Collective Atomic Recoil in Ultracold Atoms: Advances and … · 2007. 3. 17. · 6 Entanglement...

165
Universit ` a degli Studi di Milano Facolt ` a di Scienze Matematiche, Fisiche e Naturali Dottorato di Ricerca in Fisica, Astrofisica e Fisica Applicata Collective Atomic Recoil in Ultracold Atoms: Advances and Applications Coordinatore Prof. Rodolfo Bonifacio Tutore Prof. Rodolfo Bonifacio Tesi di Dottorato di Mary Manuela Cola Ciclo XVI Anno Accademico 2002-2003

Transcript of Collective Atomic Recoil in Ultracold Atoms: Advances and … · 2007. 3. 17. · 6 Entanglement...

  • Università degli Studi di Milano

    Facoltà di Scienze Matematiche, Fisiche e Naturali

    Dottorato di Ricerca in

    Fisica, Astrofisica e Fisica Applicata

    Collective Atomic Recoil

    in Ultracold Atoms:

    Advances and Applications

    Coordinatore Prof. Rodolfo Bonifacio

    Tutore Prof. Rodolfo Bonifacio

    Tesi di Dottorato di

    Mary Manuela Cola

    Ciclo XVI

    Anno Accademico 2002-2003

  • November 14, 2003 c© M M Cola 2003

  • In memory of Prof. Giuliano Preparata

  • If you do boast, consider this: you do not

    support the root, but the root supports you.

    (Rm. 11,18)

    The figure upwards shows evidence for a Bose Einstein condensation of sodium atoms.

    It is taken from PRL 75, 3969 (1995). The figure on the left shows evidence for a

    collective atomic recoil in a BEC. It is a courtesy of M. Inguscio and coworkers.

  • Acknowledgments

    First of all I wish to thank my tutor, Rodolfo Bonifacio, who gave me the possibility

    to approach the interesting physics of collective phenomena.

    Then I should sincerely thank Nicola Piovella, for his support and teachings during

    these years, especially for what concerned the physics of CARL.

    I also learned a lot from Matteo G.A. Paris: a great acknowledgment for his careful

    aid. He introduced me to the physics of quantum optics and quantum information.

    This let me to investigate fruitful topics in atom optics.

    A sincere thank to my Referee Francesco S. Cataliotti for his punctual and concerned

    correction of this thesis, and to Chiara Fort, Leonardo Fallani and Massimo Inguscio

    for all the hours of collaboration we spent together.

    I want to remember also other physicists with whom I often had stimulating discus-

    sions about frontiers of science, Emilio Del Giudice, Enrico Giannetto and Marco

    Giliberti.

    Thanks to Stefano Olivares, Andrea R. Rossi, Alessandro Ferraro and Gabriele

    Marchi. They keep cheerful the atmosphere of our group.

    A special thought to Carlo, for his patience and his encouragement, and to my

    friends Anna, Elisa and Federica for sharing the everyday difficulties.

    Finally I remember Giuliano Preparata: in a difficult moment of my life his passion

    for physics reminded me my passion for physics.

