Matter-Wave Self-Imaging by Atomic Center-of- Mass Motion Induced Interference

Similar documents
Raman Amplification of Matter Waves

cond-mat/ v2 16 Aug 1999

The phonon dispersion relation of a Bose-Einstein condensate

Absorption-Amplification Response with or Without Spontaneously Generated Coherence in a Coherent Four-Level Atomic Medium

In Situ Imaging of Cold Atomic Gases

Quantum optics of many-body systems

B 2 P 2, which implies that g B should be

Collective atomic recoil in a moving Bose-Einstein condensate: From superradiance to Bragg scattering

K. Campbell, Gretchen, E. Leanhardt, Aaron, Mun, Jongchul, Boyd, Micah, W. Streed, Erik, Ketterle, Wolfgang, Pritchard, David E.

Ultracold atoms and molecules

Optimization of transfer of laser-cooled atom cloud to a quadrupole magnetic trap

Sequential Superradiant Scattering from Atomic Bose Einstein Condensates

Dmitriy Churin. Designing high power single frequency fiber lasers

SUB-NATURAL-WIDTH N-RESONANCES OBSERVED IN LARGE FREQUENCY INTERVAL

Precision Interferometry with a Bose-Einstein Condensate. Cass Sackett. Research Talk 17 October 2008

Atomic Coherent Trapping and Properties of Trapped Atom

Laser Cooling and Trapping of Atoms

Quantum correlations and atomic speckle

HYPERFINE STRUCTURE CONSTANTS IN THE 102D3/2 AND 112D 3/2 STATES OF 85Rb M. GLOW

COPYRIGHTED MATERIAL. Index

Optical Lattices. Chapter Polarization

SYRTE - IACI. AtoM Interferometry dual Gravi- GradiOmeter AMIGGO. from capability demonstrations in laboratory to space missions

Lecture 4: Superfluidity

Atom interferometry. Quantum metrology and fundamental constants. Laboratoire de physique des lasers, CNRS-Université Paris Nord

Workshop on Topics in Quantum Turbulence March Experiments on Bose Condensates

Two-photon nonlinearity in general cavity QED systems

Interferometry and precision measurements with Bose-condensed atoms

Physics 598 ESM Term Paper Giant vortices in rapidly rotating Bose-Einstein condensates

Atom Microscopy via Dual Resonant Superposition

Ryytty, Pasi & Kaivola, Matti Pulsed standing-wave mirror for neutral atoms and molecules

6. Interference of BECs

David McIntyre. A slow beam of laser cooled rubidium atoms will be used as the matter-wave source. The atom

Lecture 3. Bose-Einstein condensation Ultracold molecules

The phases of matter familiar for us from everyday life are: solid, liquid, gas and plasma (e.f. flames of fire). There are, however, many other

Large Momentum Beamsplitter using Bloch Oscillations

NanoKelvin Quantum Engineering

Matter waves in time-modulated complex light potentials

UNIVERSITY OF SOUTHAMPTON

Manipulation of single neutral atoms in optical lattices

NON LINEAR PULSE EVOLUTION IN SEEDED AND CASCADED FELS

High-Speed Quadratic Electrooptic Nonlinearity in dc-biased InP

Temporal modulation instabilities of counterpropagating waves in a finite dispersive Kerr medium. II. Application to Fabry Perot cavities

Part I. Principles and techniques

PHYSICAL REVIEW LETTERS

Shau-Yu Lan 藍劭宇. University of California, Berkeley Department of Physics

Saturation Absorption Spectroscopy of Rubidium Atom

Quantum entanglement and light propagation through Bose-Einstein condensate (BEC) M. Emre Taşgın

Γ43 γ. Pump Γ31 Γ32 Γ42 Γ41

Development of the Zeeman Slower for the Ultra-cold Atomic Interference Experiment

5. Gross-Pitaevskii theory

Optical Multi-wave Mixing Process Based on Electromagnetically Induced Transparency

Bose-Einstein condensates in optical lattices

Supplementary Figure 1 Level structure of a doubly charged QDM (a) PL bias map acquired under 90 nw non-resonant excitation at 860 nm.

