of cavity quantum electrodynamics 89<strong>Cavity</strong> transmissi<strong>on</strong> T andnormalized absolute coupling g1.00.80.60.40.2absolute atom-cavity coupling gcavity transmissi<strong>on</strong> Tmean cavity transmissi<strong>on</strong>0.00.0 0.5 1.0 1.5 2.0Positi<strong>on</strong> al<strong>on</strong>g the cavity axis [λ probe]Figure 4.4: Expected cavity transmissi<strong>on</strong> <strong>for</strong> an atom oscillating al<strong>on</strong>g the cavity axis.If an atom would be permanently coupled to the cavity field with the coupling strengthof g 0 /2π = 18 MHz, we would expect no cavity transmissi<strong>on</strong>. However, due to oscillati<strong>on</strong>of the atom, the expected mean cavity transmissi<strong>on</strong> is 3.2 % of the maximal <strong>on</strong>e.with T the atomic temperature and Ω rad the radial oscillati<strong>on</strong> frequency in the trap givenby expressi<strong>on</strong> (1.16). For the realistic experimental parameters of w DT = 42 µm andU 0 = 1.4 mK and supposing T ≈ 125 µK (the Doppler temperature), 1 we get Ω rad =2π × 2.3 kHz and w rad = 6.2 µm. Another way to determine w rad would be to directlymeasure the width of the atomic fluorescence spot from the image of a trapped atom.Since the oscillati<strong>on</strong> amplitude w rad is much larger than the periodicity of the cavitymode of λ probe /2 = 0.426 µm, the atom oscillating in the DT al<strong>on</strong>g the cavity axis willprobe all possible atom-cavity coupling strengths from zero to maximum. If the bandwidthof the transmissi<strong>on</strong> measurement is smaller than Ω rad , which is the case in our experiment,the detected cavity transmissi<strong>on</strong> is an average over all possible transmissi<strong>on</strong> levels during<strong>on</strong>e oscillati<strong>on</strong> period.To determine the mean transmissi<strong>on</strong>, we numerically calculate the cavity transmissi<strong>on</strong>while the atom moves al<strong>on</strong>g the cavity standing-wave mode. The numerical method isbased <strong>on</strong> solving the master equati<strong>on</strong> of the coupled atom-cavity system in the presenceof dissipati<strong>on</strong>, see Secs. 4.2.3 – 4.2.4. We c<strong>on</strong>sider the res<strong>on</strong>ant case, i.e. when the cavityres<strong>on</strong>ance, the probe laser frequency, and the atomic res<strong>on</strong>ance frequency are all equal,and the maximum atom-cavity coupling strength of g 0 = 2π ×18 MHz. The thick curve inFig. 4.4 corresp<strong>on</strong>ds to the spatial variati<strong>on</strong> of the coupling rate g al<strong>on</strong>g the cavity axis,which is proporti<strong>on</strong>al to the local field amplitude. The thin line is the cavity transmissi<strong>on</strong>,T , calculated <strong>for</strong> the local coupling g. Naturally the regi<strong>on</strong>s of high g and low T (andvice versa) coincide. The dashed horiz<strong>on</strong>tal line indicates the corresp<strong>on</strong>ding mean cavitytransmissi<strong>on</strong> ¯T of about 3.2 %.1 Note that the real atom temperature can be significantly different from the Doppler temperature dueto complicated heating/cooling dynamics inside the cavity mode, see e.g. Sec. 4.5.3.


90 Chapter 4: Deterministic atom-cavity couplingLet us now c<strong>on</strong>sider the radial oscillati<strong>on</strong>s of atoms transversal to the cavity mode(al<strong>on</strong>g y directi<strong>on</strong>). Taking into account equati<strong>on</strong>s (4.2)–(4.3) and supposing <strong>for</strong> simplicitythat an atom is trapped at an antinode of the cavity standing-wave field, the atom-cavitycoupling rate al<strong>on</strong>g the y directi<strong>on</strong> varies as]g(y) = g 0 exp[− y2. (4.28)The average coupling experienced by an atom during its radial transversal oscillati<strong>on</strong> canbe found asg =∫ ∞where p(y) is the atomic distributi<strong>on</strong> in the DT al<strong>on</strong>g y given byp(y) =−∞1√2π wradexpw 2 0p(y) g(y) dy , (4.29)[− y22w 2 rad]. (4.30)For our experimental parameters we get g ≈ 0.965 g 0 . Thus, the reducti<strong>on</strong> of the couplingrate due to the radial transversal oscillati<strong>on</strong>s is 3.5 % and can be neglected compared toother effects reducing g 0 .In general, the coupling rate g can be smaller than the expected <strong>on</strong>e, g 0 , <strong>for</strong> instancebecause of not precisely placing the atom into the mode center. Moreover, the value of g 0is calculated <strong>for</strong> the str<strong>on</strong>gest transiti<strong>on</strong> in a cesium atom, namely between the outermostZeeman states |F =4, m F =±4〉 → |F ′ =5, m ′ F=±5〉. There<strong>for</strong>e, any mixing of m F states(e.g., due to n<strong>on</strong>-circular polarizati<strong>on</strong> of the probe laser) further reduces g and increases¯T . For instance, if the maximal coupling is half the expected <strong>on</strong>e, the mean transmissi<strong>on</strong>¯T is 6.4 % of T max . For the extreme case of g = 0.1 g 0 , ¯T = 0.35 T max .4.2.6 C<strong>on</strong>clusi<strong>on</strong>Summarizing, the presence of the atom-cavity coupling manifests itself via the splittingof the cavity transmissi<strong>on</strong> line in the res<strong>on</strong>ant regime (ω a = ω c ) or via the shift of theres<strong>on</strong>ance in the dispersive regime (|ω a − ω c | ≫ g). If the dissipati<strong>on</strong> is str<strong>on</strong>g, the broadres<strong>on</strong>ance lines screen these spectrum changes. Whereas in the str<strong>on</strong>g coupling regime,i.e. <strong>for</strong> g ≫ (κ, γ), the modificati<strong>on</strong>s to the spectrum of the coupled system are dramaticand should be directly observable.In all experiments presented in this chapter (Secs. 4.4 – 4.5), we detect changes ofthe transmissi<strong>on</strong> of the probe laser through the cavity while transporting few or singleatoms into and out of the cavity mode. The laser is res<strong>on</strong>ant with the empty cavityproviding high cavity transmissi<strong>on</strong> in the absence of atoms. Even <strong>on</strong>e atom placed insidethe cavity mode and coupled to it str<strong>on</strong>gly modifies the system’s spectrum and resultsin a reducti<strong>on</strong> of the probe laser transmissi<strong>on</strong>. The degree of transmissi<strong>on</strong> changes givesus the in<strong>for</strong>mati<strong>on</strong> <strong>on</strong> the coupling strength in the system. Ideally, that is in the str<strong>on</strong>gcoupling regime, the transmissi<strong>on</strong> should completely vanish.


4.3 Locating the cavity mode 914.3 Locating the cavity modeObviously, after integrating the cavity holder into our main experimental setup, there aretwo main requirements to be fulfilled in order to allow deterministic transport of an atominto the cavity mode: The DT should pass through the center of the cavity mode and thepositi<strong>on</strong> of the cavity mode al<strong>on</strong>g the DT axis should be known precisely. For that wefirst need to precisely locate the cavity mode relative to the DT and to the MOT. Then,we should move either the cavity or the DT beams such that the cavity mode perfectlycrosses the DT axis.The coarse pre-alignment of the cavity <strong>on</strong>to the DT is described in Sec. 3.4. Next,to locate the cavity mode precisely, we use transported atoms as a probe. Several experimentalmethods used <strong>for</strong> this purpose are presented below in the course of the secti<strong>on</strong>.Although using different physical processes, all of them are based <strong>on</strong> the interacti<strong>on</strong> oftrapped atoms with the probe laser coupled into the cavity.After the separati<strong>on</strong> between the cavity mode and the DT axis is determined, we makethem intersect by horiz<strong>on</strong>tally shifting <strong>on</strong>e of them or both. Large relative displacementcan be achieved by moving the cavity horiz<strong>on</strong>tally, perpendicular to the DT axis by meansof the 3-D positi<strong>on</strong>er attached to the cavity holder, see Fig. 3.10. For displacements smallerthan several tens of micrometer we shift the DT axis horiz<strong>on</strong>tally by synchr<strong>on</strong>ously tiltingthe last DT mirrors, see Fig. 2.8.4.3.1 Using optical pumpingThe cavity mode can be located by means of optical pumping induced by the probe laserwhich is coupled into the cavity and res<strong>on</strong>ant with the atomic transiti<strong>on</strong> F = 4 → F ′ = 4.Since the probe laser acts <strong>on</strong>ly <strong>on</strong> atoms in the F = 4 state, <strong>on</strong> average two scatteredphot<strong>on</strong>s are enough to pump an atom placed inside the mode from the F = 4 ground stateto F = 3. The maximum of the pumping efficiency measured <strong>for</strong> different cavity positi<strong>on</strong>relative to the DT and <strong>for</strong> different transportati<strong>on</strong> distances indicates the locati<strong>on</strong> of thecavity mode.Experimental sequenceThe experimental sequence realizing this method c<strong>on</strong>sists of the following steps. First, weload the DT with several atoms and prepare them in the F = 4 state. For this purpose,we illuminate the atoms with the MOT repumping laser (res<strong>on</strong>ant with the transiti<strong>on</strong>F = 3 → F ′ = 4) <strong>for</strong> typically 10 ms after transferring them from the MOT into the DT.Then, we transport the atoms over a fixed distance towards the cavity mode excited bythe probe laser. The power of the probe laser coupled to the cavity is typically 0.3 nWresulting in 0.16 mW intra-cavity power (see equati<strong>on</strong> (3.27)). After a short waiting timeof typically 1 ms, which is enough to pump the atoms placed at the field maximum to theF = 3 state, the standing wave is moved back to its original positi<strong>on</strong>. The hyperfine state ofthe atoms is detected by exposing them to a σ + -polarized “push-out” laser, res<strong>on</strong>ant withthe F = 4 → F ′ = 5 cycling transiti<strong>on</strong> [100, 19]. It pushes any atom in the F = 4 state outof the dipole trap, whereas atoms in F = 3 remain untouched and thus stay trapped. The


92 Chapter 4: Deterministic atom-cavity coupling(a)(b)Survival probability [%]Survival probability [%]1008060402010004.50 4.55 4.60 4.65 4.7080604020Transportati<strong>on</strong> distance [mm]0-40 -30 -20 -10 0 10 20 30 40Radial cavity displacement [µm]Figure 4.5: Locating the cavity mode using optical pumping by the probe laser. Themeasurement of the distance between the MOT and the cavity mode (a) and the alignmentof the cavity mode <strong>on</strong>to the DT (b) are per<strong>for</strong>med by maximizing the efficiency ofthe optical pumping <strong>for</strong> a fixed power of the probe laser.survival probability of the atoms, which equals the probability of the populati<strong>on</strong> transferbetween the two hyperfine states, is determined by transferring the surviving atoms backinto the MOT and counting them there. The probability is measured <strong>for</strong> different radialpositi<strong>on</strong> of the cavity relative to the trap and <strong>for</strong> different transportati<strong>on</strong> distances. Themaximum probability indicates the locati<strong>on</strong> of the cavity mode.ResultsThe results of this method are presented in Fig. 4.5. Each data point c<strong>on</strong>tains a c<strong>on</strong>tributi<strong>on</strong>from 30 experimental shots with <strong>on</strong> average 5 atoms each. In the first measurement (a)we vary the transportati<strong>on</strong> distance. When it is too short, the atoms stop outside the cavitymode and stay unaffected by the probe laser. Thus, they do not change their F = 4


4.3 Locating the cavity mode 93state and are subsequently removed out of the trap by applying the push-out laser. For theoptimal transport, the atoms are pumped into the F = 3 state and thus remain trappedafter the applicati<strong>on</strong> of the push-out laser. If the transport is too l<strong>on</strong>g, the atoms arestopped <strong>on</strong> the other side of the cavity mode. However, they still temporarily interactwith the cavity field while crossing it. That is why the survival probability <strong>for</strong> large distancesis higher than that <strong>for</strong> short <strong>on</strong>es. To eliminate this asymmetry in the recordedsignal, <strong>on</strong>e can switch the probe laser <strong>on</strong> <strong>on</strong>ly when the atoms are fully stopped. Withoutapplying the push-out laser, we detect about 80 % of atoms survived the loading of theDT and the transport. This determines the maximal survival probability measured here.From the measured survival probability we determine the optimal transportati<strong>on</strong> distanced trans of 4.59(±0.02) mm. Since after transferring the atoms into the DT they arespatially distributed al<strong>on</strong>g its axis over the size of the MOT of 10 µm, d trans is defined herebetween the centers of the MOT and of the cavity mode. As l<strong>on</strong>g as the MOT size, whichcan be c<strong>on</strong>sidered here as a size of our “atomic probe”, is smaller than the error of d trans ,it does not influence the measurement precisi<strong>on</strong> <strong>for</strong> the present measurement statistics.To precisely align the cavity mode <strong>on</strong>to the DT, we per<strong>for</strong>m the same experimentwith the fixed transportati<strong>on</strong> distance of 4.59 mm, but <strong>for</strong> different radial displacementsof the cavity (in the horiz<strong>on</strong>tal plane perpendicular to the DT axis). The result of thismeasurement is presented in Fig. 4.5(b). As in (a), the signal is saturated to about70 %, measured independently. Yet, in c<strong>on</strong>trast to (a), this signal is symmetric, since theatoms do not cross the cavity mode. By fitting the data with a Gaussian functi<strong>on</strong> wedetermine the optimal displacement of the cavity relative to its initial positi<strong>on</strong> set by acoarse prealignment with an uncertainty of about ±2.0 µm. Although the choice of the fitfuncti<strong>on</strong> is arbitrary and does not properly describe the physics under the measured line<strong>for</strong>m, it is still justified here <strong>for</strong> determining the center of the symmetric distributi<strong>on</strong>.The precisi<strong>on</strong> of both measurements presented is mainly limited by the statistical errorand by the saturati<strong>on</strong> of the measured signal, which broadens its maximum and makes itdifficult to fit it with an appropriate peak functi<strong>on</strong>. The saturati<strong>on</strong> can be reduced by usingeither a smaller probe laser power or a shorter waiting time in the cavity, whereas the statisticalerror can be decreased by using larger number of data points. Note that we cannotuse much more atoms per shot, because of the impossibility of counting them in the MOT.4.3.2 Using atom lossesThe next two methods of locating the cavity mode do not require the manipulati<strong>on</strong> of theatomic states, namely state preparati<strong>on</strong>, optical pumping, and state detecti<strong>on</strong>. Instead,they are based <strong>on</strong> detecting the losses of atoms from the dipole trap induced by the intenseprobe laser coupled to the cavity mode and res<strong>on</strong>ant with the F = 4 → F ′ = 5 cyclingtransiti<strong>on</strong> of the trapped atoms. In both following methods we transport the trappedatoms over a known distance towards the cavity mode, induce str<strong>on</strong>g scattering of theprobe laser phot<strong>on</strong>s which leads to the heating of the atoms out of the DT, and detecttheir survival probability as a functi<strong>on</strong> of the transportati<strong>on</strong> distance.The easiest experimental sequence is the following. First, we load the DT with a