  • Contents

    Introduction iv

    1 Classical CARL 7

    1.1 Equations of motion . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

    1.2 The FEL limit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

    1.2.1 The undamped case . . . . . . . . . . . . . . . . . . . . . . . . 14

    1.2.2 The damping case: adiabatic limit . . . . . . . . . . . . . . . . 15

    1.3 Linear stability analysis . . . . . . . . . . . . . . . . . . . . . . . . . 16

    1.4 Experimental realizations . . . . . . . . . . . . . . . . . . . . . . . . . 19

    1.5 Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

    2 Quantum CARL 23

    2.1 First quantization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

    2.2 Linear regime . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

    2.3 Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

    3 Quantum field theory 33

    3.1 The CARL-BEC model . . . . . . . . . . . . . . . . . . . . . . . . . . 33

    3.2 Coupled-modes equations . . . . . . . . . . . . . . . . . . . . . . . . . 37

    3.3 Linearized three-mode model . . . . . . . . . . . . . . . . . . . . . . . 40

    3.4 Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

    4 Superradiant Rayleigh scattering and matter waves amplification 43

    4.1 Directional matter waves produced by spontaneous scattering . . . . 44

    4.2 Dicke superradiance and emerging coherence . . . . . . . . . . . . . . 47

    4.3 Evidence for decoherence . . . . . . . . . . . . . . . . . . . . . . . . . 49

    i

  • ii Contents

    4.4 “Seeding” the superradiance . . . . . . . . . . . . . . . . . . . . . . . 51

    4.5 Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

    5 Superradiant Rayleigh scattering from a moving BEC 57

    5.1 Theoretical analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

    5.2 Experimental features . . . . . . . . . . . . . . . . . . . . . . . . . . 62

    5.3 Seeding the superradiance . . . . . . . . . . . . . . . . . . . . . . . . 65

    5.4 Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70

    6 Entanglement generation 71

    6.1 The Hamiltonian model . . . . . . . . . . . . . . . . . . . . . . . . . 72

    6.2 Spontaneous emission and small-gain regime . . . . . . . . . . . . . . 74

    6.3 Solution of the linear quantum regime . . . . . . . . . . . . . . . . . . 76

    6.4 Three mode entanglement . . . . . . . . . . . . . . . . . . . . . . . . 80

    6.5 High-gain regime . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83

    6.5.1 The quasi-classical recoil limit ρ À 1 . . . . . . . . . . . . . . 836.5.2 The quantum recoil limit ρ ≤ 1 . . . . . . . . . . . . . . . . . 85

    6.6 Atom-atom and atom-photon entanglement . . . . . . . . . . . . . . . 86

    6.7 Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89

    7 Radiation to atom quantum mapping 91

    7.1 The entangled state . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93

    7.2 The Bell measurement . . . . . . . . . . . . . . . . . . . . . . . . . . 94

    7.3 The displacement operation . . . . . . . . . . . . . . . . . . . . . . . 95

    7.4 The readout system . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96

    7.5 Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97

    8 Effects of decoherence and losses on entanglement generation 99

    8.1 Dissipative Master Equation . . . . . . . . . . . . . . . . . . . . . . . 99

    8.2 Solution of the Fokker-Plank equation . . . . . . . . . . . . . . . . . . 100

    8.3 Evolution from vacuum and expectation values . . . . . . . . . . . . . 101

    8.4 Numerical analysis for the relevant working regimes . . . . . . . . . . 103

    8.5 Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109

    Conclusion and Outlook 111

  • Contents iii

    A General solution of the linear model 115

    B Wigner functions 119

    C Homodyne and multiport homodyne detection 123

    C.1 Matrix notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123

    C.2 Balanced homodyne detection . . . . . . . . . . . . . . . . . . . . . . 124

    C.3 Double homodyne detection . . . . . . . . . . . . . . . . . . . . . . . 126

    D Continuous variable teleportation as conditional measurement 129

    D.1 Conditional quantum state engineering . . . . . . . . . . . . . . . . . 130

    D.2 Joint measurement of two-mode quadratures . . . . . . . . . . . . . . 133

    D.3 Fidelity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135

    E Solution of the Fokker-Planck equation 137

  • Introduction

    At the basis of most phenomena in atomic, molecular, and optical physics is the

    dynamical interaction between optical and atomic fields.

    In many ways, recently developed Bose Einstein Condensates (BECs) of trapped

    alkali atomic vapors [1, 2] are the atomic analog of the optical laser. In fact, with

    the addition of an output coupler, they are frequently referred to as “atom lasers”

    [3]. Despite many interesting and important differences, the main similarity is that

    both optical lasers and atomic BECs involve large numbers of identical bosons oc-

    cupying a single quantum state. As a result, the physics of lasers and BECs involves

    stimulated processes which, due to Bose enhancement, often completely dominate

    the spontaneous processes which play central roles in the non-degenerate regime.

    Just like the discovery of the laser led to the development of nonlinear optics, so

    too the advent of BEC has led to remarkable experimental successes in the field of

    nonlinear atom optics [4, 5, 6, 7, 8, 9].

    The regimes of nonlinear optics and nonlinear atom optics, therefore, represent

    limiting cases, where either the atomic or optical field is not dynamically independent

    because it follows the other field in some adiabatic manner which allows for its

    effective elimination. Outside of these two regimes the atomic and optical fields

    are dynamically independent, and neither field can be eliminated. The dynamics of

    coupled quantum degenerate atomic and optical fields in this intermediate regime is

    the topic of “quantum atom optics”, namely that extension of atom optics where the

    quantum state of a many-particle matter-wave field is being controlled, characterized

    and used in novel applications. Some advances and applications in this field will be

    the object of this thesis.

    One of the more relevant system in quantum atomic optics is composed of a BEC

    driven by a far off-resonant pump laser and coupled to a single mode of an optical

    1

  • 2 Introduction

    ring cavity. The mechanism that lies below this kind of physics is the so-called

    Collective Atomic Recoil Lasing (CARL) in his fully quantized version.