High Resolution Laser Spectroscopy of Cesium Vapor Layers with Nanometric Thickness

A Study of Matter-Wave Diffraction in Bose- Einstein Condensates for Atom Interferometry Applications

Dark pulses for resonant two-photon transitions

Multipath Interferometer on an AtomChip. Francesco Saverio Cataliotti

Study on positronium Bose-Einstein condensation

arxiv:physics/ v2 [physics.atom-ph] 12 Dec 2005

High-resolution hyperfine spectroscopy of excited states using electromagnetically induced transparency

Laser stabilization via saturated absorption spectroscopy of iodine for applications in laser cooling and Bose-Einstein condensate creation

Formation of Narrow Optical Resonance by Micrometer Thin Rb- Vapor Layer

Building Blocks for Quantum Computing Part IV. Design and Construction of the Trapped Ion Quantum Computer (TIQC)

Conference on Research Frontiers in Ultra-Cold Atoms. 4-8 May Generation of a synthetic vector potential in ultracold neutral Rubidium

Transit time broadening contribution to the linear evanescent susceptibility

PROGRESS TOWARDS CONSTRUCTION OF A FERMION ATOMIC CLOCK FOR NASA S DEEP SPACE NETWORK

Atomic filter based on stimulated Raman transition at the rubidium D1 line

Observation of large continuous-wave two-photon optical amplification

4 FEL Physics. Technical Synopsis

Ion traps. Trapping of charged particles in electromagnetic. Laser cooling, sympathetic cooling, optical clocks

Early time dynamics of strongly coupled ultracold neutral Ca + and Ca 2+ plasmas

Introduction to Classical and Quantum FEL Theory R. Bonifacio University of Milano and INFN LNF

Controlling Spin Exchange Interactions of Ultracold Atoms in Optical Lattices

The amazing story of Laser Cooling and Trapping

arxiv:quant-ph/ v3 17 Nov 2003

PROGRESS TOWARDS CONSTRUCTION OF A FERMIONIC ATOMIC CLOCK FOR NASA S DEEP SPACE NETWORK

arxiv: v1 [physics.ins-det] 25 May 2017

Vectorial structure and beam quality of vector-vortex Bessel Gauss beams in the far field

Laser cooling of 173 Yb for isotope separation and precision hyperfine spectroscopy

Observation of Saturation of Fidelity Decay with an Atom Interferometer

arxiv:quant-ph/ v1 17 Mar 2005

STIMULATED RAMAN ATOM-MOLECULE CONVERSION IN A BOSE-EINSTEIN CONDENSATE. Chisinau, Republic of Moldova. (Received 15 December 2006) 1.

Supplementary information for Quantum delayed-choice experiment with a beam splitter in a quantum superposition

Observation of spectral enhancement in a soliton fiber laser with fiber Bragg grating

Nonreciprocal Bloch Oscillations in Magneto-Optic Waveguide Arrays

gives rise to multitude of four-wave-mixing phenomena which are of great

Differential Brillouin gain for improving the temperature accuracy and spatial resolution in a long-distance distributed fiber sensor

Experimental Study of Quantum Chaos with Cold Atoms

Confining ultracold atoms on a ring in reduced dimensions

Optical manipulation of molecular ensembles in the gas phase

Ytterbium quantum gases in Florence

Chapter 7: Quantum Statistics

Towards compact transportable atom-interferometric inertial sensors

Simultaneous cooling of axial vibrational modes in a linear ion trap

pulses. Sec. III contains the simulated results of the interaction process and their analysis, followed by conclusions in Sec. IV.