94 Chapter 4: Deterministic atom-cavity coupling(a)Survival probability [%]1008060402004.56 4.58 4.60 4.62 4.64Transportati<strong>on</strong> distance [mm](b)Survival probability [%]1009080706050-50 -40 -30 -20 -10 0 10 20 30 40 50Transverse dipole trap displacement [µm]Figure 4.6: Locating the cavity mode by detecting atom losses induced by the intenseprobe laser. The trap displacement is measured relative to its original positi<strong>on</strong> set bythe coarse DT pre-alignment. Straight lines are fits described by equati<strong>on</strong> (4.31).small known number of atoms, transport them over a known distance towards the cavitymode, switch <strong>on</strong> the probe laser <strong>for</strong> a time l<strong>on</strong>g enough to heat the str<strong>on</strong>gly coupledatoms out of the trap, transport the surviving atoms back, and finally count them againafter reloading into the MOT. In this way we measure the survival probability of thetrapped atoms depending <strong>on</strong> the transportati<strong>on</strong> distance. The maximum of the atomlosses, i.e. the minimum survival probability, corresp<strong>on</strong>ds to the cavity mode positi<strong>on</strong>. Toavoid atomic populati<strong>on</strong> in the F = 3 state, not coupled to the probe laser, we use anauxiliary repumping laser, res<strong>on</strong>ant with the F = 3 → F ′ = 4 transiti<strong>on</strong> and shined inal<strong>on</strong>g the DT axis during the experiment.The measured dependencies are shown in Fig. 4.6. In (a) we vary the transportati<strong>on</strong>distance. The graph (b) corresp<strong>on</strong>ds to the radial alignment of the DT <strong>on</strong>to the cavitymode. In c<strong>on</strong>trast to the previous calibrati<strong>on</strong> method, where the cavity holder wasdisplaced manually, here we displace the dipole trap axis in the horiz<strong>on</strong>tal plane (i.e. perpendicularto the cavity axis) in an automated way by synchr<strong>on</strong>ously tilting the last DTmirrors, see Fig. 2.8. Each point in (a) and (b) is measured from 20 shots with <strong>on</strong> average14 and 5 atoms each, respectively. The difference in atom numbers results in differentstatistical errors in two graphs. The waiting time in the cavity is 10 ms. The intracavity


4.3 Locating the cavity mode 95probe laser power in (a) is about 0.4 mW, whereas in (b) the power was slightly reducedin order not to saturate the measured minimum to zero.To find the minimum of both dependencies we fit them with a suitably chosen fitfuncti<strong>on</strong>y(x) = a |x − b| c + d . (4.31)This functi<strong>on</strong> describes well the measured dependencies which do not show a prominentminimum, see straight lines in Fig. 4.6). The found optimal transportati<strong>on</strong> distance is4.610(±0.001) mm and the optimal DT alignment <strong>on</strong>to the cavity mode is found with aprecisi<strong>on</strong> of about ±3.0 µm. Although the graph (a) is saturated almost to zero and (b) isnot, the precisi<strong>on</strong> of the both measurements is about the same. As in the previous locatingmethod, the transportati<strong>on</strong> distance is measured between the centers of the MOT and ofthe cavity mode. The precisi<strong>on</strong> is better than that of the optical pumping method due tothe use of a better fit functi<strong>on</strong> and can be increased by the same means, e.g., by increasingthe number of the measured points.<str<strong>on</strong>g>Single</str<strong>on</strong>g> atoms <strong>for</strong> probing the cavity modeIn principle, the two previous methods of locating the cavity mode are sensitive to theinitial positi<strong>on</strong> of the atoms, since all of them are transported over the same distancetowards the cavity regardless of their unknown initial locati<strong>on</strong>s. To get rid of thisuncertainty we can use the technique of the precise positi<strong>on</strong> c<strong>on</strong>trol of individual atoms<strong>for</strong> probing the cavity mode. Now, we load the the DT <strong>on</strong>ly with <strong>on</strong>e atom using ournumber-locked loading technique (Sec. 1.4). Then, by means of our positi<strong>on</strong> c<strong>on</strong>troltechnique (Sec. 2.5) we shift the trapped atom to a chosen positi<strong>on</strong> al<strong>on</strong>g the DT axisclose to the cavity mode which has to be probed. After switching <strong>on</strong> the intense probelaser, the standing-wave trap is moved back. As be<strong>for</strong>e, the survival of the atom isdetected by loading it into the MOT. The typical calibrati<strong>on</strong> curves in this case look verysimilar to those shown in Sec. 4.3.2, although <strong>for</strong> achieving the same measurement errorwe can use about 5–6 times less atoms. However, requiring positi<strong>on</strong> c<strong>on</strong>trol technique,this method takes l<strong>on</strong>ger experimental time because of the need to take an ICCD image<strong>for</strong> the initial positi<strong>on</strong> measurement.4.3.3 Using losses of many atoms determined from CCD imageThe last method we have developed <strong>for</strong> locating the cavity mode yields a small statisticalerror because of using a larger number of atoms and is insensitive to the initial atompositi<strong>on</strong> in the DT. As the previous method it is also based <strong>on</strong> detecting losses of atomsinduced by the probe laser. However, here, the atom losses are detected by comparingtwo images of the trapped atoms recorded be<strong>for</strong>e and after transporting them towards thecavity, inducing their losses, and moving them back.In the first step, we load a large number of typically several tens of atoms into the DT.To broaden their spatial distributi<strong>on</strong> in the trap, which normally extends <strong>on</strong>ly over theMOT size of roughly 10 µm, we switch off <strong>on</strong>e of the DT beams <strong>for</strong> 1 ms. In the absence


96 Chapter 4: Deterministic atom-cavity coupling( a)2.0counts [10 ]61.51.00.5( b)0.020 40 60 80 100 120 140 160 180 200pixel1.5counts [10 ]61.00.50.020 40 60 80 100 120 140 160 180 200pixel( c)60ratio [%]4020020 40 60 80 100 120 140 160 180 200pixelFigure 4.7: Locating the cavity mode by inducing the losses of many atoms and detectingthem with a CCD camera. First, we load the DT with several tens atoms and takea CCD image of them. Then, we transport the atoms towards the cavity over a fixedtransportati<strong>on</strong> distance and induce their losses by shining in the probe laser through thecavity. Finally, we move the remaining atoms back and take the sec<strong>on</strong>d image. We repeatthis measurement 40 times and add up all initial images (a) and all final images (b). Theatom losses are detected by building the ratio (c) of the two images. The positi<strong>on</strong> ofthe maximum losses corresp<strong>on</strong>ds to the locati<strong>on</strong> of the cavity mode. Note that here <strong>on</strong>eCCD pixel corresp<strong>on</strong>ds to 0.4967 µm.


4.3 Locating the cavity mode 97of the standing-wave structure, the atoms freely expand al<strong>on</strong>g the trap axis during thistime. Next, we take a CCD image of the trapped atoms and then transport them over thefixed transportati<strong>on</strong> distance of 4.6 mm in the directi<strong>on</strong> of the cavity. By applying theintense probe laser <strong>for</strong> several millisec<strong>on</strong>ds we remove those atoms from the trap, whichare placed within the cavity mode. After transporting the remaining atoms back to theirinitial positi<strong>on</strong> we take the sec<strong>on</strong>d image. For a better signal-to-noise ratio we repeat thisexperimental sequence 40 times and then add up all initial images and all final images.The resulting two images are shown in Fig. 4.7. Image (a) shows the initial homogeneousdistributi<strong>on</strong> of the atoms al<strong>on</strong>g the DT axis. Yet, the final atom distributi<strong>on</strong> inthe image (b) has a dip corresp<strong>on</strong>ding to the absent atoms which were placed into thecavity mode and then lost after applying the probe laser. To analyze these images, we bintheir pixel counts in the vertical directi<strong>on</strong> within a narrow horiz<strong>on</strong>tal regi<strong>on</strong> denoted bytwo red lines. This procedure is the same as that applied <strong>for</strong> measuring the positi<strong>on</strong> ofindividual atoms, see Sec. 2.2. The obtained histograms are shown in Fig. 4.7 below thecorresp<strong>on</strong>ding images.To find the exact distributi<strong>on</strong> of the lost atoms, we normalize the initial distributi<strong>on</strong>of atoms <strong>on</strong>to their final distributi<strong>on</strong>. For this purpose, we build a ratio of the initialhistogram, N be<strong>for</strong>e , to the final histogram, N after , of the corresp<strong>on</strong>ding distributi<strong>on</strong>s aftersubtracting from them the background histogram, N bg , as followsR = N be<strong>for</strong>e − N bgN after − N bg. (4.32)The background level N bg is measured from the same images by binning their pixels withina horiz<strong>on</strong>tal stripe of the same width as that denoted by the horiz<strong>on</strong>tal lines, but located<strong>on</strong> the image away from the illuminated atoms. The obtained histogram R is shown inFig. 4.6(c). A fit with a Gaussian functi<strong>on</strong> yields a peak positi<strong>on</strong> of x peak = 36.46 µmwith an uncertainty of ±0.15 µm. Note that with this method we locate the cavity moderelative to the CCD image. Thus, <strong>for</strong> the transportati<strong>on</strong> distance of d tran = 4.6 mm, thecenter of the cavity mode is located at a distance d mode = x peak + d tran = 4636.5(±0.2) µmfrom the object point <strong>on</strong> the DT axis, which has the first pixel of the CCD camera as itsimage. The whole calibrati<strong>on</strong> procedure takes about 3–4 minutes.Radial positi<strong>on</strong> of the cavity modeTo align the DT radially, i.e. perpendicularly to the cavity mode, we per<strong>for</strong>m theexperiment described above several times <strong>for</strong> different horiz<strong>on</strong>tal displacements of thetrap axis. For each displacement we measure the height of the ratio histogram, seeFig. 4.8. The maximum of the measured dependence corresp<strong>on</strong>ds to the situati<strong>on</strong> whenthe DT goes perfectly through the center of the cavity mode. Here, to determine thepositi<strong>on</strong> of the maximum, we fit the data with a Gaussian functi<strong>on</strong> yielding the optimalDT positi<strong>on</strong> with a precisi<strong>on</strong> of ±0.2 µm.


98 Chapter 4: Deterministic atom-cavity coupling5040Peak height [%]3020100-10 -5 0 5 10 15Transverse dipole trap displacement [µm]Figure 4.8: Aligning the DT <strong>on</strong>to the cavity mode using many atom losses. For eachtrap displacement we per<strong>for</strong>m the measurement of Fig. 4.7 and determine the height ofthe ratio histogram (c). By fitting a Gaussian functi<strong>on</strong> to the obtained dependence, seethe straight line, we determine the optimal horiz<strong>on</strong>tal alignment of the dipole trap.4.3.4 Summary and discussi<strong>on</strong>Summarizing, we have developed and tested several different methods of locating the modeof our high-finesse optical cavity. Although all of them use atoms as a probe, they arebased <strong>on</strong> different physical processes, use different experimental techniques, and have theirown features and advantages.The optical-pumping method is very sensitive to low laser intensities and thus can bealso used to detect weak fields, i.e. weak light scattered off the mirrors. Besides, it isso-called background-free, since in the absence of the cavity mode or the light field thedetected signal stays low.The easiest in realizati<strong>on</strong> is the sec<strong>on</strong>d method based <strong>on</strong> detecting the atom lossesinduced by the probe laser filling the cavity mode. It does not require the manipulati<strong>on</strong>of atomic states and the <strong>on</strong>ly demand is to compare the final number of atoms with theinitial <strong>on</strong>e. In order to be not sensitive to the initial atom positi<strong>on</strong> in the DT, this methodcan be modified by applying the positi<strong>on</strong> c<strong>on</strong>trol technique to precisely positi<strong>on</strong> singleatoms inside the cavity mode. However, it requires the CCD camera and thus l<strong>on</strong>germeasurement times <strong>for</strong> getting sufficient statistics with single atoms.Finally, the last calibrati<strong>on</strong> method allows us to use more (up to 10 times) atomsper shot, thus significantly reducing the statistical error of the measurement. Moreover,by taking CCD images of the atoms be<strong>for</strong>e and after the transport, we get not <strong>on</strong>lythe in<strong>for</strong>mati<strong>on</strong> about the survival probability of the atoms, but also about its positi<strong>on</strong>dependence. In this way, we effectively test all transportati<strong>on</strong> distances simultaneously.


4.4 Transport of many atoms into the cavity 99Because of the slow drifts of the experimental elements (the cavity holder, the DTmirrors, the ICCD camera, etc.) <strong>on</strong> the time scale of several days, the calibrati<strong>on</strong> ofthe mode positi<strong>on</strong> should be per<strong>for</strong>med regularly. In the experiments described in thefollowing part of this chapter we have alternately used all these methods, depending <strong>on</strong>the planned experiment. For instance, if the experiment does not require the use of theCCD camera, we calibrate the cavity positi<strong>on</strong> by the first two methods. In this way wedo not need to look <strong>for</strong> the optimal parameters <strong>for</strong> illuminating the trapped atoms <strong>on</strong>lyin order to locate the cavity.4.4 Transport of many atoms into the cavityAfter the cavity is locked and its mode is located, we can tackle the main goal of the currentresearch project – deterministic coupling of atoms to the mode of a high-finesse cavity andits detecti<strong>on</strong>. First, we examine our system with many (from few to several tens) atoms inorder to qualitatively study the influence of different important experimental parameters<strong>on</strong>to the trapped atoms and the atom-cavity coupling.For transporting several atoms we do not use the method of precise positi<strong>on</strong>ing ofsingle atoms inside the cavity. Since the trapped atoms are distributed al<strong>on</strong>g the DTover about 10 µm (diameter of the MOT), after a correct transport all of them should besimultaneously coupled to the cavity mode, having a width of about 46 µm, although withdifferent coupling rates. By transporting many atoms into the cavity, we are less sensitiveto the local atom-cavity coupling strength experienced by each atom. Even if at least <strong>on</strong>eof them is str<strong>on</strong>gly coupled, the cavity transmissi<strong>on</strong> should completely vanish. Besides,according to equati<strong>on</strong> (4.16), the simultaneous coupling of N atoms effectively increasesthe coupling strength by a factor of √ N. There<strong>for</strong>e, in our very first experiment we useseveral (<strong>on</strong> average 5) atoms and move them into the cavity mode while observing thetransmissi<strong>on</strong> of the probe laser through the cavity, the decrease of which should be theevidence of atom-cavity coupling.The experimental sequence c<strong>on</strong>sists of the following main steps. First, we load severalatoms into the DT and optically pump them into the F = 4 ground state (the probe laseris res<strong>on</strong>ant with the F = 4 → F ′ = 5 transiti<strong>on</strong>). Then, we transport the atoms intothe cavity mode while detecting the probe transmissi<strong>on</strong> through the cavity. After 3 mswaiting time the DT is moved back to its initial positi<strong>on</strong> and the remaining atoms, if any,are loaded back into the MOT and counted there.One of the recorded transmissi<strong>on</strong> traces showing the complete vanish of the cavitytransmissi<strong>on</strong> when the atoms are held inside the cavity is presented in Fig. 4.9. Thetransmitted probe power is measured with the analog APD, see Sec. 3.3.3. From itsoutput voltage we subtract a c<strong>on</strong>stant background, which corresp<strong>on</strong>ds to the APD offsetvoltage and to not-filtered stray light. As indicated in the figure by the horiz<strong>on</strong>tal bar,the probe laser is switched <strong>on</strong> and then off again at about 1 ms and 14 ms, respectively,thus showing that the zero level indeed corresp<strong>on</strong>ds to no probe transmissi<strong>on</strong>. At 2 ms westart to transport the prepared atoms towards the cavity and after a 4 ms transportati<strong>on</strong>


100 Chapter 4: Deterministic atom-cavity couplingtransportprobe laser <strong>on</strong>Transmitted probe laser power [pW]65432100 2 4 6 8 10 12 14 16Time [ms]Figure 4.9: Vanishing cavity transmissi<strong>on</strong> in the presence of several (about 5) atoms inthe cavity mode. The transmissi<strong>on</strong> of the probe laser is measured with the analog APD.Negative transmissi<strong>on</strong> appearing <strong>on</strong> the graph stems from subtracting the APD offsetvoltage and noisy stray light level from the measured signal. The atom transport startsat 2 ms and lasts 4 ms. The arrival of atoms into the cavity mode is accompanied by avanishing probe transmissi<strong>on</strong>. After about 1 ms all atoms are lost and the high cavitytransmissi<strong>on</strong> level is restored.durati<strong>on</strong> they arrive into the cavity mode. As a c<strong>on</strong>sequence of the atom-cavity coupling,the cavity transmissi<strong>on</strong> drops to zero at this moment.The observed increase in transmissi<strong>on</strong> following the drop indicates that all atoms arelost from the DT. This is c<strong>on</strong>firmed by detecting no surviving atoms in the MOT at theend of the sequence. On the other hand, about 90 % of the atoms survive the wholeexperimental sequence if the probe and lock lasers are blocked. There<strong>for</strong>e, the atom lossesare presumably induced by <strong>on</strong>e or both of those lasers, as will be investigated below inSecs. 4.4.1 and 4.4.2.Noteworthy, the increase of the transmissi<strong>on</strong> in traces, which show the transmissi<strong>on</strong>drop, is always sudden. This can be explained if a single atom leads to complete extincti<strong>on</strong>of the transmissi<strong>on</strong> due to its str<strong>on</strong>g coupling to the cavity mode. In this case, there isno transmissi<strong>on</strong> until the very last atom leaves the trap. However, if <strong>on</strong>e atom does notsaturate the cavity, the rapid transmissi<strong>on</strong> change can nevertheless occur if last atoms arelost so fast that the detecti<strong>on</strong> bandwidth does not allow us to resolve the intermediatetransmissi<strong>on</strong> levels in the detected signal corresp<strong>on</strong>ding to different numbers of still coupledatoms. In the current experiment we cannot distinguish these two cases and <strong>on</strong>ly in Sec. 4.5we will see that <strong>on</strong>e atom does not make the cavity transmissi<strong>on</strong> vanish completely.