    The CARL mechanism was originally proposed as a new mechanism for the

    generation of coherent light [10, 11, 12]. It consist of three main ingredients: (1) a

    gas of two-level atoms (the active medium) (2) a strong pump laser, which drives the

    two-level atomic transition, and (3) a ring cavity which supports an electromagnetic

    mode (the probe) counterpropagating with respect to the pump. Under suitable

    conditions, the operation of the CARL results in the generation of a coherent light

    field due to the following mechanism. First, a weak probe field is initiated by noise,

    either optical in the form of spontaneously emitted light, or atomic in the form of

    density fluctuations in the atomic gas which backscatters the pump. Once initiated,

    the probe combines with the pump field to form a weak standing wave which acts

    as a periodic optical potential. The center-of-mass motion of the atoms on this

    potential results in a bunching, i.e. a modulation of their density, very similar to

    the combined effects of the wiggler and the light field leads to electron bunching

    in a Free Electron Laser (FEL) [13]. This bunching process is then seen by the

    pump laser as the appearance of a polarization grating in the active medium, which

    results in stimulated backscattering into the probe field. The resulting gain in the

    probe strength further amplifies the magnitude of the standing wave field, generating

    more bunching followed by an increase in stimulated backscattering, and so on. This

    positive feedback mechanism give rise to an exponential growth of both the probe

    intensity and the atomic bunching which leads to the perhaps surprising result that

    the presence of the ring cavity turns the ordinarily stable system of an atomic gas

    driven by a strong pump laser into an unstable system.

    CARL effect was verified experimentally by Bigelow et al. in a hot atomic cell

    [14]. Related experiments by Courtois et al. [15], using cold cesium atoms, and

    by Lippi et al. [16], using hot sodium atoms, measured the recoil induced small-

    signal probe gain, which was interpreted in terms of coherent scattering from an

    induced polarization grating. However, these experiments missed a probe feedback

    mechanism, which is necessary to see the long time scale instability which charac-

    terizes the CARL. The first unambiguous experimental proof of the CARL effect

    has been obtained only very recently [17] in a system of cold atoms in a collision-less

    environment.

  • Introduction 3

    In chapter 1 we review the essential conceptual framework of the original CARL

    showing that self-bunching via an exponential instability can occur under very gen-

    eral conditions. The original CARL theory considers the atoms as classical point

    particles moving in the optical potential generated by the light fields. From an

    atom optics point of view, this correspond to a “ray atom optics” treatments of the

    atomic field, in analogy with the ordinary ray optics treatment of electromagnetic

    fields. This description is valid provided that the characteristic wavelength of the

    matter-wave field remains much smaller than the characteristic length scale of any

    atom-optical element in the system. Such length, for the atomic field, is its De

    Broglie wavelength, determined by the atomic mass and the temperature T of the

    gas. The central atom-optical element of the CARL is the periodic optical potential,

    which acts as a diffraction grating for the atoms, and has the characteristic length

    scale of half the optical wavelength. Hence the classical description is valid provided

    that the temperature is high enough so that the thermal De Broglie wavelength is

    much smaller than the optical wavelength.

    However, the spectacular recent advances in atomic cooling techniques makes

    it likely that CARL experiments using ultracold atomic samples can and will be

    performed in the next future. In particular, subrecoil temperatures can now be

    achieved almost routinely. So CARL theory has been extended to this “wave atom

    optics” regime [18]. In this regime matter-wave diffraction plays a dominant role in

    the CARL dynamics. The main drawback of the semiclassical model is that, as it

    considers the center-of-mass motion of the atoms as classical, it cannot describe the

    discreteness of the recoil velocity, as has been observed in the experiment of Ref.[19]

    for an atomic sample below the recoil temperature. So, to extend the model in the

    region of ultracold atoms, a quantum mechanical description of the center-of-mass

    motion of the atoms should be included. In chapter 2 we present a way to work

    out this program simply performing a first quantization of the external variables of

    the atoms, position and momentum. This simple model gives a description of all

    the features of the considered system and in particular allows to define the main

    different regimes. In the conservative regime (no radiation losses), the quantum

    model depends on a single collective parameter, ρ, that can be interpreted as the

    average number of photons scattered per atom in the classical limit. When ρ À 1,the semiclassical CARL regime is recovered, with many momentum levels populated

  • 4 Introduction

    at saturation. On the contrary, when ρ ≤ 1, the average momentum oscillatesbetween zero and ~~q, where ~q is the difference between the incident and the scatteredwave vectors, and a periodic train of 2π hyperbolic secant pulses is emitted. In the

    dissipative regime (large radiation losses) and in a suitable quantum limit ρ À 1 ,a sequential superradiant scattering occurs, in which after each process atoms emit

    a π hyperbolic secant pulse and populate a lower momentum state. These results

    describe the regular arrangement of the momentum pattern observed in experiments

    of superradiant Rayleigh scattering from a BEC [20].