Optical time-domain differentiation based on intensive differential group delay

Introduction to Atomic Physics and Quantum Optics

Recollision processes in strong-field QED

Model Answer (Paper code: AR-7112) M. Sc. (Physics) IV Semester Paper I: Laser Physics and Spectroscopy

Raman-Induced Oscillation Between an Atomic and Molecular Gas

Transcription:

Georgia Southern University From the SelectedWorks of Marvin G. Payne December 18, 2008 Matter-Wave Self-Imaging by Atomic Center-of- Mass Motion Induced Interference Ke Li L. Deng E. W. Hagley Marvin G. Payne, Georgia Southern University M. S. Zhan, Chinese Academy of Sciences Available at: https://works.bepress.com/marvin_payne/3/

Matter-Wave Self-Imaging by Atomic Center-of-Mass Motion Induced Interference Ke Li, 1,2,3 L. Deng, 1 E. W. Hagley, 1 M. G. Payne, 1 and M. S. Zhan 2,3 1 Physics Laboratory, National Institute of Standards and Technology, Gaithersburg, Maryland 20899, USA 2 State Key Laboratory of Magnetic Resonance, Atomic and Molecular Physics, Wuhan Institute of Physics and Mathematics, Chinese Academy of Sciences, Wuhan 430071, China 3 Center for Cold Atom Physics, Chinese Academy of Sciences, Wuhan 430071, China (Received 25 August 2008; published 18 December 2008) We demonstrate matter-wave self-imaging resulting from atomic center-of-mass motion-based interference. We show that non-negligible atomic center-of-mass motion and an instantaneous Doppler shift can drastically change the condensate momentum distribution, resulting in a periodic collapse and the recurrence of condensate diffraction probability as a function of the stationary light-field pulsing time. The observed matter-wave self-imaging is characterized by an atomic center-of-mass motion induced population amplitude interference in the presence of the light field that simultaneously minimizes all high (n 1) diffraction orders and maximizes the zeroth diffraction component. DOI: 10.1103/PhysRevLett.101.250401 Deflection and diffraction of atoms by light is a research area that can be traced back more than 40 years to early pioneering theoretical and experimental studies [1 11]. The recent development of laser cooling and trapping technology, and the realization of Bose-Einstein condensates (BECs), have greatly advanced the research activities of coherent atom optics, a rapidly developing new research area of modern atomic physics. In comparison with thermal atomic sources, atomic BECs as atom-optical sources have the advantages of high phase space density and high spatial localization, and are therefore well suited for high resolution time-domain matter-wave diffraction studies. Indeed, various groups have demonstrated Bragg diffraction of condensates [12,13], pulsed standing-wave diffraction of a condensate [14], the time-domain matter-wave Talbot effect [15], condensate coherence-time measurements [16], Bragg-diffraction-based quasicontinuous atom lasers [17], and coherent matter-wave mixing [18]. Diffraction of ultracold atoms by optical standing waves usually can be classified into two time regimes. In the Raman-Nath regime the optical standing wave is pulsed for a very short time so that the atom center-of-mass (c.m.) motion and instantaneous Doppler frequency shift can be neglected [2,14 16]. The characteristic bidirectional momentum distribution of the diffracted ultracold atoms and BEC components is well described by Bessel functions evenly distributed on both sides of the zero momentum component. In the Bragg regime, the optical grating is pulsed for a much longer time, and the stringent phase matching condition results in a unidirectional diffraction pattern. In this Letter using an optical standing wave formed with counterpropagating beams of significant intensity difference we demonstrate atomic c.m. motion-based bidirectional, high-order matter-wave self-imaging and condensate momentum oscillation. This phenomenon exists in the non-raman-nath regime that is also far away from the PACS numbers: 03.75. b, 42.25.Fx, 67.85.Hj typical Bragg regime. We further note that the matter-wave self-imaging effect reported here is not related to the temporal, matter-wave Talbot effect reported previously [15]. To the best of our knowledge, such a full matterwave self-imaging resulting from atomic c.m. motion induced interference has never been demonstrated before. The experiment reported here was performed with a 87 Rb condensate in the jf ¼ 2;m F ¼ 2i hyperfine ground state. We first collect 10 8 atoms in a double magnetooptical trap and transfer the cold atoms into a quadrupole-ioffe-configuration trap for further evaporative cooling, producing an almost pure condensate that typically contains 10 5 atoms. We expand the condensate for 3ms[19] and then illuminate the condensate with a pair of counterpropagating laser pulses of selected pulse duration and uneven intensities. In our setup the pulsed standingwave light fields are applied along the long dimension of the condensate (Fig. 1). The strong pump field (wave vector ^k 1 ¼ ^k L, I 1 215 mw=cm 2 ) propagates in the þ ^x direction, whereas a much weaker pump field (wave vector ^k 2 ¼ ^k L, I 2 14 mw=cm 2 ) propagates collinearly in the ^x direction. Both fields come from the same laser (wavelength L ¼ 780 nm) that is linearly polarized in ^z direction, detuned ¼ 2ð 1:5Þ GHz below the F 0 ¼ 3, m F 0 ¼ 2 state [20], and locked to a saturation absorption cell maintained under constant temperature. Immediately after irradiation by the pump laser pulses we allow 30 ms of ballistic expansion of the condensate before imaging its momentum distribution with on-resonance light. In Fig. 2 we show a series of images of condensate momentum distribution for different standing-wave pulse durations. In obtaining these images, we have used optical powers of 20 and 1.3 mw for the two pump fields (beam radius r 1=e 1:7 mm), respectively. The images are taken after a 30 ms time of flight. With a 1 s pump pulse the condensate momentum distribution has only the p ¼ 0@k L 0031-9007=08=101(25)=250401(4) 250401-1 Ó 2008 The American Physical Society