4.4 Transport of many atoms into the cavity 101Transmitted probe laser power [pW]3.53.02.52.01.51.00.5transport0.00 2 4 6 8 10 12Time [ms](a) P LOCK= 25 µW(b) P LOCK= 0.33 µW(c) with repumper <strong>on</strong>Figure 4.10: Influence of lock laser power and repumping laser <strong>on</strong> the transmissi<strong>on</strong>decrease. At 4 ms the atoms enter the cavity mode. The power of the lock laser coupledinto the cavity is 25 µW in (a) and 0.33 µW in (b) and (c). An additi<strong>on</strong>al repumping laseris applied al<strong>on</strong>g the DT in (c). The transmissi<strong>on</strong> increase in each graph is not sudden as,e.g., in Fig. 4.9 because 64 individual traces have been averaged to yield each respectivetrace.4.4.1 Lock laser and repumping laserDifferent recorded and analyzed individual transmissi<strong>on</strong> signals have a different durati<strong>on</strong>of the low-transmissi<strong>on</strong> phase. Yet, in some experimental shots no transmissi<strong>on</strong> drop isobserved at all. In order to quantitatively characterized the observed transmissi<strong>on</strong> changes,we repeat the same experimental sequence many times and then average the measuredAPD signals. The result of averaging over 64 single traces <strong>for</strong> three different sets ofexperimental parameters is shown in Fig. 4.10.In the measurement (a) the lock laser power coupled into the cavity is 25 µW. Thetransmissi<strong>on</strong> does not completely vanish when atoms enter the cavity mode at 4 ms indicatingthat not all traces out of 64 had a transmissi<strong>on</strong> drop to zero. The observed increasein transmissi<strong>on</strong> following the drop reveals the loss of all atoms within less than 0.3 ms.Suspecting the lock laser to be <strong>on</strong>e of the reas<strong>on</strong>s <strong>for</strong> the atom losses, we reduceits power to about 0.33 µW in the next measurement (b). Indeed, the drop durati<strong>on</strong> isextended to about 1 ms, although it is still short. After 3 ms we move the standing waveto its initial positi<strong>on</strong> and detect that about 25 % of the atoms have survived the wholeexperimental procedure, although the initial transmissi<strong>on</strong> level in Fig. 4.10(b) is quicklyrestored. This surprising result can be explained by optical pumping of the atoms intothe F =3 ground state. The pumping is induced by the probe laser, which off-res<strong>on</strong>antlyexcites the atoms into the F ′ = 4 state, from where their are free to decay into the darkF =3 state. There the atoms are not sensitive to the cavity field, resulting in the restored


102 Chapter 4: Deterministic atom-cavity coupling200Storage time [ms]15010050010 100Transmitted probe laser power [pW]Figure 4.11: Storage time in the cavity depending <strong>on</strong> probe laser power. The storagetime is defined as the time until all atoms are lost. Each point is measured from 20transmissi<strong>on</strong> traces with several tens atoms each.transmissi<strong>on</strong>. After moving the DT back, these atoms can be loaded into the MOT andsubsequently detected there.In order to depopulate the F = 3 state we apply an additi<strong>on</strong>al repumping laser al<strong>on</strong>gthe DT in the measurement (c). It is derived from the MOT repumper and is res<strong>on</strong>antwith the F =3→F ′ =4 transiti<strong>on</strong>. Its power of 0.2 µW provides an efficient pumping ratewhile keeping the unwanted light scattering off the cavity mirrors low. Now, the drop inthe cavity transmissi<strong>on</strong> is more significant compared to that of (b).4.4.2 Reducing the probe laser powerIn spite of the reduced lock laser power, the atoms are nevertheless lost after several millisec<strong>on</strong>ds.Further lowering of the lock power does not improve the situati<strong>on</strong>. Thus, wec<strong>on</strong>clude that the main heating is now induced by a too high power of the probe laser.Despite the Rabi splitting in the coupled atom-cavity system, which shifts the systemres<strong>on</strong>ance away from the probe laser frequency, the small fracti<strong>on</strong> of the probe power offres<strong>on</strong>antlycoupled into the cavity can still affect the atoms and eventually heat them outof the DT.A straight<strong>for</strong>ward way to increase the lifetime of atoms in the cavity is thus to reducethe power of the probe laser. Figure 4.11 shows the storage time <strong>for</strong> different powers ofthe probe laser. Here we define the storage time as the time until all atoms are lost andmeasure it between the transmissi<strong>on</strong> drop and its following increase to a half maximumlevel after averaging over many single traces. Note that in this measurement we havesignificantly increased the storage time relative to that of Fig. 4.10(c) by using more(about 50) atoms per shot, see next Sec. 4.4.3.


4.4 Transport of many atoms into the cavity 103Transmitted probe laser power [pW]43210average # ofatoms/shot817223551620.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35Time [s]Figure 4.12: <strong>Cavity</strong> transmissi<strong>on</strong> <strong>for</strong> different numbers of coupled atoms. Each traceis an average over 50 individual measurements resulting in a slow transmissi<strong>on</strong> increaseinstead of an instantaneous <strong>on</strong>e when all atoms are lost. See text <strong>for</strong> details.By reducing the probe power, the storage time increases almost exp<strong>on</strong>entially. Sinceno saturati<strong>on</strong> in the storage time is observed, the probe laser power should be furtherreduced. However, since the minimal power detectable with the analog APD used hereis about 0.5 pW, the use of an even lower probe powers requires a more sensitive lightdetector. We there<strong>for</strong>e use a single phot<strong>on</strong> counting module in all measurements describedin Sec. 4.5. This allows us to reduce the power of the probe laser by a factor of about100 and c<strong>on</strong>sequently to increase the storage time of single atoms over several sec<strong>on</strong>ds, seee.g. Fig. 4.19.4.4.3 More atoms, l<strong>on</strong>ger storage timeAccording to equati<strong>on</strong> (4.16), the Rabi splitting is larger <strong>for</strong> a lager number of atoms.Thus, simultaneously coupling more atoms to the cavity mode shifts its res<strong>on</strong>ance furtheraway from the probe frequency resulting in less probe power coupled to the cavity.C<strong>on</strong>sequently, <strong>on</strong>e expects the increase of the storage time of the atoms.Figure 4.12 shows the cavity transmissi<strong>on</strong> measured <strong>for</strong> different numbers of transportedatoms. Each trace presented is an averaged over 50 single measurements. Theaverage number of atoms <strong>for</strong> the three first measurements was measured as usually fromthe MOT fluorescence, while in the last three measurements it was estimated from theMOT loading time since the fluorescence from more than about 25 atoms saturates theAPD and cannot be measured anymore. Since two-atom losses in the DT are more prob-


104 Chapter 4: Deterministic atom-cavity coupling200Storage time [ms]1501005000 10 20 30 40 50 60 70Average number of atoms per shotFigure 4.13: Storage time in the cavity depending <strong>on</strong> the atom number. The observedn<strong>on</strong>-logarithmic dependence shows a collective effect of many atoms which significantlyincreases their storage time. See text <strong>for</strong> details.able when loading more atoms [56], the estimated atom numbers in these measurementscan be systematically higher than actual <strong>on</strong>es by up to 20 %.The measured dependence of the storage time t s <strong>on</strong> the number of atoms initiallycoupled to the cavity, N 0 , is shown in Fig. 4.13. For large N 0 , t s increases steeply andn<strong>on</strong>-linearly. This result is opposite to that of “usual” atom losses where each atom getslost independent <strong>on</strong> the current number of trapped atoms. For instance, if the lifetime ofatoms is limited by collisi<strong>on</strong>s with background gas, which is a probabilistic process, theatom number decays exp<strong>on</strong>entially as N(t) = N 0 e −t/τ and the time t s when all atomsare <strong>on</strong> average lost reads t s = τ ln N 0 . For the case of the atom losses induced by aheating process [101], see e.g. Fig. 3.16 <strong>for</strong> the losses due to AOM phase noise, the atomdecay can be still approximated by an exp<strong>on</strong>ential <strong>for</strong> large times, thus also leading to thelogarithmically dependence of t s .The result in Fig. 4.13 looks more like exp<strong>on</strong>ential rather than logarithmic dependence.There<strong>for</strong>e, we observe that the atoms simultaneously coupled to the cavity collectivelyprotect themselves from quickly being lost. This result agrees with our assumpti<strong>on</strong> thatthe more atoms we couple to the cavity, the less probe power and thus heating theyexperience.4.4.4 Multiple transportIn a dem<strong>on</strong>strati<strong>on</strong> experiment per<strong>for</strong>med with many atoms, we shuttle several tens ofatoms many times between the MOT positi<strong>on</strong> and the cavity mode. The cavity transmissi<strong>on</strong>is thereby “switched <strong>on</strong> and off”, see Fig. 4.14. At the end of the sequence all atomsare lost. Note that, during this measurement, the atoms have passed more that 16 cmdistance in total, which is indeed a macroscopic distance in atomic physics.


4.5 A single atom inside the cavity 105Transmitted probe laser power [pW]65432100 50 100 150 200 250Time [ms]Figure 4.14: Multiple transport of a large number (several tens) of atoms into the cavity.The atoms are shuttled between the MOT positi<strong>on</strong> and the cavity mode staying insidethe latter <strong>for</strong> 2 ms. The total transportati<strong>on</strong> distance is about 16 cm.Summarizing, in the first experiments, carried out with many atoms, we have foundout the influence of some important experimental parameters <strong>on</strong> the lifetime of atoms inthe DT while being coupled to the cavity. We have investigated the effect of the probeand lock laser power, of the presence of the repumping laser, and of the number of initiallycoupled atoms. In order to detect single atoms in the cavity, we have to further reducethe probe laser power and to use a single phot<strong>on</strong> counting detector <strong>for</strong> observing thecavity transmissi<strong>on</strong>. This is realized in the next secti<strong>on</strong> where a series of measurementswith <strong>on</strong>ly <strong>on</strong>e atom coupled to the cavity is presented.4.5 A single atom inside the cavityIn the following, I describe experiments <strong>on</strong> deterministic transport and coupling of singleatoms to the cavity mode. The atoms are placed inside the cavity using our precise positi<strong>on</strong>c<strong>on</strong>trol technique. The cavity transmissi<strong>on</strong> is detected by a single phot<strong>on</strong> counting module,allowing us to significantly reduce the probe laser power and to avoid the loss of atomsduring the measurement. The atom-cavity coupling is indicated by a reducti<strong>on</strong> of theprobe transmissi<strong>on</strong>.4.5.1 Probe laser parametersThe main parameters of the probe laser which are of relevance to our experiments are itspower, polarizati<strong>on</strong>, and detuning from the atomic res<strong>on</strong>ance.


106 Chapter 4: Deterministic atom-cavity couplingPowerThe transmissi<strong>on</strong> of the probe laser is detected by the SPCM, see Sec. 3.3.3. The measuredoptical power is related to the detected count rate, R SPCM , byP SPCM = R SPCM hν probe (4.33)with ν probe denoting the frequency of the probe laser. According to equati<strong>on</strong> (3.28), thetransmitted power is related to the intra-cavity power by the transmissi<strong>on</strong> of the cavitymirror T . The detecti<strong>on</strong> efficiency of the SPCM, including phot<strong>on</strong> losses in the opticalsystem between the cavity and the detector, is η total ≈ 0.089, see equati<strong>on</strong> (3.18). Thus,the intra-cavity light power is given byP intra = 1 TR SPCM hν probeη total. (4.34)Similarly, we find the intra-cavity phot<strong>on</strong> number ofN intra = 1 TR SPCM τ round−tripη total, (4.35)where τ round−trip = 2L/c is the round-trip time in the cavity of a length L.As an example, 10 kHz count rate of the SPCM corresp<strong>on</strong>ds to P intra = 80 nW andN intra = 0.09 phot<strong>on</strong>s. Typically, we use probe powers resulting in a detector count rateof 2–6×10 4 counts/s.Polarizati<strong>on</strong>The cavity shows birefringence induced by mechanical stress <strong>on</strong> the glued mirror substrates,see Sec. 3.3.4. The two birefringent modes have linear polarizati<strong>on</strong> and are orientedparallel and perpendicular to the cavity holder axis, respectively. All experimentsin this chapter have been per<strong>for</strong>med with the probe laser linearly polarized al<strong>on</strong>g the DTaxis.FrequencyThe probe laser is derived from the MOT cooling laser locked to the F = 4 → F ′ = 3/5crossover transiti<strong>on</strong>, see Sec. 3.3.2, and is thus ∆ 4→3/5 /2π = −226.12 MHz detuned belowthe F = 4 → F ′ = 5 transiti<strong>on</strong> (<strong>for</strong> cesium level structure see Appendix A). Its frequencyis further c<strong>on</strong>trolled by two AOMs, see Fig. 3.12, set up in a double-pass c<strong>on</strong>figurati<strong>on</strong>.The first modulator, AOM 1, is driven in the −1 st diffracti<strong>on</strong> order with a modulati<strong>on</strong>frequency in the range of 80–170 MHz. The sec<strong>on</strong>d modulator, AOM 3, is driven in the+1 st diffracti<strong>on</strong> order with a frequency of 180–300 MHz. Thus, the combinati<strong>on</strong> of thetwo AOMs allows us to tune the probe laser frequency by ∆ AOM /2π ≈ ±200 MHz aboutthe atomic F = 4 → F ′ = 5 transiti<strong>on</strong>.Apart from the laser frequency, the detuning of the probe laser also depends <strong>on</strong> theexact atomic res<strong>on</strong>ance frequency. In the far-off-res<strong>on</strong>ance dipole trap, the atomic levels