    In chapter 3 we derive a quantum field theory model of a gas of bosonic two-level

    atoms which interact with a strong, classical, undepleted pump laser and a weak,

    quantized optical ring cavity mode, both of which are as usual assumed to be tuned

    far away from atomic resonances. Starting from the second-quantized hamiltonian

    of the system, we will write an effective model for the time evolution of the ground

    state atomic field operator and for the probe field operator (the CARL-BEC model),

    adiabatically eliminating the excited state atomic field operator and including effects

    of atom-atom collisions.

    In chapter 4 we review the experimental situations, such as superradiant Rayleigh

    scattering and matter waves amplification, that can be interpreted with the full

    quantistic version of CARL model in the dissipative regime, where the radiation

    emission is superradiant.[21]

    In chapter 5 we analyze some experiment performed at European Laboratory for

    Non-linear Spectroscopy (LENS) in Florence about superradiant Rayleigh scattering

    from a moving BEC. This allows to investigate the influence of the initial velocity of

    the condensate on superradiant Rayleigh scattering. The experiment gives evidence

    of a damping of the matter-wave grating which depends on the initial velocity of

    the condensate. We describe this damping in terms of a phase-diffusion decoherence

    process, in good agreement with the experimental results. Moreover we analyze the

    effect of seeding superradiance by a weak signal directed in the opposite direction

    with respect to the pump laser.

    One important consideration is to determine to which extent the quantum state

    of a many-particle atomic field like a BEC can be optically manipulated. In the

    single-particle case, the answer to this problem is known to a large extent. This is

    the domain of atom optics [22], where a number of optical elements for matter waves

  • Introduction 5

    have now been developed, including gratings, mirrors, interferometers, resonators,

    etc. But these optical elements manipulate just the atomic field “density”, or at most

    first-order coherence properties. However, Schrödinger fields possess a wealth of

    further properties past their first-order coherence, including atom statistics, density

    correlation functions. In chapter 6 we investigate the properties, such as quantum

    fluctuations and entanglement, of the quantum system BEC-radiation in the linear

    regime for a good cavity regime. We obtain new analytical results, calculating

    explicitly the statistical properties for atoms and photons and evaluating the state

    of the coupled BEC-light system evolved from vacuum. In the limit of undepleted

    atomic ground state and unsaturated probe field, the quantum CARL Hamiltonian

    reduces to that for three coupled modes. By calculating the exact evolution of the

    state from the vacuum of the three modes we demonstrate that the evolved state

    is a fully inseparable three mode Gaussian one. Moreover we show how this three

    mode Gaussian state can provide a valuable source of atom-atom and atom-photon

    entanglement.

    Entanglement is a crucial resource in the manipulation of quantum information,

    and quantum teleportation [23, 24, 25, 26, 27] is perhaps the most impressive ex-

    ample of quantum protocol based on entanglement. It realizes the transferral of

    (quantum) information between two distant parties that share entanglement. There

    is no physical move of the system from one player to the other, and indeed the two

    parties need not even know each other’s locations. Only classical information is ac-

    tually exchanged between the parties. However, due to entanglement, the quantum

    state of the system at the transmitter location (say Alice) is mapped onto a different

    physical system at the receiver location (say Bob). In chapter 7 we show a scheme

    to realize radiation to atom continuous variable quantum mapping, i.e. to teleport

    the quantum state of a single mode radiation field onto the collective state of atoms

    with a given momentum out of a BEC. The atoms-radiation entanglement needed

    for the teleportation protocol is established through the CARL three linear model

    studied in chapter 6, whereas the coherent atomic displacement is obtained by the

    same interaction with the probe radiation in a classical coherent field.

    In chapter 8 the results obtained in chapter 6 are extended to include the effects

    of losses either due to the optical cavity or to the depletion of atomic modes. The

    calculation are performed by means of the Master equation formalism and a system-

  • 6 Introduction

    atic comparison with respect to the ideal case is given. The results of this chapter

    are promising and give indications for future experiments.

  • Chapter 1

    Classical CARL

    The CARL, a kind of hybrid between the FEL [13] and the ordinary laser, with

    physical features common to both, was thought and presented like a source of tunable

    coherent radiation. Its essential conceptual framework was first outlined in Ref.[10].