FIG. 1 (color online). (a) Schematics of experiment setup and directions of laser propagation. (b) A typical diffraction image obtained after about 30 ms time of flight. (c) Schematic energylevel diagram and laser couplings used in deriving Eq. (1). component. After a 6 s pulse more condensate components of nonzero momenta have developed and we observe the depletion of the 0@k L component. After an 8 s pulse the zero momentum component is repopulated. At 25 s, however, populations in all high-order momentum components are very small in comparison with the stationary condensate (0@k L ). That is, the condensate momentum distribution has collapsed completely and matter-wave self-imaging is achieved. As the standing-wave pulse time is further increased, the multiorder bidirectional condensate momentum diffraction distribution reappear. At FIG. 2 (color online). Matter-wave self-imaging and condensate momentum distribution collapse and recurrence in a pulsed optical standing wave. Field of view: 2:7 mm 0:63 mm (see text for other parameters) and condensate time of flight is typically 30 ms. 250401-2 40 s the bidirectional high-order momentum distributions are clearly seen. To explain the observed condensate momentum distribution collapse, recurrence, and matter-wave self-imaging observed near P 25 s, we consider an atomic system that includes two electronic states and n momentum states [Fig. 1(c)] [20]. In the rest frame of an atom each twophoton process that involves absorption of one photon from one beam and emission of one photon to the other beam leads to the atom, which has mass M and was initially at rest, acquiring a velocity V r ¼2@k L =M, which results in a two-photon recoil energy E r =@ ¼ ¼ 2@k 2 L =M ¼ 8 r, where r ¼ @k 2 L =ð4mþ is the onephoton recoil frequency. The atomic c.m. motion thus leads to an instantaneous Doppler shift in angular frequency! Doppler ¼ 2ð2n@k L Þ! L =ðmcþ ¼4n@k 2 L =M ¼ 16n r (n ¼ 0; 1; 2;...). The equation of motion of the probability amplitude of the nth momentum state of the atomic ground electronic state wave function for such a standing wave formed by counterpropagating beams of significant intensity difference can be written as [21] (see discussion on asymmetry later) nðp; tþ ¼iZ½ n 1e i t þ nþ1e iþt Š; (1) where ¼! Doppler ¼16 r ðn 1 2Þ. The real quantity Z ¼ 1 2 = is the two-photon Rabi frequency, qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi j ¼ I j =ð2i s Þ (with ¼ 2 6:07 MHz and I s 2:5 mw=cm 2 for the transitions used) is the forward (j ¼ 1) (backward, j ¼ 2) pump field Rabi frequency, is one-photon detuning, r ¼ 3:75 khz is rubidium atom one-photon recoil frequency. The exponential factors in Eq. (1) include the contributions from atomic c.m. motion. In the Raman-Nath regime, the pulsed optical grating is considered to be instantaneous and atomic motion is neglected. Thus, all exponential factors are replaced by unity. Consequently, the momentum distribution of the condensate components is characterized by the Bessel functions of order n [1 11]. When the optical grating is pulsed for an appreciable time, however, deviation from the Raman- Nath approximation becomes important and the exponential factors in Eq. (1) must be included. This results in nonnegligible corrections due to atomic c.m. motion. We show below that this correction is at the root of the observed condensate momentum oscillation and full matter-wave self-imaging with a period that is different from the fractional and integral Talbot time reported previously [15]. In general, Eq. (1) cannot be solved analytically. The physical origin of the observed condensate self-imaging, however, can be understood by rearranging Eq. (1) as nðp; tþ ¼iZ½ n 1 þ nþ1šþiz½ n 1ðe i t 1Þ þ nþ1ðe iþt 1ÞŠ: (2) It can be shown that in the Raman-Nath limit where the second square bracket, which represents the contribution from the atomic c.m. motion, is neglected, Eq. (2) yields