4.5 A single atom inside the cavity 10710excited F’=5 stateground F=4 stateLight shift [MHz]0-10-20σ − σ +π-30-5 -4 -3 -2 -1 0 1 2 3 4 5Zeeman m FstatesFigure 4.15: Light shift of the atomic excited and ground states in the presence of thelinearly polarized DT laser (power of P YAG = 3 W per beam, beam radius of w DT =42 µm). The str<strong>on</strong>gest σ − , π, and σ + transiti<strong>on</strong>s are shown as arrows.are subject to the ac Stark shift (light shift), see Fig. 4.15. The m F -dependent shifts<strong>for</strong> the ground (F = 4) and the excited (F ′ = 5) atomic levels are calculated by takinginto account the multi-level structure of a cesium atom interacting with a n<strong>on</strong>-res<strong>on</strong>antclassical electro-magnetic field [55, App. A]. The relevant DT parameters used are thepower of each DT beam of P YAG = 3 W and the beam radius of w DT = 42 µm.While the light shift of the Zeeman multiplicity of the (6S 1/2 , F = 4) ground statedoes not depend <strong>on</strong> the m F sublevels, the excited (6P 3/2 , F ′ = 5) state is split in m Fdue to the coupling to the higher lying S and D states [55]. The light shift ∆ ls /2π of aspecific transiti<strong>on</strong> is then the difference of the shifts of the corresp<strong>on</strong>ding m F sublevels.The largest shift of −33 MHz is found <strong>for</strong> the π-transiti<strong>on</strong> between the m F = 0 states.The outer-most σ + - and σ − -transiti<strong>on</strong>s have the smallest shift of −15 MHz.In Ref. [102] a str<strong>on</strong>g m F - and positi<strong>on</strong> dependence of the transiti<strong>on</strong> frequency hasbeen avoided by using the state-insensitive trapping of atoms with the DT operating atthe “magic” wavelength of 935 nm <strong>for</strong> cesium. At this frequency, the sum of ac Starkshifts coming from different optical transiti<strong>on</strong>s results in the ground 6S 1/2 and excited6P 3/2 states both being shifted downwards by comparable amounts. The dependence <strong>on</strong>(F ′ , m ′ F) is small and thus does not significantly change the atomic frequency.Summing up the effects given above, the laser detuning from the atomic res<strong>on</strong>ance,∆ p = ω p − ω a , is given by∆ p = ∆ 4→3/5 + ∆ AOM + ∆ ls . (4.36)


108 Chapter 4: Deterministic atom-cavity couplingGuiding magnetic fieldThe orientati<strong>on</strong> of the guiding magnetic field sets the polarizati<strong>on</strong> of the probe laser: Thelinearly-polarized probe laser with a perpendicular magnetic field can induce σ + - and σ − -transiti<strong>on</strong>s at the same time, whereas with a parallel orientati<strong>on</strong> of the guiding field to thelight polarizati<strong>on</strong> <strong>on</strong>e can drive <strong>on</strong>ly π-transiti<strong>on</strong>s. There<strong>for</strong>e, since the light shift ∆ ls isdifferent <strong>for</strong> different transiti<strong>on</strong>s, the frequency detuning ∆ p depends <strong>on</strong> the orientati<strong>on</strong>of the guiding field and <strong>on</strong> the m F sublevel occupied.The occupati<strong>on</strong> of the Zeeman levels depends <strong>on</strong> the optical pumping caused by theprobe laser. For instance, in the case of no Zeeman splitting, the π-light pumps about34 % of the atomic populati<strong>on</strong> into the m F = 0 state and about 24 % into each adjacentm F = ±1 state. In c<strong>on</strong>trast, if exposed to the (σ + + σ − ) light, 39 % of the populati<strong>on</strong> isc<strong>on</strong>centrated in each outermost Zeeman state. In this way <strong>on</strong>e can calculate the effectivelight shift averaged over the atomic ensemble. However, since we do not know the realDT power affecting the atoms, the exact calculati<strong>on</strong> of the light shifts in our system willbe per<strong>for</strong>med in the future after a more precise characterizati<strong>on</strong> of the system.The atom-cavity coupling rate depends <strong>on</strong> the Zeeman states first of all because of them F dependence of the atom-cavity detuning. In additi<strong>on</strong>, the transiti<strong>on</strong> strength definingthe coupling rate depends <strong>on</strong> the specific m F → m ′ Ftransiti<strong>on</strong>: The coupling g 0 , given byequati<strong>on</strong> (4.22), is calculated <strong>for</strong> the str<strong>on</strong>gest atomic transiti<strong>on</strong> |F = 4, m F = 4〉 → |F ′ =√5, m ′ F= 5〉, whereas the coupling <strong>for</strong> the transiti<strong>on</strong> |F = 4, m F = 0〉 → |F ′ = 5, m ′ F= 0〉 is9/5 times weaker, see Table A.2 in Appendix A <strong>for</strong> dipole matrix elements.C<strong>on</strong>cluding, the orientati<strong>on</strong> of the guiding magnetic field, which defines the polarizati<strong>on</strong>of the probe laser and thus the atomic distributi<strong>on</strong> over the m F states, has a direct influence<strong>on</strong> the atom-cavity coupling. Such an influence has been indeed experimentally observedin Sec. 4.5.5.4.5.2 Experimental sequenceA typical experimental sequence <strong>for</strong> all experiments presented below c<strong>on</strong>sists of the followingmain steps. First, we use our technique of number-triggered loading of the DT inorder to trap <strong>on</strong>ly <strong>on</strong>e atom [56]. Next, we apply the positi<strong>on</strong> c<strong>on</strong>trol procedure <strong>for</strong> placingthe atom inside the cavity [25]: We take an image of the atom, determine its positi<strong>on</strong> inthe trap, calculate the distance to the center of the cavity mode, and then transport theatom there. Shortly be<strong>for</strong>e the atom is transported, we start to observe the transmissi<strong>on</strong>of the probe laser through the cavity. After a fixed observati<strong>on</strong> time (typically 2 s), thedipole trap is moved back to its original positi<strong>on</strong> and the sec<strong>on</strong>d CCD image is recorded.Finally, the surviving atom, if any, is loaded back into the MOT. Thus, during each experiment,we obtain three different types of in<strong>for</strong>mati<strong>on</strong> <strong>on</strong> the system: first, the atom’spositi<strong>on</strong> in the DT be<strong>for</strong>e and after transport inferred from the two CCD images; sec<strong>on</strong>d,transmissi<strong>on</strong> of the probe and lock lasers through the cavity detected by the SPCM andthe APD, respectively; third, survival of the atom, detected at the end of the sequence inthe MOT.


4.5 A single atom inside the cavity 10930atom in the cavity25SPCM counts [10 3 /s]201510500.0 0.5 1.0 1.5 2.0 2.5Time [s]Figure 4.16: The rapid changes of the atom-cavity coupling seen in the cavity transmissi<strong>on</strong>signal are assigned to hopping of the atom between potential wells of the DT.The horiz<strong>on</strong>tal bar indicates the time when the atom is hold inside the cavity betweentwo transports.4.5.3 Hopping atomsA typical cavity transmissi<strong>on</strong> signal in the presence of <strong>on</strong>e atom can be seen in Fig. 4.16.Here, the atom enters the cavity at 0.03 s and is removed from the cavity mode at 2.64 s.These events are accompanied with the corresp<strong>on</strong>ding decrease and increase of the transmissi<strong>on</strong>.Other changes in the cavity transmissi<strong>on</strong> indicate rapid changes of the atomcavitycoupling strength which is proporti<strong>on</strong>al to the local probe laser intensity seen bythe trapped atom. These changes can arise from the atom hopping between different potentialwells of our standing-wave DT. Such hopping can move the atom away from orcloser to the center of the cavity mode, thus changing its coupling.The phenomen<strong>on</strong> of jumping atoms in the DT is often observed in our experimentduring their c<strong>on</strong>tinued imaging with the CCD camera when the illuminati<strong>on</strong> of the atomsis not optimal [55]. Any laser cooling process str<strong>on</strong>gly depends <strong>on</strong> the laser detuningrelative to the atomic res<strong>on</strong>ance frequency. If it changes sign, the laser radiati<strong>on</strong> can, e.g.,induce heating of the atom instead of cooling. An atom trapped in the dipole potential andoscillating there has a complicated time dependence of the res<strong>on</strong>ance frequency becauseof the positi<strong>on</strong>-dependent ac Stark shift of the atomic levels. Moreover, this variati<strong>on</strong>depends <strong>on</strong> the atom temperature. Thus, in general, illuminating the atoms in the DTcan result in alternating cooling and heating: At some moment the atom can be heatedout of the trap and, being still under illuminati<strong>on</strong> of the cooling lasers, cooled down againand subsequently recaptured in a different potential well of the DT.The same effect occurs inside the cavity mode where the probe laser plays a role of theilluminating laser field, although <strong>on</strong>ly al<strong>on</strong>g <strong>on</strong>e directi<strong>on</strong>. The trapped atom can undergo


110 Chapter 4: Deterministic atom-cavity coupling25Cumulative distributi<strong>on</strong>20151050-40 -30 -20 -10 0 10 20Hopping distance d hop[µm]Figure 4.17: Hopping of atoms al<strong>on</strong>g the DT. The hopping distance d hop is measuredas the difference of the atom positi<strong>on</strong> be<strong>for</strong>e transporting the atom into the cavity modeand after moving it back by the same transportati<strong>on</strong> distance.complicated dynamics in the crossed fields of the probe laser and the DT including hoppingbetween DT potential wells. As seen from the transmissi<strong>on</strong> traces, see e.g. Fig. 4.16, theatom can hop both into the regi<strong>on</strong> of a higher probe laser intensity and of a lower <strong>on</strong>e.A more direct way to observe the atom hopping is to compare the atom positi<strong>on</strong> be<strong>for</strong>etransporting it into the cavity mode and after moving it back over the same transportati<strong>on</strong>distance. For this purpose, we measure the hopping distance, d hop , as the differencebetween the final and initial positi<strong>on</strong> of the atom in the DT measured from two CCDimages taken at the corresp<strong>on</strong>ding instants. Of course, we c<strong>on</strong>sider <strong>on</strong>ly those experimentalshots where the atom is not lost until the end of the sequence, which happens in about30 % of the cases.A str<strong>on</strong>g correlati<strong>on</strong> between changes in the cavity transmissi<strong>on</strong> and the measuredd hop has been observed. If the transmissi<strong>on</strong> stays c<strong>on</strong>stant during the whole observati<strong>on</strong>time, d hop is always equal to zero within the measurement error. In c<strong>on</strong>trast, Fig. 4.17shows the cumulative distributi<strong>on</strong> of the measured hopping distances d hop <strong>on</strong>ly <strong>for</strong> thetraces where the cavity transmissi<strong>on</strong> shows at least <strong>on</strong>e change during the observati<strong>on</strong>time. One can distinguish two regi<strong>on</strong>s of the most probable d hop : around about −30 µmand 10 µm, respectively. They are separated by about twice the waist of the cavity modeof w 0 = 23.2 µm. We c<strong>on</strong>clude that the atoms hop until they are trapped at the edges ofthe cavity mode At these points of reduced probe laser intensity, the heating of the atomsand thus their further hopping is suppressed. The asymmetry of the two regi<strong>on</strong>s withrespect to the zero point can be explained by the unprecise knowledge <strong>on</strong> the locati<strong>on</strong>of the cavity mode al<strong>on</strong>g the DT. Indeed, be<strong>for</strong>e this set of measurements, we have notprecisely located the cavity mode using <strong>on</strong>e of the experimental procedures described inSec. 4.3.2, but instead we have used an old calibrati<strong>on</strong> from a previous measurement day.


4.5 A single atom inside the cavity 111As a result, the atoms have been positi<strong>on</strong>ed about 10 µm from the center of the cavitymode in the MOT directi<strong>on</strong>.By per<strong>for</strong>ming the same measurements in the absence of the probe laser, we observeneither atom losses nor hopping. Thus, the heating of the atoms in these measurementsas well as their subsequent cooling are induced with the assistance of the probe laser.4.5.4 Influence of the lock laserIn Sec. 4.4.1 we have already seen evidence that the lock laser also affects the atoms placedinside the cavity mode. Below I present a more detailed investigati<strong>on</strong> of these effects.Dipole interacti<strong>on</strong>Being res<strong>on</strong>ant with the cavity mode, the lock laser <strong>for</strong>ms a standing wave between thecavity mirrors. Due to its blue detuning with respect to the atomic transiti<strong>on</strong>, the locklaser produces a repulsive standing-wave potential, given by equati<strong>on</strong> (1.7), which pushesthe atoms away from regi<strong>on</strong>s of high laser intensity, i.e. from its antinodes.The dipole potential al<strong>on</strong>g the cavity axis <strong>for</strong>med by the lock laser and the DT is shownin Fig. 4.18. The slow varying envelope corresp<strong>on</strong>ds to the trapping potential of our DTin the vertical (radial) directi<strong>on</strong> al<strong>on</strong>g the cavity axis with a width of w DT = 42 µm anda maximum depth of 1.4 mK. This potential is modulated by the repulsive (i.e. positive)standing-wave potential <strong>for</strong>med by the lock laser. The height of the potential “hills” ofU lock = 0.8 mK results from 1.4 µW of coupled lock laser power. Note that the wavelengthof the lock laser in the figure is stretched by a factor of 5 to visualize the individualpotential hills. From now <strong>on</strong>, we distinguish three cases: the lock laser dipole potentialU lock is smaller, about equal, or larger than the kinetic energy, E atom , of the trapped atomoscillating vertically in the DT. This will define whether the atom will move al<strong>on</strong>g thecavity mode quasi-freely or if it will be c<strong>on</strong>fined between hills of the repulsive lock laserpotential.Figure 4.19 shows the experimental investigati<strong>on</strong> of the influence of the atom c<strong>on</strong>finemental<strong>on</strong>g the cavity axis caused by the lock laser standing wave <strong>on</strong> the cavity transmissi<strong>on</strong>.The lock laser powers coupled into the cavity mode are 0.28 µW (a), 1.4 µW (b),and 2.8 µW (c) producing 160 µK (a), 0.8 mK (b), and 1.6 mK (c) high dipole potential,respectively. The graphs <strong>on</strong> the left represent typical single transmissi<strong>on</strong> traces, while theright graphs are averages over about 20 individual measurements.If U lock < E atom , the oscillati<strong>on</strong>s al<strong>on</strong>g the cavity axis are not suppressed. Then,according to the results of Sec. 4.2.5, the oscillating atom will probe regi<strong>on</strong>s of differentatom-cavity coupling strength resulting in the measured average transmissi<strong>on</strong> which issignificantly different from zero. Thus, the reduced, but not completely vanished cavitytransmissi<strong>on</strong> will stay c<strong>on</strong>stant until the atom leaves the cavity.This case is illustrated in the measurement (a) in Fig. 4.19, where the lock laserproduces the dipole potential of a height of 160 µK. All recorded traces show c<strong>on</strong>stanttransmissi<strong>on</strong> levels during the whole observati<strong>on</strong> time until the atom is moved out of the