    The ordinary laser and the FEL share an important physical trait: they gen-

    erate electromagnetic waves through a noise-initiated process of self-organization.

    In an ordinary laser, for example, the energy is stored initially as excitation of in-

    ternal degrees of freedom of the active medium, while in a FEL it is brought into

    the interaction region as translational kinetic energy of the incident electron beam.

    The spectral character of laser light is constrained mainly by the gain profile of the

    active medium, while in a FEL the frequency of the emitted radiation is assigned

    by the speed of the incident electrons, and can be varied, in principle, over a very

    wide range; hence, the FEL is intrinsically a widely tunable source. Furthermore,

    the laser gain originates from the induced atomic polarization, under the constraint

    that a suitable population inversion exists in the active medium; in a FEL, instead,

    amplification of coherent radiation follows the spontaneous emergence of a suffi-

    ciently large electron bunching, i.e. the appearance of a periodic spatial structure

    in the form of a longitudinal grating on the scale of the electromagnetic wavelength.

    Hence, light amplification in a FEL is the result of a coherent scattering process

    from the grating structure created within the active medium, and it comes at the

    expense of a recoil in the momentum of the individual electrons.

    The physics of the FEL and of the atomic lasers are unified in the CARL. The

    active medium now is a collection of two level atoms initial in their lower state and

    7

  • 8 Chapter 1. Classical CARL

    exposed to a strong pump field. For appropriate values of the parameters the atoms,

    through an exponential instability, can amplify a weak probe field counterpropagat-

    ing with respect to the pump. As in the laser the active medium is characterized by

    bound states which play a key role in the amplification process but do not posses a

    population inversion. Common to the FEL, instead, is the existence of a reservoir of

    momentum that can be transformed partly into radiation through a kind of cooper-

    ative scattering. Furthermore, optical gain is initiated by the growth of a bunching

    parameter. What happens is that the incident optical wave creates an atomic po-

    larization wave in the medium. This polarization couples with the backscattered

    radiation, creating a longitudinal self-consistent pendulum potential which traps

    and than bunches the particles giving rise to a coherent scattering.

    1.1 Equations of motion

    The CARL model is based on the Hamiltonian of a collection of two-level atoms

    interacting with a strong pump field and a weak optical probe counterpropagating

    with respect to the pump. In addition to the internal atomic degrees of freedom,

    which are typical of laser models, the CARL model take explicit account of the

    center-of-mass motion. The explicit form of the Hamiltonian is

    Ĥ = ~ω1â†1â1 + ~ω2â†2â2 + ~ω0

    N∑j=1

    Ŝzj +N∑

    j=1

    p̂2j2m

    +i~

    (g1â

    †1

    N∑j=1

    Ŝ−j e−ik1ẑj + g2â

    †2

    N∑j=1

    Ŝ−j eik2ẑj − h.c.

    )(1.1)

    where N is the number of atoms in the interaction volume V , ω1,2 = ck1,2 are

    the carrier frequency of the probe and pump fields, respectively, k1 and k2 are the

    corresponding wave numbers and ω0 is the atomic transition frequency when the

    atoms are at rest relative to the observer. The couplings constants are defined as

    gi = µ

    [ωi

    2~²0V

    ]1/2i = 1, 2 (1.2)

    where µ is the modulus of the atomic dipole moment. Ŝzj and Ŝ±j are the standard

    effective angular-momentum operators (in units of ~) describing the evolution ofthe internal degrees of freedom of the j-th atom so that Ŝzj measures one-half the

  • 1.1. Equations of motion 9

    difference between the excited and ground state populations of the j-th atom; ẑj

    and p̂j denote, respectively, the position and momentum operators of the center of

    mass of the j-th atom and â†i (i = 1, 2) are the photon creation operators of the

    pump field (index 1) and of the probe field (index 2). The operators obey the usual

    commutation relations:

    [Ŝzi, Ŝ

    ±j

    ]= ±δijŜ±j ,

    [Ŝ+i , Ŝ

    −j

    ]= 2δijŜzi (1.3)

    [ẑi, p̂j] = i~δij (1.4)[âi, â

    †j

    ]= δij (1.5)

    The Hamiltonian (1.1) admits two constants of the motion,

    N∑j=1

    p̂j + ~k1â†1â1 − ~k2â†2â2 = constant (1.6)