the result for a pulse time P, nðp; P Þ¼e ix i n J n ðxþ n ; (3) p where x ¼ Z P =2, n ¼ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ðp þ 2n@k L Þ and the Dirac function enforces the even-photon processes. Thus, the population of the nth-order diffracted component R (corresponds to 2n@k L momentum) is P n ¼ þ1 1 dpj nðp; P Þj 2 ¼½J n ðxþš 2 as it should be. When the atomic c.m. motion is included the second square bracket in Eq. (2) leads to a solution of the atomic ground state population amplitude of the form nðp; P Þ¼e ix i n ½ n J n ðxþþg n ð; x; nþmþš; (4) nþmþ, where where ¼ r =Z. We note that G n ð; x; jmj ðnþ1þ, is a complex function containing the sum (over integer m) of integrals of high-order Bessel functions. Although the exact analytic expression of G n ð; x; nþmþ cannot be obtained, its physical implication is clear and sufficient for one to understand the experimental observations. In the Raman-Nath limit, the condensate momentum distribution oscillations in the time domain are solely determined by the minima of the Bessel functions given in Eq. (3). Based on the properties of Bessel functions one concludes that full matter-wave self-imaging cannot occur in the Raman-Nath regime [see Fig. 3(c) and discussion below). In the non-raman-nath regime, where Eq. (2) applies, the non-negligible atomic c.m. motion leads to an interference, and the collapse and revival of the condensate momentum distribution are now determined by the minima of the quantity in the square bracket in Eq. (4). The physical meaning of the function G n ð; x; nþmþ is that it modifies and shifts minima of the diffraction amplitudes as a function of pulse duration, and therefore permits the possibility of simultaneously achieving minima of all n 1 order diffraction amplitudes. We emphasize that simultaneously achieving minima of all high-order amplitudes is a prerequisite for a full matter-wave self-imaging where the population distribution collapses completely to a condensate of zero momentum (0@k L ). The observed full matter-wave self-imaging is different from the temporal matter-wave Talbot effect reported before [15]. In the temporal matter-wave Talbot effect, two successive optical gratings are pulsed on for such a short time ( P 100 ns) that the effect of the atomic c.m. motion is totally negligible during the grating irradiation time. The condensate self-imaging is achieved by the alignment and antialignment of the atomic optical Bloch vector which precesses in the absence of any external field. We have numerically integrated Eq. (1) to obtain the diffraction probabilities for the first few significant diffraction orders. Figure 3(a) shows that at P 23 s the zeroth-order diffraction probability is near 100%, whereas all high orders are very small. This agrees well with the 250401-3 FIG. 3. (a) Probabilities of nth-order diffraction as a function of pulse duration. The numerics were done with 3 n þ3 and C 4 ¼ C 4 ¼ 0 (calculation with n ¼5 have yielded the same results). Parameters are chosen so that 1 2 =ð2 r Þ¼ 25, which is very close to the experimental value of 22. The matter-wave self-imaging occurs at t ¼ P 23 s (see arrow). Solid curve, n ¼ 0; dash-dotted curve, n ¼ 1; dashed curve, n ¼ 2; dotted curve (not visible), n ¼ 3. (b) Probabilities of nth-order diffraction as a function of pulse duration for 1 2 =ð2 r Þ¼ 34. The first full matter-wave self-imaging point occurs at t ¼ P 56 s and partial collapses of momentum distribution can be seen [curve styles are the same as in (a)]. We note that the full matter-wave self-imaging effect due to atomic c.m. motion is characterized only by an interference that simultaneously minimizes all n 1 diffraction orders and maximizes the zeroth diffraction component. (c) Raman-Nath limit diffraction probability P n ¼jJ n ðxþj 2 with n ¼ 0;...; 3 as functions of dimensionless argument x. Note that jj 0 ðxþj 2 reaches a maximum again when jj 1 ðxþj 2 becomes zero (see the arrows), indicating the simultaneous disappearance of the first-order diffraction and repopulation of the zeroth-order diffraction even when the atomic c.m. motion is neglected. At the same x, however,n> 1 diffraction components are significant. observed time of P ¼ 25 s when all n 1 orders are very small. Figure 3(b) shows the numerically calculated diffraction probabilities for 1 2 =ð2 r Þ¼34. It is interesting to notice that the first full matter-wave self-imaging time moves to P 56 s. In addition, partial collapse of selected diffraction orders, a phenomenon that is well known in near-field imaging, are seen before the first full self-imaging is achieved. This indicates that the twophoton Rabi frequency can shift the time location of the first full matter-wave self-imaging, raising the possibility of coherently controlled and time-tunable full matter-wave self-imaging [22]. It is instructive to explain the key enabler of the matterwave self-imaging effect. First, we point out that the full matter-wave self-imaging stems out of the non-negligible atomic c.m. motion in the non-raman-nath regime.