112 Chapter 4: Deterministic atom-cavity couplingDipole trap potential [mK]1.00.50.0-0.5-1.0-1.5U lockU DT-60 -40 -20 0 20 40 60Distance from the DT axis al<strong>on</strong>g the cavity axis [µm]Figure 4.18: Dipole potential <strong>for</strong>med by crossed DT and lock laser. The wavelength ofthe lock laser is stretched by a factor of 5 to visualize the individual potential hills.cavity. Noteworthy, n<strong>on</strong>e of them has a transmissi<strong>on</strong> lower than about 50 %. There<strong>for</strong>e, wec<strong>on</strong>clude that in this case the atomic kinetic energy is higher than the lock laser potentialallowing the atom to oscillate al<strong>on</strong>g the probe laser standing wave, thus averaging thecavity transmissi<strong>on</strong>.If the kinetic energy of the atom is about the lock laser potential (U lock ≈ E atom ),the atoms can hop between the different wells of the lock laser trap. The higher thetemperature, the more frequent the hops. This happens because the atom temperature isnot a static property, but changes permanently, both increasing and decreasing, depending<strong>on</strong> many experimental parameters, e.g., local light shift of the atomic transiti<strong>on</strong>, the probelaser power, the coupling strength, etc. Similar hops, although between the potential wellsof the DT, have been observe be<strong>for</strong>e in Sec. 4.5.3.In measurement (b) of Fig. 4.19 the lock laser potential of about 0.8 mK height whichcorresp<strong>on</strong>ds to the total dipole potential shown in Fig. 4.18. The cavity transmissi<strong>on</strong>shows rapid changes, similar to those in Fig. 4.16. However, the sec<strong>on</strong>d image of theatoms taken after moving them out of the cavity shows no atom hopping al<strong>on</strong>g the DTaxis. There<strong>for</strong>e, the coupling strength changes due to hops of atoms al<strong>on</strong>g the cavity axisbetween different potential wells of the lock laser. Note that the transmissi<strong>on</strong> level nowvaries over a wide range of values from very low to maximum.Finally, if U lock > E atom , the lock laser potential prevents the atom from moving al<strong>on</strong>gthe cavity axis. In the combinati<strong>on</strong> with the attractive DT, the resulting potential c<strong>on</strong>finesthe atoms al<strong>on</strong>g the cavity axis to better than λ lock /2 – the periodicity of the lock laserstanding wave. The local atom-cavity coupling strength seen by an atom c<strong>on</strong>fined at theminimum of the lock laser field depends <strong>on</strong> the local power of the probe laser. Because ofthe difference in wavelength, the standing waves of the two lasers have different periods.The beat length between them, i.e. the distance between regi<strong>on</strong>s of the same phase, isd beat =λ probe λ lock= 22.3 µm. (4.37)2(λ probe − λ lock )


4.5 A single atom inside the cavity 113singleaveraged(a)SPCM counts [10 3 /s]605040302010504030201000.0 0.2 0.4 0.6 0.8 1.000.0 0.2 0.4 0.6 0.8 1.0Time [s]Time [s](b)SPCM counts [10 3 /s]4030201025201510500.0 0.5 1.0 1.5 2.0Time [s]00.0 0.5 1.0 1.5 2.0Time [s](c)SPCM counts [10 3 /s]4025302015201010500.0 0.5 1.0 1.5 2.000.0 0.5 1.0 1.5 2.0Time [s]Time [s]Figure 4.19: <strong>Cavity</strong> transmissi<strong>on</strong> <strong>for</strong> different powers of the lock laser: 0.28 µW (a),1.4 µW (b), and 2.8 µW (c) producing a 160 µK (a), 0.8 mK (b), and 1.6 mK (c) highdipole potential, respectively. Left column: typical single traces, right column: averagedtraces over about 20 single measurements. Horiz<strong>on</strong>tal bars indicate the time betweenmoving an atom into and out of the cavity mode.


114 Chapter 4: Deterministic atom-cavity couplingThus, the distance between the regi<strong>on</strong>s of the maximum coupling of the c<strong>on</strong>fined atom tothe cavity field and of no coupling is <strong>on</strong>ly d beat /2 = 11.2 µm.The case of a high lock laser potential is presented in the measurement (c) in Fig. 4.19.Although the hops are still probable (from time to time the atom can gain enough energyfrom some heating process to hop over the potential wall), they happen less often. Forthe time being the placing precisi<strong>on</strong> of an atom al<strong>on</strong>g the cavity axis is low and the atommay end up with different coupling strength to the cavity field. There<strong>for</strong>e, any cavitytransmissi<strong>on</strong> from low to high can be observed if placing the atom into the cavity whichis indicated the still relatively hight averaged transmissi<strong>on</strong>.Un<strong>for</strong>tunately, even higher powers of the lock laser could not be tested because of thesignificantly increased heating rate of the atoms caused by the lock laser which leads tofast atom loss. Already <strong>for</strong> the lock laser powers in (b) and (c) the survival probabilityis about 65 % and 30 %, respectively, measured by loading the surviving atoms back intothe MOT at the end of the experimental sequence.Placement precisi<strong>on</strong>To insure a high coupling, the precisi<strong>on</strong> of the placement of an atom into the lock laserstanding wave should be much better than d beat /2. This precisi<strong>on</strong> is mainly limited bythe amplitude of the radial oscillati<strong>on</strong>s of atoms in the DT (al<strong>on</strong>g the cavity axis), w rad ,be<strong>for</strong>e entering the cavity mode, see Sec. 4.2.5. For the typical experimental parametersand <strong>for</strong> the Doppler temperature of atoms, w rad ≈ 7 µm which is comparable to the halfbeat length of d beat /2. Hence, even <strong>for</strong> perfect alignment of the DT, i.e., with its axiscrossing the overlapping antinodes of the probe and lock laser standing waves, the radialoscillati<strong>on</strong> of the atom in the DT will prevent a precise placement into the maximum ofthe cavity field. Note that the real atom temperature can be significantly higher than theDoppler <strong>on</strong>e due to complicated heating/cooling dynamics of the probe and lock lasers,see Sec. 4.5.3.The placing precisi<strong>on</strong> can be c<strong>on</strong>siderably improved by cooling the atom into its groundoscillatory state which has a radial spread of <strong>on</strong>ly√a 0 == 130 nm, (4.38)2m Ω radwhere the radial oscillati<strong>on</strong> frequency Ω rad is given by expressi<strong>on</strong> (1.16). The corresp<strong>on</strong>dingcooling, which is also required <strong>for</strong> many realistic schemes of atom-atom entanglement ina cavity, is <strong>on</strong>e of the next major projects in our group, see Outlook.A further effect limiting the placing precisi<strong>on</strong> is the drifts of the DT positi<strong>on</strong> al<strong>on</strong>g thecavity axis due to, e.g., thermal drifts of optical elements in the beam path. The standardmethod to estimate this drift is to regularly take images <strong>on</strong> the DT loaded with manyatoms, sum up the detected fluorescence al<strong>on</strong>g the trap axis, and determine the center ofthe obtained distributi<strong>on</strong>, which gives us the vertical positi<strong>on</strong> of the DT axis. Typicallythese slow drifts are up to 5 µm per hour.


4.5 A single atom inside the cavity 115Scattering rateApart from the dispersive dipole interacti<strong>on</strong>, the lock laser can also off-res<strong>on</strong>antly exciteatoms. For instance, <strong>for</strong> an incoming laser power of 1 µW the scattering rate <strong>for</strong> an atomplaced at the field maximum is 24 phot<strong>on</strong>s/s, see equati<strong>on</strong> (1.8), and does not depend <strong>on</strong>the hyperfine ground state. The recoil energy of cesium is E rad /k B = 99 nK. There<strong>for</strong>e,in the absence of all other loss mechanisms, the scattering of the lock laser should resultin a lifetime of atoms stored in a 1 mK deep DT of at least several minutes and is thusnegligible in our experiments.Parametric heating of atomsAlthough the scattering of the lock laser phot<strong>on</strong>s inside the cavity is low, we do observe astr<strong>on</strong>g reducti<strong>on</strong> of the storage time of atoms if using several ten µW of lock laser poweror even less. For instance, in the measurement of Fig. 4.19(c) with 2.8 µW laser power,almost all atoms stay trapped inside the cavity <strong>for</strong> 2 sec<strong>on</strong>ds until we move them out ofthe cavity mode. However, the atoms are always lost if we switch the probe laser off <strong>for</strong><strong>on</strong>ly 10 ms in this experimental sequence after the atoms are placed inside the cavity. Thisindicates str<strong>on</strong>g heating of the atoms caused by the lock laser which is counteracted bysome cooling mechanism originating from the probe laser.The observed atom losses are attributed to parametric heating of atoms in thelock laser dipole trap. 2 This heating results from fluctuati<strong>on</strong>s of the depth of the locklaser potential appearing at the twice radial oscillati<strong>on</strong> frequency of the DT (typicallyseveral kHz). The power fluctuati<strong>on</strong>s of the lock laser inside the cavity arise mainly fromits fluctuating coupling into the cavity mode due to n<strong>on</strong>-ideal stabilizati<strong>on</strong> of the cavitylength <strong>on</strong>to the lock laser. Thus, the parametric heating can be reduced by improvingthe per<strong>for</strong>mance of the cavity stabilizati<strong>on</strong> scheme or by reducing the external noise.4.5.5 Guiding magnetic fieldThe guiding magnetic field splits the atomic Zeeman levels and determines the quantizati<strong>on</strong>axis in the system. The probe laser is linearly polarized al<strong>on</strong>g the DT axis in all presentedexperiments. Thus, recalling Sec. 4.5.1, if the guiding field is applied parallel to thepolarizati<strong>on</strong> of the probe laser, the laser can induce <strong>on</strong>ly π-transiti<strong>on</strong>s with ∆m F = 0, seeFig. 4.15. However, if the quantizati<strong>on</strong> axis is perpendicular to the laser linear polarizati<strong>on</strong>(e.g., oriented al<strong>on</strong>g the cavity or orthog<strong>on</strong>al to both the cavity and the DT), then bothσ + - and σ − -transiti<strong>on</strong>s with ∆m F = ±1 can be induced.In our experiment the applicati<strong>on</strong> of guiding fields is accomplished by three orthog<strong>on</strong>alpairs of coils surrounding the vacuum cell. Typically, we apply magnetic fields of aboutseveral Gauss. Figure 4.20 shows cavity transmissi<strong>on</strong> signals averaged over 20 single traces<strong>for</strong> three different orientati<strong>on</strong>s of the guiding field. In each trace the atoms enter the cavitymode at 0.075 s indicated by a drop in the detected transmissi<strong>on</strong>. After 2 s of observati<strong>on</strong>the remaining atoms are removed from the mode.2 For analysis of various heating mechanisms in a dipole trap see Ref. [51].


116 Chapter 4: Deterministic atom-cavity coupling(a)3SPCM counts [10 /s]604020B-field al<strong>on</strong>g the DT-transiti<strong>on</strong>s00.0 0.5 1.0 1.5 2.0(b)3SPCM counts [10 /s]604020B-field al<strong>on</strong>g the cavity - and -transiti<strong>on</strong>s00.0 0.5 1.0 1.5 2.0(c)3SPCM counts [10 /s]604020B-field perpendicular to cavity and DT - and -transiti<strong>on</strong>s00.0 0.5 1.0 1.5 2.0Time [s]Figure 4.20: Averaged cavity transmissi<strong>on</strong> <strong>for</strong> different orientati<strong>on</strong>s of the guiding magneticfield.The result of measurement (a) with <strong>on</strong>ly π-transiti<strong>on</strong>s possible differs str<strong>on</strong>gly fromother two measurements (b) and (c) where <strong>on</strong>ly σ + - and σ − -transiti<strong>on</strong>s can be induced.Note that the str<strong>on</strong>gest π-transiti<strong>on</strong> is √ 9/5 times weaker than the str<strong>on</strong>gest σ-transiti<strong>on</strong>between outermost Zeeman states (see Table A.2). Thus, we would expect to observethe larger transmissi<strong>on</strong> drop <strong>for</strong> the measurements (b) and (c) compared to (a). Thisdiscrepancy can be qualitatively explained by different probe laser detuning <strong>for</strong> differentlight polarizati<strong>on</strong>s effecting the atom-cavity coupling strength as shown in Fig. 4.15.The two AOMs in these measurements shift the probe laser frequency by∆ AOM /2π = 245 MHz. There<strong>for</strong>e, according to equati<strong>on</strong> (4.36), the probe detun-


4.6 C<strong>on</strong>clusi<strong>on</strong> 117ing <strong>for</strong> the str<strong>on</strong>gest π- and σ-transiti<strong>on</strong>s is ∆ π p/2π = −14.5 MHz and ∆ σ p/2π = 4 MHz,respectively (see Sec. 4.5.1). One sees that the large detuning ∆ π p should decrease theatom-cavity coupling even further, thus c<strong>on</strong>tradicting the results shown in Fig. 4.20.However, in the previous experiments [55] we have observed that due to abberati<strong>on</strong>s anddiffracti<strong>on</strong> the actual trap depth measured from the oscillati<strong>on</strong> frequencies of the trapcan be up to 3.3 times smaller than what we calculate from equati<strong>on</strong> (1.7). If we supposethe real trap depth to be half of the expected <strong>on</strong>e, then the corrected probe detunings are∆ π p/2π = 2 MHz and ∆ σ p/2π = 11.6 MHz. Now, the larger atom-cavity coupling rate <strong>for</strong>the circularly polarized field can be diminished by the larger detuning from the atomicfrequency. A more quantitative analysis of the expected coupling rate should includethe atomic populati<strong>on</strong>s over m F levels, the m F -dependence of the light shifts and of thetransiti<strong>on</strong> strengths, as well as the experimentally measured dipole trap depth.4.6 C<strong>on</strong>clusi<strong>on</strong>Following significant achievements in c<strong>on</strong>trolling single atoms in our standing-wave dipoletrap and a successful design, assembly, and stabilizati<strong>on</strong> of our high-finesse optical res<strong>on</strong>ator,we are now able to realize deterministic atom-cavity coupling requiring rathercomplex experimental sequences. A typical sequence includes cooling and trapping ofatoms in the MOT, loading them into the dipole trap, transporting over half a centimeterinto the cavity mode, probing the atom-cavity system with an external probe laser, andfinally moving the remaining atoms back and analyzing them in the MOT.The first experiments using many atoms helped us to better understand our new systemas well as to test the compatibility of different comp<strong>on</strong>ents and techniques. We haveanalyzed the influence of some important experimental parameters <strong>on</strong> the lifetime of atomsin the DT coupled to the cavity, mainly of the probe and lock laser power, of the presenceof the repumping laser, and of the number of initially coupled atoms.For coupling and detecting <strong>on</strong>e atom in the cavity mode, we have further reduced theprobe laser power and used a single phot<strong>on</strong> counting detector <strong>for</strong> observing the cavitytransmissi<strong>on</strong>. By means of number-triggered loading of the dipole trap with a singleatoms and sub-micrometer positi<strong>on</strong> c<strong>on</strong>trol we prepare and place single atoms into thecavity mode <strong>on</strong> demand. By observing the cavity transmissi<strong>on</strong> we obtain in<strong>for</strong>mati<strong>on</strong> <strong>on</strong>the coupling and dynamics in the system. A series of measurements presented shows thatwe are able to keep single atoms trapped inside the cavity <strong>for</strong> several sec<strong>on</strong>ds. Althoughthe atoms hop inside the mode, their recapture indicates the presence of some coolingmechanisms preventing atom losses and keeping atoms coupled to the cavity.