    N∑j=1

    Ŝzj + â†1â1 + â

    †2â2 = constant; (1.7)

    the first represents the conservation of the total momentum and the second the

    conservation of the number of excitations. If we combine Eqs.(1.6) and (1.7) and

    eliminate the number operator of the driving field, â†2â2, we can also write

    N∑j=1

    (p̂j + ~k2Ŝzj) + ~(k1 + k2)â†1â1 = constant (1.8)

    whose obvious physical implication is that the expectation value of the number op-

    erator for the probe field, â†1â1, can grow either as the result of a loss of internal

    atomic energy or a decrease of the center-of-mass kinetic energy. This setting rep-

    resents a generalization of the basic mechanism by which energy is produced in the

    laser and in the FEL; in fact, the laser Hamiltonian does not involve the momentum

    and position operators p̂j and ẑj, while the FEL Hamiltonian does not include the

    angular-momentum operators, descriptive of the internal degrees of freedom of the

    active medium.

    In this chapter we analyze the dynamical evolution of the coherent atomic recoil

    lasing within the framework of the standard semiclassical approximation. First we

    construct the Heisenberg equations of motion for the relevant operators and map

  • 10 Chapter 1. Classical CARL

    the operator equation into their c-number counterparts in the usual factorized form

    obtaining

    dzjdt

    =pjm

    (1.9)

    dpjdt

    = −~k1g1a∗1Sje−ik1zj + ~k2g2a∗2Sjeik2zj + c.c. (1.10)

    da1dt

    = −iω1a1 + g1N∑

    j=1

    Sje−ik1zj (1.11)

    dS−jdt

    = −iω0S−j + 2g1a1Szjeik1zj + 2g2a2Szje−ik2zj (1.12)dSzjdt

    = − (g1a∗1S−j e−ik1zj + g2a∗2S−j eik2zj + c.c.). (1.13)

    where we have assumed that a2 is a constant real number. This program is ac-

    complished by introducing the appropriate slowly varying variables ao1, ao2, and Sj

    according to the definitions

    a1(t) = a01(t) exp

    {−i

    [ω2 +

    k1 + k2m

    p̄(0)

    ]t

    }(1.14)

    a2(t) = a02 exp {−iω2t} (1.15)

    S−j (t) = Sj(t) exp(−ik2(zj + ct)) (1.16)

    where p̄(0) ≡ mv̄(0) is the average initial momentum of the atoms. Furthermore,we define the new position and momentum variables θj(t) and δpj(t)

    θj(t) = (k1 + k2)

    [zj − p̄(0)

    mt

    ](1.17)

    δpj(t) = pj(t)− p̄(0), (1.18)

    and the population difference between the ground and excited states of the j-th

    atom,

    Dj(t) = −2Szj(t) (1.19)The required equations of motion take the form

    dθjdt

    =k1 + k2

    mδpj (1.20)

    d

    dtδpj = −~k1g1a0∗1 Sje−iθj + ~k2g2a0∗2 Sj + c.c. (1.21)

    da01dt

    = iδ1,2a01 + g1

    N∑j=1

    Sje−iθj (1.22)

  • 1.1. Equations of motion 11

    dSjdt

    = i

    [ω2c

    δpjm

    + δ2,0

    ]Sj − g1a01Djeiθj − g2a02Dj − γ⊥Sj, (1.23)

    dDjdt

    =(2g1a

    0∗1 Sje

    iθj + 2g2a0∗2 Sj + c.c.

    )− γ‖(Dj −Deqj ) (1.24)

    where we have introduced the detuning parameters

    δ2,1 = (k1 + k2)[v̄(0)− vr,1], (1.25)δ2,0 = k2[v̄(0)− vr,2], (1.26)

    with

    vr,1 =ω1 − ω2ω1 + ω2

    c, vr,2 =ω0 − ω2

    ω2c (1.27)

    and added phenomenological decay terms to the polarization and population equa-

    tions (Deqj = 1 because each atom is assumed to be in the ground state as it enters

    the interaction region). Note that the two resonance conditions δ2,0 = δ2,1 = 0,

    taken together, imply

    ω1 =ω0

    1− β0 , ω2 =ω0

    1 + β0, (1.28)

    with β0 = v̄(0)/c, i.e. they imply a resonance between the atomic transition fre-

    quency ω0 and the Doppler-shifted frequencies of the probe and the pump beams.