Second, observation of the simultaneous disappearance of the first-order diffraction and the repopulation of the zeroth-order diffraction does not lead to the conclusion of matter-wave self-imaging. In fact, one can show [Fig. 3(c)] that even in the Raman-Nath regime the disappearance of the first-order diffraction indeed coincides with the repopulation of the zeroth-order diffraction. This is due to the properties of the Bessel functions, not the atomic c.m. motion induced interference. However, the higher (n>1) diffraction orders in this regime are not small, indicating the lack of matter-wave self-imaging. We emphasize that the key indicator of the full matter-wave self-imaging is the simultaneous minimization of all higher orders (not just n ¼ 1 order) and maximization of the zeroth-order. This indicator is clearly shown in our work [23]. For this reason we believe that we have demonstrated for the first time a full matter-wave self-imaging effect. Finally, we comment on the apparent asymmetry exhibited in the images shown in Fig. 2. In this work, we study momentum oscillation of a condensate as a function of atomic center-of-mass motion in a pulsed optical grating formed by counterpropagating beams with a significant intensity difference propagating along the long axis of the condensate. These two new features along with the beam propagation geometry have profound consequences. At certain interaction time we have observed that high-order components are stronger than the lower order diffraction components. This is to be expected as in near-field optical imagining multiple near-field imaging planes can be observed in which high-order components are stronger than low-order components [15]. Other causes of asymmetry stem out of the ac Stark shifts associated with uneven beam intensities and matter-wave superradiant effects [24]. Both of these effects can make the asymmetry described above more pronounced. For instance, the far-detuned strong beam leads to a matter-wave superradiant effect that favors forward scattering (i.e., þn orders in the direction of the strong beam) [24]. This weak matter-wave amplification feature has never been studied in near-field atom optics. It is important, however, to realize that these effects, which are important for accurate atom number distribution predictions, do not have significant bearing on the matterwave self-imaging point in time. This is because the selfimaging is mainly the work of an instantaneous Dopplershift induced interference. Thus, for the purpose of demonstrating the matter-wave self-imaging effect, the theoretical model given in Eq. (1) is sufficient. It provides a simple but reasonably accurate picture of the underlying physics and explains satisfactorily the observed matterwave self-imaging time. It is, however, incapable of providing an accurate count of numbers of atoms in individual diffraction peaks since we have neglected ac Stark shifts, matter-wave superradiant effects, and the weak beam absorption. A detailed