118 Chapter 4: Deterministic atom-cavity coupling


Chapter 5C<strong>on</strong>clusi<strong>on</strong> and outlookThe major achievement of this work is the deterministic coupling of a single atom to themode of a high-finesse optical res<strong>on</strong>ator. I have presented a set of techniques allowing usto place and hold a single atom inside the cavity mode <strong>on</strong> demand. The coupling betweenatom and cavity field manifests itself by a str<strong>on</strong>g reducti<strong>on</strong> of the cavity transmissi<strong>on</strong>probed by a weak laser. Moveover, we are able to move the atom out of the cavity andto per<strong>for</strong>m a final state analysis. All these steps are preparatory <strong>for</strong> deterministic cavity<strong>QED</strong> experiments with trapped atoms, in particular <strong>for</strong> possible applicati<strong>on</strong>s in quantumin<strong>for</strong>mati<strong>on</strong> processing.The first part of this thesis was devoted to establishing a sub-micrometer positi<strong>on</strong>c<strong>on</strong>trol of single atoms using our “optical c<strong>on</strong>veyor belt” technique and the fluorescenceimaging of trapped atoms. It allows us to place single atoms precisely at the center of adistant cavity mode independently of their initial locati<strong>on</strong> in the dipole trap. Next, I havepresented the manufacturing, assembly, and test of a miniature ultra-high finesse opticalcavity stabilized to the atomic transiti<strong>on</strong>. Its properties allow us to optimally integratethe cavity into the present atom-trapping experiment and to operate the atom-cavitysystem in the str<strong>on</strong>g coupling regime. After transporting a single atom into the cavity,its presence inside the mode is detected as a decrease of the cavity transmissi<strong>on</strong>. Theatoms stay trapped and coupled to the cavity field <strong>for</strong> up to 2 sec<strong>on</strong>ds until we transportthem out of the mode. The basic obstacles preventing the maximal coupling strengthhave been investigated in a set of experiments presented at the end of the thesis.C<strong>on</strong>stant couplingThe next main task is to achieve a c<strong>on</strong>stant and maximum atom-cavity coupling rate. Wehave determined that the basic effects reducing the coupling strength are the placing of anatom not precisely at an antinode of the standing-wave field of the cavity mode, the atomicoscillati<strong>on</strong> in the dipole trap transverse to the cavity axis, a time-varying detuning betweenthe cavity res<strong>on</strong>ance frequency and the atomic transiti<strong>on</strong> resulting from the positi<strong>on</strong>andm F -dependent light shift, and driving an atom <strong>on</strong> all the differently coupled |F =119


120 C<strong>on</strong>clusi<strong>on</strong> and outlook4, m F 〉 → |F ′ = 5, m ′ F〉 transiti<strong>on</strong>s. Based <strong>on</strong> this analysis, we are now planning a set ofmodificati<strong>on</strong>s to our experimental setup. We are still working <strong>on</strong> improving the relativefrequency stability between the cavity and the lock laser. This measure should reduce theintra-cavity power fluctuati<strong>on</strong>s of the lock laser which lead to parametric heating of theatoms, and thus should allow us to increase the lock laser power in order to better c<strong>on</strong>finetrapped atoms al<strong>on</strong>g the cavity axis in order to avoid the oscillating coupling strength.Moreover, better c<strong>on</strong>finement of atoms inside the cavity mode can be achieved by using amore sophisticated geometry of the intra-cavity dipole trap, e.g., by simultaneously usinghigher transversal modes of the cavity [103].Spatial filtering of the dipole trap beams by using optical fibers is supposed to reduceheating of the cavity mirrors, thus improving the cavity stability. In additi<strong>on</strong>, opticalfibers should provide better pointing stability of the dipole trap leading to a c<strong>on</strong>stant trapdepth and a reproducible atom-cavity coupling. Next, characterizati<strong>on</strong> of our dipole trapshould allow us to precisely compensate ac Stark shifts of atomic levels in order to set theright atom-cavity detuning <strong>for</strong> the future experiments.Even if the best possible external stability is achieved, <strong>on</strong>e cannot avoid fluctuati<strong>on</strong>sof the coupling after inserting a thermal atom into the cavity. Thus the next project is toimplement a method of cooling the atoms in the dipole trap and/or in the cavity.Raman and cavity coolingBy cooling an atom far below the Doppler temperature, we can better localize it at thebottom of a trapping potential well. Besides making the atom-cavity coupling rate morec<strong>on</strong>stant, the cooling reduces an undesired variati<strong>on</strong> of the ac Stark shift arising fromthe atomic thermal moti<strong>on</strong> in the dipole trap. Moreover, some applicati<strong>on</strong>s in quantumin<strong>for</strong>mati<strong>on</strong> processing require an atom to be in its oscillatory ground state, which is theultimate cooling limit in an optical lattice.Efficient ground state cooling of atoms in optical lattices has been achieved using Ramansideband cooling [104, 105]. We have dem<strong>on</strong>strated that we can resolve the moti<strong>on</strong>alsidebands due to the axial oscillati<strong>on</strong> of the atoms in our dipole trap [100, 106], which isa necessary prerequisite <strong>for</strong> these cooling schemes to work.The presence of the cavity mode, which is str<strong>on</strong>gly coupled to the atom and whichmechanically affects its moti<strong>on</strong> [107], can give rise to novel cooling mechanisms [108]which are not observed in free space. Such cavity-induced cooling has been dem<strong>on</strong>stratedexperimentally in Refs. [102, 109]. The first evidence of a cooling process in our cavityis the recapture of hopping atoms. Presently, possible implementati<strong>on</strong>s of cavity coolingschemes are discussed in our group.Atom-atom entanglementThe generati<strong>on</strong> and the c<strong>on</strong>trolled manipulati<strong>on</strong> of an entangled state is an importantbenchmark <strong>for</strong> quantum in<strong>for</strong>mati<strong>on</strong> processing experiments. In a cavity <strong>QED</strong> system theinteracti<strong>on</strong> between two atoms, which is required <strong>for</strong> their entangling, can be based <strong>on</strong>


121the exchange of a cavity phot<strong>on</strong>. Despite a simple c<strong>on</strong>cept and a wide range of possibleapplicati<strong>on</strong>s, the entanglement between two ground-state atoms in an optical cavity hasnot been realized yet.There are basically three main approaches to entangle atoms in a cavity: deterministic,adiabatic, and measurement-induced. Deterministic entanglement schemes rely <strong>on</strong>the coherent energy exchange between two atoms simultaneously coupled to the mode of ahigh-finesse cavity. One of the most promising schemes is based <strong>on</strong> a res<strong>on</strong>ant four-phot<strong>on</strong>Raman process involving the cavity mode and an auxiliary laser field [110]. However,the best possible fidelity of this scheme <strong>for</strong> our experimental parameters is at most 85 %under c<strong>on</strong>diti<strong>on</strong> of a precise realizati<strong>on</strong> of a π/2 four-phot<strong>on</strong> pulse [55]. In c<strong>on</strong>trast,adiabatic entanglement schemes are not so sensitive to experimental parameters. However,being in general slower than res<strong>on</strong>ant processes, they are still subject to dissipati<strong>on</strong>[111]. A completely different class of entanglement schemes relies <strong>on</strong> measurement inducedentanglement [112, 113], but being probabilistic they are not well suitable <strong>for</strong> quantumin<strong>for</strong>mati<strong>on</strong> processing.Despite recent progress in manufacturing high-quality mirrors, the achievable valuesof the cooperativity parameter are still too low to permit the high fidelity generati<strong>on</strong> ofentangled atomic states with existing proposals. There<strong>for</strong>e, we are still seeking <strong>for</strong> themost promising entanglement scheme that can be implemented in our system.


122 C<strong>on</strong>clusi<strong>on</strong> and outlook


Appendix AAtomic data <strong>for</strong> 133 CsThe most important properties of the cesium atom relevant <strong>for</strong> this work are summarizedin Table A.1, and a level scheme of the first excited states is shown in Fig. A.1. Most ofthe data is extracted from the data collecti<strong>on</strong> of D. Steck [114].General propertiesMass m 2.21 × 10 −25 kg6 2 S 1/2 ground state hyperfine splitting ω hfs 2π × 9.1926 GHzD 1 -lineWavelength λ D1 894.35 nmNatural lifetime 6 2 P 1/2 τ D1 34.9 nsDecay rate Γ D1 2π × 4.56 MHzD 2 -lineWavelength λ D2 852.11 nmNatural lifetime 6 2 P 3/2 τ D2 30.5 nsDecay rate Γ D2 2π × 5.22 MHzTransiti<strong>on</strong> dipole matrix element 4.479 ea 0〈J = 1/2||er||J ′ = 3/2〉 3.797 × 10 −29 C mDipole moment (see Table A.2) d 3.167 ea 0| F = 4, m F = ±4 〉 | F ′ = 5, m ′ F= ±5 〉 2.685 × 10 −29 C mSaturati<strong>on</strong> intensity I 0 11 W/m 2| F = 4, m F = ±4 〉 | F ′ = 5, m ′ F= ±5 〉Doppler temperature T D 125 µKRecoil energy E rad /k B 99 nKTable A.1: Some physical properties of the 133 Cs atom.123


124 APPENDIX A. ATOMIC DATA FOR 133 CS6 2 P3/2251.4 MHz201.5 MHz151.3 MHzF´=5F´=4F´=3F´=26 2 P1/2F´=4F´=3D2-transiti<strong>on</strong>852 nmD1-transiti<strong>on</strong>894 nm6 2 S1/29 192.6 MHzF=4F=3Figure A.1: Level scheme of the first excited states in 133 Cs.The dipole matrix elements <strong>for</strong> specific |F =4, m F 〉 → |F ′ =5, m ′ F〉 transiti<strong>on</strong>s are listedin Table A.2 as multiples of 〈J =1/2||er||J ′ =3/2〉 = 4.479 ea 0 [114]. The transiti<strong>on</strong>s m F →m ′ F= m F are coupled by π-polarized light, whereas the transiti<strong>on</strong>s m F → m ′ F= m F + 1are coupled by σ + -polarized light.m F =-4 m F =-3 m F =-2 m F =-1 m F =0 m F =1 m F =2 m F =3 m F =4m ′ F=m F +1√190√130√115√19√16√730√1445√25√12m ′ F=m F -√110√8-45√7-30√4-15√5-18√4-15√7-30-√845√1-10Table A.2: Dipole matrix elements <strong>for</strong> |6 2 S 1/2 , F = 4, m F 〉 → |6 2 P 3/2 , F ′ = 5, m ′ F〉 transiti<strong>on</strong>sin 133 Cs, expressed as multiples of 〈J =1/2||er||J ′ =3/2〉.


Appendix B<strong>Cavity</strong> ringingWhen the length of a cavity is scanned faster than the cavity decay time, while beingexcited by a m<strong>on</strong>ochromatic light field, the cavity output shows an amplitude oscillati<strong>on</strong>sin its normal decay profile (see Ref. [87] and references therein). This cavity ringing effectcan be described classically. For this purpose, we c<strong>on</strong>sider the cavity as a damped harm<strong>on</strong>icoscillator with a time-dependent res<strong>on</strong>ance frequency driven by an external periodic <strong>for</strong>ce.The differential equati<strong>on</strong> describing the oscillator’s dynamics readsd 2dt 2 E(t) + 2κ d dt E(t) + ω2 (t) E(t) = A e iω 0t ,(B.1)where κ is the cavity decay rate, ω(t) is its time-dependent res<strong>on</strong>ance frequency, A andω 0 are the amplitude and the frequency of an incident laser field, respectively. The intracavityfield is given byE(t) = E(t) e −iφ(t) e iω 0t(B.2)with E(t) denoting the complex field amplitude and φ(t) denoting the phase. Since thecavity frequency is linearly scanned over ω 0 , its frequency and phase can be expressed asφ(t) = S t 2 /2 (B.3a)ω(t) = ω 0 − S t. (B.3b)The sweep rate S is defined as S = ω FWHM /τ sweep , where τ sweep is the sweep time throughthe cavity linewidth. Since ω FWHM is the inverse of the cavity ring-down time, τ CRD , weget S = (τ sweep τ CRD ) −1 .In an approximati<strong>on</strong> of a slowly varying amplitude and phase of the intracavity fieldcompared to ω 0 , equati<strong>on</strong> (B.1) can be rewritten as a first-order differential equati<strong>on</strong>di AE(t) + κ E(t) = − e iφ(t) . (B.4)dt 2 ω 0For the initial c<strong>on</strong>diti<strong>on</strong> ε(t = −∞) = 0 the soluti<strong>on</strong> to (B.4) readsE(t) = − i A ∫ ∞2ω 00[S(t −]τ)2exp − i − κ τ dτ.2(B.5)125


126 APPENDIX B. CAVITY RINGING1.0Normalized cavity output intensity0.80.60.40.20.00 20 40 60 80 100Scan time [a.u.]Figure B.1: Theoretical cavity ringing signal. The graph is the soluti<strong>on</strong> to equati<strong>on</strong>(B.7) <strong>for</strong> κ = 1 and τ sweep = 0.2 τ CRD .This integral can be taken analytically in term of the error functi<strong>on</strong> 1E(t) = − i A (1 − i) √ π2ω 0 2 √ S(× 1 − erf[exp − i S t22[ (1 − i)√S(κ − iS t)− iS]t)2− i(κ ×2S]). (B.6)Finally, the intensity of the intracavity fields is I(t) = E ∗ (t) E(t). Since the cavityoutput is proporti<strong>on</strong>al to its intracavity intensity, the cavity ringing signal can be describedby[∣ (1 − i)]∣ ∣∣2I out (t) = T I 0 exp(−2κt) ∣1 − erf √ (κ − iS t)(B.7)Swith I 0 = (πA 2 )/(8Sω 0 ) and T the mirror transmissi<strong>on</strong>. The exp<strong>on</strong>ential term in (B.7)describes the cavity decay, while the term with the error functi<strong>on</strong> represents the cavityringing, i.e. oscillati<strong>on</strong>s with increasing frequency. For κ = 1/(2τ CRD ) = 1 and <strong>for</strong>τ sweep = 0.2 τ CRD , which seems to be a reas<strong>on</strong>able proporti<strong>on</strong> <strong>for</strong> our experiments, thetime dependence of the cavity output intensity is shown in Fig. B.1.1 The error functi<strong>on</strong> is defined as erf(x) = 2/ √ π ∫ x0 exp(−t2 ) dt.


Bibliography[1] E. M. Purcell, Sp<strong>on</strong>taneous transiti<strong>on</strong> probabilities in radio-frequency spectroscopy,Phys. Rev. 69, 681 (1946)[2] E. T. Jaynes and F. W. Cummings, Comparis<strong>on</strong> of Quantum and SemiclassicalRadiati<strong>on</strong> Theory with Applicati<strong>on</strong> to the Beam Maser, Proc. IEEE 51, 89 (1963)[3] J. M. Raim<strong>on</strong>d, M. Brune, and S. Haroche, Manipulating quantum entanglementwith atoms and phot<strong>on</strong>s in a cavity, Rev. Mod. Phys. 73, 565–582 (2001)[4] G. Raithel, C. Wagner, H. Walter, L. M. Narducci, and M. O. Scully, The micromaser:Providing ground <strong>for</strong> quantum physics, in P. Berman, ed., <strong>Cavity</strong> QuantumElectrodynamics, p. 57, Academic Press, New York (1994)[5] R. Miller, T. E. Northup, K. M. Birnbaum, A. Boca, A. D. Boozer, and H. J. Kimble,Trapped atoms in cavity <strong>QED</strong>: coupling quantized light and matter, J. Phys. B: At.Mol. Opt. Phys. 38, S551–S565 (2005)[6] G. Rempe, T. Fischer, M. Hennrich, A. Kuhn, T. Legero, P. Maunz, P. Pinkse, andT. Puppe, <str<strong>on</strong>g>Single</str<strong>on</strong>g> atoms and single phot<strong>on</strong>s in cavity quantum electrodynamics, inN. P. Bigelow, J. H. Eberly, C. R. Stroud, and I. A. Walmsley, eds., Coherence andQuantum Optics VIII, pp. 241–248, Kluwer Academic / Plenum Publishers, NewYork (2003)[7] M. A. Nielsen and I. L. Chuang, Quantum computati<strong>on</strong> and quantum in<strong>for</strong>mati<strong>on</strong>,Cambridge University Press (2000)[8] P. Shor, Polynomial-time algorithms <strong>for</strong> prime factorizati<strong>on</strong> and discrete logarithms<strong>on</strong> a quantum computer, SIAM J. Comp. 26, 1484 (1997)[9] L. Grover, Quantum mechanics helps in searching <strong>for</strong> a needle in a haystack, Phys.Rev. Lett. 97, 325 (1997)[10] T. Pellizzari, S. A. Gardiner, J. I. Cirac, and P. Zoller, Decoherence, C<strong>on</strong>tinousObservati<strong>on</strong> and Quantum Computing: A <strong>Cavity</strong> <strong>QED</strong> Model, Phys. Rev. Lett 75,3788 (1995)127