    As our final step we introduce the so-called universal scaling and cast the working

    equations in dimensionless form. For simplicity we let k1 ≈ k2 ≡ k = ω/c, g1 ≈g2 ≡ g and, furthermore, define the dimensionless parameter

    ρ′=

    (g√

    N

    ωR

    )2/3∝

    (N

    V

    )1/3(1.29)

    where ωR = 2~k2/m is the single-photon recoil frequency shift, the scaled time τ ,and the new dependent variables Pj and A1,2 according to the definitions

    τ = ωRρ′t Pj =

    δpj~kρ′

    A1,2 =a01,2√Nρ′

    (1.30)

    where A2 is real for definiteness. The parameter ρ′

    is a measure of the collective

    effects; in fact g√

    N is the collective spontaneous Rabi frequency of an ensemble of

    N two-level atoms [28, 29].

  • 12 Chapter 1. Classical CARL

    With these definitions the final form of the CARL equations of motion is

    dθjdτ

    = Pj (1.31)

    dPjdτ

    = −A∗1e−iθjSj − A1eiθjS∗j + 2A2ReSj, (1.32)

    dA1dτ

    = iδA1 +1

    N

    N∑j=1

    Sje−iθj (1.33)

    dSjdτ

    =i

    2(Pj + 2∆)Sj − ρDj(A1eiθj + A2)− Γ⊥Sj, (1.34)

    dDjdτ

    = [2ρ(A′∗1 e

    −iθj + A2)Sj + h.c.]− Γ‖(Dj −Deqj ), (1.35)

    where the remaining parameters are defined as follow:

    δ =δ2,1ωRρ

    ′ ∆ =δ2,0ωRρ

    ′ (1.36)

    Γ⊥ =γ⊥

    ωRρ′ Γ‖ =

    γ‖ωRρ

    ′ . (1.37)

    Eqs. (1.31-1.35) form a closed, self-consistent set of equations for the internal and

    translational atomic degrees of freedom, coupled to the pump field A2 and the probe

    field A1, whose amplification was the main objective of the original works on CARL.

    In arriving at this result, as already mentioned, we have assumed k1 ≈ k2 ≡ k.If k1 6= k2, Eqs. (1.31-1.35) are still valid in the so-called Bambini-Ranieri frame[30, 31] moving with a velocity vr,1 [see Eq.(1.27)], where the transformed frequencies

    coincide. We note that for a nearly resonant interaction, i.e. δ2,1 ≈ 0, it follows thatv̄(0) ≈ vr,1. We can distinguish two cases:

    (a) Non relativistic particles [v̄(0) ¿ c]; in this case we have ω1 ≈ ω2 and ourequations are valid in the laboratory frame.

    (b) Relativistic particles [v̄(0) ≈ c]; in this case it follows that

    ω1 =c + vr,1c− vr,1ω2 =

    1 + β01− β0ω2 ≈ 4γ

    2ω2, (1.38)

    where γ = (1 − β20)−1/2; hence ω1 can be considerably larger than ω2. Thus, thisformulation can also account for the dynamics of relativistic particles; of course, in

    this case one needs an additional Lorentz transformation of Eqs.(1.31-1.35) back to

    the laboratory frame. We will never deal here with this case because our purpose is

    to show how the CARL model can be extended to the ultracold atoms regime.

  • 1.1. Equations of motion 13

    Eqs. (1.34) and (1.35) are the optical Bloch equations, generalized for the in-

    clusion of the atomic translational motion. In addition to the familiar detuning

    term ∆Sj in the polarization equation (1.34), one may note the appearance of the

    time dependent detuning contribution PjSj/2 resulting from the recoil suffered by

    atoms under the action of the pump and probe fields. If we ignore the probe field

    (A1 = 0), Eqs. (1.31-1.35) describe the usual cooling process for time long compared

    to Γ−1⊥ and Γ−1‖ . If, instead, we set A2 = 0 and Sj = 1, for all j, the modified Eqs.

    (1.31-1.33) reduce to the traditional FEL equations.

    The structure of Eq.(1.33) indicates that the probe field A1 can be amplified

    in the presence of an atomic polarization (but without the need of an initial pop-

    ulation inversion) if the phase of the polarization has the appropriate value and

    if the atomic positions are properly bunched. If the scaled position variables are

    uniformly distributed between 0 and 2π, just as one has at the beginning of the evo-

    lution, no macroscopic field source exists even if the atomic polarization variables

    Sj are maximized for all values of j because

    N∑j=1

    e−iθj = 0 (1.39)

    Eqs. (1.31-1.35) for a wide range of the parameters predict the development of an

    exponential instability for the probe field and for the bunching parameter

    B =

    ∣∣∣∣∣1

    N

    N∑j=1

    e−iθj

    ∣∣∣∣∣ . (1.40)

    The result of this instability is the growth of a macroscopic field and the spontaneous

    creation of a longitudinal spatial structure in the initially uniform atomic beam with

    a periodicity that matches the wavelength of the reflected field. This type of behavior

    can be easily demonstrated by numerical integration of Eqs. (1.31-1.35)

    We now want to show that, under certain approximations, the atomic degrees

    of freedom can be eliminated and the CARL equations made formally identical to

    the FEL-model equations with universal scaling [13, 32]. This allows the simple

    description of a Hamiltonian instability leading to exponential growth of the probe

    field and of the bunching parameter. It will be demonstrated both without and in

    presence of atomic damping.