prediction of numbers of atoms in each diffraction peak requires a full numerical simulation beyond the scope of this work that must include these effects. In conclusion, we have studied matter-wave momentum oscillations in a standing light wave field. We have demonstrated a new matter-wave self-imaging effect by experimentally investigating bidirectional, high-order condensate momentum distribution collapse and revival. The observed matter-wave self-imaging results from an atomic c.m. motion induced interference in a finite temporal optical standing wave. It is characterized by the simultaneous suppression of all higher (n 1) diffraction orders and momentum collapse into the zeroth-order condensate component. We thank Dr. Thomas Lucatorto and Dr. Jay Vaishnav for valuable suggestions and comments. We thank Dr. Lü Baolong of WIPM for assistance. K. L and M. S. Z. acknowledge financial support by the National Natural Science Foundation of China. [1] S. Altshuler, L. M. Frantz, and R. Braunstein, Phys. Rev. Lett. 17, 231 (1966). [2] R. J. Cook and A. F. Bernhardt, Phys. Rev. A 18, 2533 (1978). [3] A. F. Bernhardt and B. W. Shore, Phys. Rev. A 23, 1290 (1981). [4] P. E. Moskowitz et al., Phys. Rev. Lett. 51, 370 (1983). [5] P. L. Gould, G. A. Ruff, and D. E. Pritchard, Phys. Rev. Lett. 56, 827 (1986). [6] P. J. Martin et al., Phys. Rev. A 36, 2495 (1987). [7] P. J. Martin et al., Phys. Rev. Lett. 60, 2194 (1988). [8] P. L. Gould et al., Phys. Rev. A 43, 585 (1991). [9] M. Lindberg, Appl. Phys. B 54, 467 (1992). [10] W. Zhang and D. F. Walls, Phys. Rev. A 49, 3799 (1994). [11] Atom Interferometry, edited by P. R. Berman (Academic, San Diego, 1997). [12] M. Kozuma et al., Phys. Rev. Lett. 82, 871 (1999). [13] J. Stenger et al., Phys. Rev. Lett. 82, 4569 (1999). [14] Yu. Ovchinnikov et al., Phys. Rev. Lett. 83, 284 (1999). [15] L. Deng et al., Phys. Rev. Lett. 83, 5407 (1999). [16] E. W. Hagley et al., Phys. Rev. Lett. 83, 3112 (1999). [17] E. W. Hagley et al., Science 283, 1706 (1999). [18] L. Deng et al., Nature (London) 398, 218 (1999). [19] We have a condensate with Thomas-Fermi radii of r 0 ¼ 14 m and L ¼ 75 m in the radial and axial directions. [20] We used the D 2 line for this experiment. The D 1 line scheme corresponds more closely to Eq. (1). [21] Equation (1) neglects any attenuation of the weak counterpropagating beam. It also neglects any presence of matterwave superradiant effects and ac Stark shifts. [22] Studies of ac Stark shifts, matter-wave superradiant effect and weak amplification, and coherently controlled full matter-wave self-imaging will be reported later. [23] We note that due to the conditions of the setup, the experiment reported in Ref. [14] did not exhibit any observable n>1 diffraction orders and therefore is incapable of demonstrating matter-wave self-imaging effect. [24] S. Inouye et al., Science 285, 571 (1999); R. Gonifacio et al., Opt. Commun. 233, 155 (2004); Özgür E. Müstecaplioğlu and L. You, Phys. Rev. A 62, 063615 (2000). 250401-4