128 BIBLIOGRAPHY[11] J. I. Cirac, P. Zoller, H. J. Kimble, and H. Mabuchi, Quantum state transfer andentanglement distributi<strong>on</strong> am<strong>on</strong>g distant nodes in a quantum network, Phys. Rev.Lett. 78, 3221 (1997)[12] G. R. Guthörlein, M. Keller, K. Hayasaka, W. Lange, and H. Walther, A single i<strong>on</strong>as a nanoscopic probe of an optical field, Nature 414, 49 (2001)[13] J. Ye, D. W. Vernooy, and H. J. Kimble, Trapping of <str<strong>on</strong>g>Single</str<strong>on</strong>g> <str<strong>on</strong>g>Atoms</str<strong>on</strong>g> in <strong>Cavity</strong> <strong>QED</strong>,Phys. Rev. Lett. 83, 4987 (1999)[14] S. Nußmann, M. Hijlkema, B. Weber, F. Rohde, G. Rempe, and A. Kuhn, Submicr<strong>on</strong>Positi<strong>on</strong>ing of <str<strong>on</strong>g>Single</str<strong>on</strong>g> <str<strong>on</strong>g>Atoms</str<strong>on</strong>g> in a Microcavity, Phys. Rev. Lett. 95, 173602 (2005)[15] C. J. Hood, T. W. Lynn, A. C. Doherty, A. S. Parkins, and H. J. Kimble, TheAtom-<strong>Cavity</strong> Microscope: <str<strong>on</strong>g>Single</str<strong>on</strong>g> <str<strong>on</strong>g>Atoms</str<strong>on</strong>g> Bound in Orbit by <str<strong>on</strong>g>Single</str<strong>on</strong>g> Phot<strong>on</strong>s, Science287, 1447 (2000)[16] P. W. H. Pinkse, T. Fischer, P. Maunz, and G. Rempe, Trapping an atom with singlephot<strong>on</strong>s, Nature 404, 365 (2000)[17] D. P. DiVincenzo, Quantum Computati<strong>on</strong>, Science 270, 255 (1995)[18] D. Schrader, I. Dotsenko, M. Khudaverdyan, Y. Miroshnychenko, A. Rauschenbeutel,and D. Meschede, Neutral atom quantum register, Phys. Rev. Lett. 93, 150501(2004)[19] S. Kuhr, W. Alt, D. Schrader, I. Dotsenko, Y. Miroshnychenko, W. Rosenfeld,M. Khudaverdyan, V. Gomer, A. Rauschenbeutel, and D. Meschede, Coherenceproperties and quantum state transportati<strong>on</strong> in an optical c<strong>on</strong>veyor belt, Phys. Rev.Lett. 91, 213002 (2003)[20] H. Mabuchi, Q. A. Turchette, M. S. Chapman, and H. J. Kimble, Real-time detecti<strong>on</strong>of individual atoms falling through a high-finesse optical cavity, Opt. Lett. 21, 1393(1996)[21] P. Münstermann, T. Fischer, P. W. H. Pinkse, and G. Rempe, <str<strong>on</strong>g>Single</str<strong>on</strong>g> slow atomsfrom an atomic fountain observed in a high-finesse optical cavity, Opt. Commun.159, 63 (1999)[22] J. A. Sauer, K. M. Fortier, M. S. Chang, C. D. Hamley, and M. S. Chapman, <strong>Cavity</strong><strong>QED</strong> with optically transported atoms, Phys. Rev. A 69, 051804(R) (2004)[23] S. Kuhr, W. Alt, D. Schrader, M. Müller, V. Gomer, and D. Meschede, Deterministicdelivery of a single atom, Science 293, 278 (2001)[24] D. Schrader, S. Kuhr, W. Alt, M. Müller, V. Gomer, and D. Meschede, An opticalc<strong>on</strong>veyor belt <strong>for</strong> single neutral atoms, Appl. Phys. B 73, 819 (2001)


BIBLIOGRAPHY 129[25] I. Dotsenko, W. Alt, M. Khudaverdyan, S. Kuhr, D. Meschede, Y. Miroshnychenko,D. Schrader, and A. Rauschenbeutel, Submicrometer positi<strong>on</strong> c<strong>on</strong>trol ofsingle trapped neutral atoms, Phys. Rev. Lett. 95, 033002 (2005)[26] Y. Miroshnychenko, W. Alt, I. Dotsenko, L. Förster, M. Khudaverdyan,D. Meschede, D. Schrader, and A. Rauschenbeutel, An atom-sorting machine, Nature442, 151 (2006)[27] R. Hughes and T. Heinrichs, A Quantum In<strong>for</strong>mati<strong>on</strong> Science and TechnologyRoadmap, Neutral Atom Approaches to Quantum In<strong>for</strong>mati<strong>on</strong> Processing and QuantumComputing (2004), http://qist.lanl.gov/pdfs/neutral_atom.pdf[28] T. W. Hänsch and A. Schawlow, Cooling of Gases by Laser Radiati<strong>on</strong>, Opt. Commun.13, 68 (1975)[29] D. Wineland and H. Dehmelt, Proposed 10 14 ∆ν < ν Laser Fluorescence Spectroscopy<strong>on</strong> Tl + M<strong>on</strong>o-I<strong>on</strong> Oscillator III (side band cooling), Bull. Am. Phys. Soc. 20, 637(1975)[30] S. Chu, L. Hollberg, J. E. Bjorkholm, A. Cable, and A. Ashkin, Three-dimensi<strong>on</strong>alViscous C<strong>on</strong>finement and Cooling of <str<strong>on</strong>g>Atoms</str<strong>on</strong>g> by Res<strong>on</strong>ance Radiati<strong>on</strong> Pressure, Phys.Rev. Lett. 55, 48 (1985)[31] E. L. Raab, M. Prentiss, A. Cable, S. Chu, and D. E. Pritchard, Trapping of NeutralSodium <str<strong>on</strong>g>Atoms</str<strong>on</strong>g> with Radiati<strong>on</strong> Pressure, Phys. Rev. Lett. 59, 2631 (1987)[32] M. H. Anders<strong>on</strong>, J. R. Ensher, M. R. Matthews, C. E. Wieman, and E. A. Cornell,Observati<strong>on</strong> of Bose-Einstein c<strong>on</strong>densati<strong>on</strong> in a dilute atomic vapor, Science 269,198 (1995)[33] K. B. Davis, M. O. Mewes, M. R. Andrews, N. J. van Druten, D. S. Durfee, D. M.Kurn, and W. Ketterle, Bose-Einstein c<strong>on</strong>densati<strong>on</strong> in a gas of sodium atoms, Phys.Rev. Lett. 75, 3969 (1995)[34] Z. Hu and H. J. Kimble, Observati<strong>on</strong> of a single atom in a magnetooptical trap, Opt.Lett. 19, 1888 (1994)[35] F. Ruschewitz, D. Bettermann, J. L. Feng, and W. Ertmer, Statistical investigati<strong>on</strong>s<strong>on</strong> single trapped neutral atoms, Europhys. Lett. 34, 651 (1996)[36] D. Haubrich, H. Schadwinkel, F. Strauch, B. Ueberholz, R. Wynands, andD. Meschede, Observati<strong>on</strong> of individual neutral atoms in magnetic and magnetoopticaltraps, Europhys. Lett. 34, 663 (1996)[37] G. A. Askaryan, Effect of gradient of high-power electromagnetic field <strong>on</strong> electr<strong>on</strong>sand atoms, Sov. Phys. JETP 15, 1088 (1962)[38] R. Grimm, M. Weidemüller, and Y. B. Ovchinnikov, Optical dipole traps <strong>for</strong> neutralatoms, Adv. At. Mol. Opt. Phys. 42, 95 (2000)


130 BIBLIOGRAPHY[39] M. J. Lang and S. M. Block, Resource Letter: Laser-based optical tweezers, Am. J.Phys. 71, 201 (2003)[40] I. Bloch, Ultracold quantum gases in optical lattices, Nature Physics 1, 23 (2005)[41] D. Jaksch, Optical lattices, ultracold atoms and quantum in<strong>for</strong>mati<strong>on</strong> processing,C<strong>on</strong>temp. Phys. 45, 367–381 (2004)[42] Masao Takamoto, Feng-Lei H<strong>on</strong>g, Ryoichi Higashi, and Hidetoshi Katori, An opticallattice clock, Nature 435, 321 (2005)[43] G. Binnig and H. Rohrer, In touch with atoms, Rev. Mod. Phys. 71, 324 (1991)[44] W. Neuhauser, M. Hohenstatt, P. Toschek, and H. Dehmelt, Localized visible Ba +m<strong>on</strong>o-i<strong>on</strong> oscillator, Phys. Rev. A 22, 1137 (1980)[45] Y. Miroshnychenko, D. Schrader, S. Kuhr, W. Alt, I. Dotsenko, M. Khudaverdyan,A. Rauschenbeutel, and D. Meschede, C<strong>on</strong>tinued imaging of the transport of a singleneutral atom, Optics Express 11, 3498 (2003)[46] H. J. Metcalf and P. van der Straten, Laser Cooling and Trapping, Springer (1999)[47] V. I. Balykin, V. G. Minogin, and V. S. Letokhov, Electromagnetic trapping of coldatoms, Rep. Prog. Phys. 63, 1429–1510 (2000)[48] P. D. Lett, R. N. Watt, C. I. Westbrook, W. D. Phillips, P. L. Gould, and H. J.Metcalf, Observati<strong>on</strong> of <str<strong>on</strong>g>Atoms</str<strong>on</strong>g> Laser Cooled below the Doppler Limit, Phys. Rev.Lett. 61, 169 (1988)[49] J. Dalibard and C. Cohen-Tannoudji, Laser cooling below the Doppler limit by polarizati<strong>on</strong>gradients: simple theoretical models, J. Opt. Soc. Amer. B 6, 2023 (1989)[50] S. Kuhr, A c<strong>on</strong>trolled quantum system of individual neutral atoms, Ph.D. thesis, UniversitätB<strong>on</strong>n (2003), available at http://hss.ulb.uni-b<strong>on</strong>n.de/diss_<strong>on</strong>line/[51] W. Alt, Optical c<strong>on</strong>trol of single neutral atoms, Ph.D. thesis, Universität B<strong>on</strong>n(2004), available at http://hss.ulb.uni-b<strong>on</strong>n.de/diss_<strong>on</strong>line/[52] L. Ricci, M. Weidemüller, T. Esslinger, A. Hemmerich, C. Zimmermann, V. Vuletic,W. König, and T. W. Hänsch, A compact grating-stabilized diode laser system <strong>for</strong>atomic physics, Opt. Commun. 117, 541 (1995)[53] W. Alt, An objective lens <strong>for</strong> efficient fluoresence detecti<strong>on</strong> of single atoms, Optik113, 142 (2002)[54] J. P. Gord<strong>on</strong> and A. Ashkin, Moti<strong>on</strong> of atoms in a radiati<strong>on</strong> trap, Phys. Rev. A 21,1606 (1980)[55] D. Schrader, A neutral atom quantum register, Ph.D. thesis, Universität B<strong>on</strong>n(2004), available at http://hss.ulb.uni-b<strong>on</strong>n.de/diss_<strong>on</strong>line/


BIBLIOGRAPHY 131[56] L. Förster, W. Alt, I. Dotsenko, M. Khudaverdyan, D. Meschede, Y. Miroshnychenko,S. Reick, and A. Rauschenbeutel, Number-triggered loading and collisi<strong>on</strong>alredistributi<strong>on</strong> of neutral atoms in a standing wave dipole trap, New J. Phys. 8, 259(2006)[57] S. B. Hill and J. J. McClelland, <str<strong>on</strong>g>Atoms</str<strong>on</strong>g> <strong>on</strong> demand: Fast, deterministic producti<strong>on</strong>of single Cr atoms, Appl. Phys. Lett. 82, 3128 (2003)[58] N. Schlosser, G. Reym<strong>on</strong>d, I. Protsenko, and P. Grangier, Sub-poiss<strong>on</strong>ian loading ofsingle atoms in a microscopic dipole trap, Nature 411, 1024 (2001)[59] N. Schlosser, G. Reym<strong>on</strong>d, and P. Grangier, Collisi<strong>on</strong>al Blockade in MicroscopicOptical Dipole Traps, Phys. Rev. Lett. 89, 023005 (2002)[60] M. Khudaverdyan, W. Alt, I. Dotsenko, L. Förster, S. Kuhr, D. Meschede, Y. Miroshnychenko,D. Schrader, and A. Rauschenbeutel, Adiabatic quantum state manipulati<strong>on</strong>of single trapped atoms, Phys. Rev. A 71, 031404 (2004)[61] M. Khudaverdyan, Addressing of individual atoms in an optical dipole trap, Master’sthesis, Universität B<strong>on</strong>n (2003)[62] F. J. Giessibl, Advances in atomic <strong>for</strong>ce microscopy, Rev. Mod. Phys. 75, 949 (2003)[63] K. Suenaga, M. Tenc, C. Mory, C. Colliex, H. Kato, T. Okazaki, H. Shinohara,K. Hirahara, S. Bandow, and S. Iijima, Element-selective single atom imaging, Science290, 2280 (2000)[64] G. R. Guthöhrlein, M. Keller, K. Hayasaka, W. Lange, and H. Walther, A single i<strong>on</strong>as a nanoscopic probe of an optical field, Nature 414, 49 (2001)[65] A. B. Mundt, A. Kreuter, C. Becher, D. Leibfried, J. Eschner, F. Schmidt-Kaler,and R. Blatt, Coupling a <str<strong>on</strong>g>Single</str<strong>on</strong>g> Atomic Quantum Bit to a High Finesse Optical<strong>Cavity</strong>, Phys. Rev. Lett. 89, 103001 (2002)[66] D. Leibfried, B. DeMarco, V. Meyer, D. Lucas, M. Barrett, J. Britt<strong>on</strong>, W. M.Itano, B. Jelenkovi, C. Langer, T. Rosenband, and D. J. Wineland, Experimentaldem<strong>on</strong>strati<strong>on</strong> of robust, high-fidelity geometric two i<strong>on</strong>-qubit phase gate, Nature422, 412 (2003)[67] J. E. Thomas and L. J. Wang, Absolute frequency measurement of the In + clocktransiti<strong>on</strong> with a mode-locked laser, Phys. Rep. 262, 311 (1995)[68] S. Bergamini, B. Darquié, M. J<strong>on</strong>es, L. Jacubowiez, A. Browaeys, and P. Grangier,Holographic generati<strong>on</strong> of microtrap arrays <strong>for</strong> single atoms by use of a programmablephase modulator, J. Opt. Soc. Am. B 21, 1889 (2004)[69] J.-L. Starck, F. Murtagh, and A. Bijaoui, Image processing and data analysis, CambridgeUniversity Press (1998)


132 BIBLIOGRAPHY[70] M. Bertero and P. Boccacci, Introducti<strong>on</strong> to inverse problems in imaging, IOP PublishingLtd (1998)[71] W. Falk, Data reducti<strong>on</strong> from experimental histograms, Nucl. Instrum. and Methodsin Phys. Res. 220, 473 (1984)[72] O. Mandel, M. Greiner, A. Widera, T. Rom, T. W. Hänsch, and I. Bloch, C<strong>on</strong>trolledcollisi<strong>on</strong>s <strong>for</strong> multiparticle entanglement of optically trapped atoms, Nature 425, 937(2003)[73] D. Jaksch, H.-J. Briegel, J. I. Cirac, C. W. Gardiner, and P. Zoller, Entanglementof <str<strong>on</strong>g>Atoms</str<strong>on</strong>g> via Cold C<strong>on</strong>trolled Collisi<strong>on</strong>s, Phys. Rev. Lett. 82, 1975 (1999)[74] R. Raussendorf and H. J. Briegel, A One-Way Quantum Computer, Phys. Rev. Lett.86, 5188 (2001)[75] R. Scheunemann, F. S. Cataliotti, T. W. Hänsch, and M. Weitz, Resolving andaddressing atoms in individual sites of a CO 2 -laser optical lattice, Phys. Rev. A 62,51801 (2000)[76] Y. Miroshnychenko, W. Alt, I. Dotsenko, L. Förster, M. Khudaverdyan,A. Rauschenbeutel, and D. Meschede, Precisi<strong>on</strong> preparati<strong>on</strong> of strings of trappedneutral atoms, New J. Phys. 8, 191 (2006)[77] Y. Miroshnychenko, W. Alt, I. Dotsenko, L. Förster, M. Khudaverdyan,D. Meschede, S. Reick, and A. Rauschenbeutel, Inserting two atoms into a singleoptical micropotential, Phys. Rev. Lett. 97, 243003 (2006)[78] P. R. Berman, ed., <strong>Cavity</strong> Quantum Electrodynamics, Advances in atomic, molecular,and optical physics, Academic Press, San Diego (1994)[79] A. Siegman, Lasers, University Science Books, Cali<strong>for</strong>nia (1986)[80] Y. Miroshnychenko, Design and test of a high finesse res<strong>on</strong>ator <strong>for</strong> single atomexperiments, Master’s thesis, Universität B<strong>on</strong>n (2002)[81] W. Rosenfeld, A high finesse optical res<strong>on</strong>ator <strong>for</strong> cavity <strong>QED</strong> experiments, Master’sthesis, Universität B<strong>on</strong>n (2003)[82] C. J. Hood, Real-Time Measurement and Trapping of <str<strong>on</strong>g>Single</str<strong>on</strong>g> <str<strong>on</strong>g>Atoms</str<strong>on</strong>g> by <str<strong>on</strong>g>Single</str<strong>on</strong>g> Phot<strong>on</strong>s,Ph.D. thesis, Cali<strong>for</strong>nia Institute of Technology, CA (2000)[83] C. J. Hood, H. J. Kimble, and J. Ye, Characterizati<strong>on</strong> of high-finesse mirrors: Loss,phase shifts, and mode structure in an optical cavity, Phys. Rev. A 64, 33804 (2001)[84] D. Z. Anders<strong>on</strong>, J. C. Frish, and C. S. Masser, Mirror reflectometer based <strong>on</strong> opticalcavity decay time, Appl. Opt. 23, 1238 (1984)[85] G. Rempe, R. J. Thomps<strong>on</strong>, H. J. Kimble, and R. Lalezari, Measurement of ultralowlosses in an optical interferometer, Opt. Lett. 17, 363 (1992)


BIBLIOGRAPHY 133[86] T. W. Lynn, Measurement and C<strong>on</strong>trol of Individual Quanta in <strong>Cavity</strong> <strong>QED</strong>, Ph.D.thesis, Cali<strong>for</strong>nia Institute of Technology, CA (2003)[87] J. Poirs<strong>on</strong>, F. Bretenaker, M. Vallet, and A. L. Floch, Analytical and experimentalstudy of ringing effects in a Fabry-Perot cavity. Applicati<strong>on</strong> to the measurement ofhigh finesses, J. Opt. Soc. Am. B 14, 2811 (1997)[88] K. An, C. Yang, R. R. Dasari, and M. S. Feld, <strong>Cavity</strong> ring-down technique and itsapplicati<strong>on</strong> to the measurement of ultraslow velocities, Opt. Lett. 20, 1068 (1995)[89] R. Drever and J. Hall, Laser phase and frequency stabilizati<strong>on</strong> using an optical res<strong>on</strong>ator,Appl. Phys. B 31, 97 (1983)[90] H. Mabuchi, J. Ye, and H. J. Kimble, Full observati<strong>on</strong> of single-atom dynamics incavity <strong>QED</strong>, Appl. Phys. B 68, 1095 (1999)[91] C. Wieman and T. W. Hänsch, Doppler-Free Laser Polarizati<strong>on</strong> Spectroscopy, Phys.Rev. Lett. 36, 1170 (1976)[92] C. P. Pearman, C. S. Adams, S. G. Cox, P. F. Griffin, D. A. Smith, and I. G.Hughes, Polarizati<strong>on</strong> spectroscopy of a closed atomic transiti<strong>on</strong>: applicati<strong>on</strong>s to laserfrequency locking, J. Phys. B: At. Mol. Opt. Phys. 35, 5141 (2002)[93] J. R. Buck, <strong>Cavity</strong> <strong>QED</strong> in Microsphere and Fabry-Perot Cavities, Ph.D. thesis,Cali<strong>for</strong>nia Institute of Technology, CA (2000)[94] H. Carmichael, An Open Systems Approach to Quantum Optics, Springer, Berlin(1993)[95] S. M. Dutra, <strong>Cavity</strong> Quantum Electrodynamics, Wiley-Intersceince (2005)[96] C. Cohen-Tannoudji, J. Dup<strong>on</strong>t-Roc, and G. Grynberg, Atom-Phot<strong>on</strong> Interacti<strong>on</strong>s:Basic Processes and Applicati<strong>on</strong>s, Wiley, New York (1992)[97] S. Haroche and J.-M. Raim<strong>on</strong>d, Exploring the Quantum, Ox<strong>for</strong>d University Press(2006)[98] M. O. Scully and M. S. Zubairy, Quantum Optics, Cambridge University Press,Cambridge (1997)[99] S. M. Tan, A computati<strong>on</strong>al toolbox <strong>for</strong> quantum and atom optics, J. Opt. B: Quant.Semiclass. Opt. 1, 424 (1999)[100] I. Dotsenko, Raman spectroscopy of single atoms, Master’s thesis, Universität B<strong>on</strong>n(2002)[101] M. E. Gehm, K. M. O’Hara, T. A. Savard, and J. E. Thomas, Dynamics of noiseinducedheating in atom traps, Phys. Rev. A 58, 3914 (1998)


134 BIBLIOGRAPHY[102] J. McKeever, J. R. Buck, A. D. Boozer, A. Kuzmich, H.-C. Nägerl, D. M. Stamper-Kurn, and H. J. Kimble, State-Insensitive Trapping of <str<strong>on</strong>g>Single</str<strong>on</strong>g> <str<strong>on</strong>g>Atoms</str<strong>on</strong>g> in an Optical<strong>Cavity</strong>, Phys. Rev. Lett. 90, 133602 (2003)[103] T. Puppe, I. Schuster, A. Grothe, A. Kubanek, K. Murr, P.W.H. Pinkse,and G. Rempe, Trapping and observing single atoms in the dark, arXiv:quantph/0702162(2007)[104] H. Perrin, A. Kuhn, I. Bouchoule, and C. Salom<strong>on</strong>, Sideband cooling of neutral atomsin a far-detuned optical lattice, Europhys. Lett. 42, 395 (1998)[105] S. E. Hamann, D. L. Haycock, G. Klose, P. H. Pax, I. H. Deutsch, and P. S. Jessen,Resolved Sideband Raman Cooling to the Ground State of an Optical Lattice, Phys.Rev. Lett. 80, 4149 (1998)[106] I. Dotsenko, W. Alt, S. Kuhr, D. Schrader, M. Müller, Y. Miroshnychenko,V. Gomer, A. Rauschenbeutel, and D. Meschede, Applicati<strong>on</strong> of electro-opticallygenerated light fields <strong>for</strong> Raman spectroscopy of trapped cesium atoms, Appl. Phys.B 78, 711 (2004)[107] P. Domokos and H. Ritsch, Mechanical effects of light in optical res<strong>on</strong>ators, JOSAB 20, 1098–1130 (2003)[108] P. Horak, G. Hechenblaikner, K. M. Gheri, H. Stecher, and H. Ritsch, <strong>Cavity</strong>-InducedAtom Cooling in the Str<strong>on</strong>g Coupling Regime, Phys. Rev. Lett. 79, 4974 (1997)[109] P. Maunz, T. Puppe, I. Schuster, N. Syassen, P. W. H. Pinkse, and G. Rempe,<strong>Cavity</strong> cooling of a single atom, Nature 428, 50 (2004)[110] L. You, X. X. Yi, and X. H. Su, Quantum logic between atoms inside a high-Q opticalcavity, Phys. Rev. A 67, 032308 (2003)[111] C. Marr, A. Beige, and G. Rempe, Entangled-state preparati<strong>on</strong> via dissipati<strong>on</strong>assistedadiabatic passages, Phys. Rev. A 68, 033817 (2003)[112] A. S. Sørensen and K. Mølmer, Measurement Induced Entanglement and QuantumComputati<strong>on</strong> with <str<strong>on</strong>g>Atoms</str<strong>on</strong>g> in Optical Cavities, Phys. Rev. Lett. 91, 097905 (2003)[113] J. Metz, M. Trupke, and A. Beige, Robust Entanglement through Macroscopic QuantumJumps, Phys. Rev. Lett. 97, 040503 (2006)[114] D. A. Steck, Cesium D Line Data (23 January 1998, Revisi<strong>on</strong> 1.5, 21 November2002), http://steck.us/alkalidata/


Publicati<strong>on</strong>s1. S. Kuhr, W. Alt, D. Schrader, I. Dotsenko, Y. Miroshnychenko, W. Rosenfeld,M. Khudaverdyan, V. Gomer, A. Rauschenbeutel, and D. Meschede, Coherenceproperties and quantum state transportati<strong>on</strong> in an optical c<strong>on</strong>veyor belt,Phys. Rev. Lett. 91, 213002 (2003)2. Y. Miroshnychenko, D. Schrader, S. Kuhr, W. Alt, I. Dotsenko, M. Khudaverdyan,A. Rauschenbeutel, and D. Meschede, C<strong>on</strong>tinued imaging of the transport of a singleneutral atom, Optics Express 11, 3498 (2003)3. I. Dotsenko, W. Alt, S. Kuhr, D. Schrader, M. Müller, Y. Miroshnychenko, V. Gomer,A. Rauschenbeutel, and D. Meschede, Applicati<strong>on</strong> of electro-optically generated lightfields <strong>for</strong> Raman spectroscopy of trapped Cesium atoms, Appl. Phys. B 78, 711 (2004)4. D. Schrader, I. Dotsenko, M. Khudaverdyan, Y. Miroshnychenko, A. Rauschenbeutel,and D. Meschede, A neutral atom quantum register, Phys. Rev. Lett. 93, 150501(2004)5. M. Khudaverdyan, W. Alt, I. Dotsenko, L. Förster, S. Kuhr, D. Meschede, Y. Miroshnychenko,D. Schrader, and A. Rauschenbeutel, Adiabatic quantum state manipulati<strong>on</strong>of single trapped atoms, Phys. Rev. A 71, 031404(R) (2005)6. I. Dotsenko, W. Alt, M. Khudaverdyan, S. Kuhr, D. Meschede, Y. Miroshnychenko,D. Schrader, and A. Rauschenbeutel, Submicrometer positi<strong>on</strong> c<strong>on</strong>trol ofsingle trapped neutral atoms, Phys. Rev. Lett. 95, 033002 (2005)7. S. Kuhr, W. Alt, D. Schrader, I. Dotsenko, Y. Miroshnychenko, V. Gomer,A. Rauschenbeutel, and D. Meschede, Analysis of dephasing mechanisms in a standingwave dipole trap, Phys. Rev. A 72, 023406 (2005)8. Y. Miroshnychenko, W. Alt, I. Dotsenko, L. Frster, M. Khudaverdyan, D. Meschede,D. Schrader, and A. Rauschenbeutel, An atom-sorting machine, Nature 442, 151(2006)9. Y. Miroshnychenko, W. Alt, I. Dotsenko, L. Förster, M. Khudaverdyan,D. Meschede, S. Reick, and A. Rauschenbeutel, Inserting two atoms into a singleoptical micropotential, Phys. Rev. Lett. 97, 243003 (2006)135


136 Publicati<strong>on</strong>s10. L. Förster, W. Alt, I. Dotsenko, M. Khudaverdyan, Y. Miroshnychenko,D. Meschede, S. Reick, and A. Rauschenbeutel, Number-triggered loading and collisi<strong>on</strong>alredistributi<strong>on</strong> of neutral atoms in a standing wave dipole trap, New J. Phys. 8,259 (2006)


AcknowledgementsMany people deserve thanks <strong>for</strong> their help and support during my research in B<strong>on</strong>n.First and <strong>for</strong>emost, I would like to thank my supervisor, Dieter Meschede, <strong>for</strong> providingme the opportunity to work <strong>on</strong> this exciting “single-atoms” experiment and to be apart of his research group. I am also grateful <strong>for</strong> his being a coordinator of the B<strong>on</strong>nInternati<strong>on</strong>al Physics Programme, which gave me a chance to start my scientific career atthe University of B<strong>on</strong>n. I would like to thank Le<strong>on</strong>id Yatsenko <strong>for</strong> his accepting to reviewmy thesis.I want to sincerely thank our post-docs, Arno Rauschenbeutel and Wolfgang Alt,<strong>for</strong> their many years of guidance and support, <strong>for</strong> their supernatural feeling of theexperiment, and their detailed answer to any questi<strong>on</strong>. I have particularly benefited fromworking al<strong>on</strong>gside our senior generati<strong>on</strong> of the PhD students: Stefan Kuhr, Wolfgang Alt,and Dominik Schrader. I acknowledge them <strong>for</strong> the inspiring atmosphere they createdwhile working <strong>on</strong> the experiment with me, and <strong>for</strong> immeasurable amount of knowledge,both experimental and theoretical, I have learned from them.I am grateful to have been a part of “ukrainian mafia” in the institute together withYevhen Miroshnychenko, Wenja Rosenfeld, and Mika Khudaverdyan. The experiments <strong>on</strong>positi<strong>on</strong> c<strong>on</strong>trol would be hardly possibly without Yevhen and Mika. My special thanksgoes to Wenja <strong>for</strong> introducing me the cavity stabilizati<strong>on</strong> – a large net of wires, BNCcables, and aluminium “black boxes”. I appreciate working with Le<strong>on</strong>id Förster duringour experiments <strong>on</strong> juggling with single atoms.The cavity results of my thesis could become possible owing to a coordinated teamwork and a friendly climate reigned in our lab. I thank Mika <strong>for</strong> his endless patienceand optimism, even during the night measurements, when we tried to recognize andimagine the cavity mode in each noisy peak generated by the experiment <strong>on</strong> random.Who would ever have thought, that over more than a week Igor and Mika would look <strong>for</strong>the cavity <strong>on</strong> the wr<strong>on</strong>g site from the MOT? I also deeply appreciate Sebastian Reick<strong>for</strong> his organizati<strong>on</strong> skills, which helped us a lot while opening the vacuum chamber,installing our “baby” inside the glass cell, or simply planning next measurements. Finally,I am thankful to Ariane Stiebeiner <strong>for</strong> her ability to quickly take the hint of what I amtrying to explain during both numerous night measurements and the lab rec<strong>on</strong>structi<strong>on</strong>.I deeply appreciate criticism and valuable suggesti<strong>on</strong>s of Sebastian, Mika, Wolfgang, and137


138 AcknowledgementsArno while reading the manuscript of my thesis.Of course, I would like to thank all other members of the group of Prof. Meschedethat I have been <strong>for</strong>tunate to have c<strong>on</strong>stant interacti<strong>on</strong> with during my time here.Special acknowledgement goes to our “support team” – Annelise Miglo, Il<strong>on</strong>a Jaschke,Fien Latumahina, and Dietmar Haubrich, <strong>for</strong> their “parental care” of all of us, <strong>for</strong> theirprotecting us in the cruel and intricate world of bureaucracy, <strong>for</strong> providing us a carelessplace called “the Institute of Applied Physics”, where we could c<strong>on</strong>centrate <strong>on</strong> “buildinga quantum computer”.And finally, last but not least, I would like to thank my darling wife, Ira, and ourdaughter, Aly<strong>on</strong>a, <strong>for</strong> their c<strong>on</strong>tinuous love, support, and patience during these years.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!