  • 14 Chapter 1. Classical CARL

    1.2 The FEL limit

    1.2.1 The undamped case

    Consider first the case in which Γ = 0 in the Bloch equations (this, in practice,

    means that we take Γ 1. In this case, we can take the time average of the atomic variables

    〈S1 = 0〉 and 〈S2〉 = −S0 where

    S0 =1

    2

    Ω∆

    Ω2 + ∆2. (1.44)

    Note that S0 is maximized for ∆ = Ω, where S0 = 1/4. Upon substituting in Eqs.

    (1.31-1.33) we obtain

    dθjdτ

    = Pj (1.45)

    dPjdτ

    = −S0(A∗e−iθj + Aeiθj) (1.46)

    dA

    dτ= −iδA + S0

    N

    N∑j=1

    e−iθj (1.47)

  • 1.2. The FEL limit 15

    where we have defined A = iA1. Note that the time average has washed out the

    absorptive part S1 of the polarization, leaving only the dispersive part S2, which

    is antisymmetric in the detuning ∆. In particular, in Eq. (1.32), the radiation

    pressure term 2A2> 1/Γ so that we can perform theadiabatic elimination of the polarization and population variables [11]. With simple

    calculations, one can show that we obtain again the FEL equations (1.49-1.51) with

    A substituted by√

    2A and

    S0 =1√2

    Ω∆

    Γ2 + ∆2 + Ω2(1.52)

    This result can be obtained by again assuming |A2|2 >> |A1|2 and Pj

  • 16 Chapter 1. Classical CARL

    1.3 Linear stability analysis

    Equations (1.49-1.51) admit two constants of motion:

    〈p〉+ |A|2 (1.53)

    and〈p2〉2

    + iS0[A∗〈e−iθ〉 − c.c] = H (1.54)

    The first represents momentum conservation, the second defines the Hamiltonian

    from which (1.49-1.51) can be derived. This Hamiltonian system is unstable.

    A linear analysis of (1.49-1.51) leads to the identification of the conditions under

    which the process of self-amplification of the spontaneous emission takes place. Let

    us consider the initial conditions (where the system has an equilibrium solution) with

    zero field, no spatial modulation of the beam of particles in which θj are randomly

    distributed. In this initial situations∑

    j e−imθj = 0 with m = 1, 2.

    Equations (1.49-1.51) are generally valid also if the initial momentum has an

    arbitrary distribution f(p0). In this case, we can label the particles by the initial

    position θ0 and initial momentum p0 so that1N

    ∑Nj=1 e

    −iθj is replaced by

    〈e−iθ(θ0,p0,τ)〉 = 12π

    ∫ 2π0

    f(p0)e−iθ(θ0,p0,τ)dθ0dp0 (1.55)

    where we have assumed a uniform distribution for θ0. One can show that, by lineariz-

    ing (1.49-1.51) around the initial situation above and using a Laplace transformation,

    we find solutions of the form3∑

    j=1

    cjeiλjτ (1.56)

    where cj depends on the initial conditions and λj are the roots of the cubic equation

    λ +

    ∫f(p0)

    (λ + p0)2dp0 = 0 (1.57)

    In particular, for a “cold” case, i.e., if f(p0) is a Dirac delta function centered at a

    value p0 = δ, the cubic equation takes the form

    λ + (λ + δ)2 + 1 = 0. (1.58)

    Note that for v0 = 0, we have δ = (ω2 − ω1)/ωRρ.

  • 1.3. Linear stability analysis 17

    Exponential growth, and thus, unstable behavior, results if the cubic Eq.(1.58)

    has one real and two complex-conjugate roots. In this case, one of the imaginary

    parts of the eigenvalues measures the exponential growth rate of the unstable solu-

    tion.

    For δ > δc = (27/4)1/3, all the roots are real and the system is stable, for

    δ < δc the system is unstable and it shows a collective instability that leads to an

    exponential growth of the probe-field intensity. In particular, for δ = 0 the gain is

    maximum and the intensity grows as

    |A|2 = e√

    3τ = eGt (1.59)

    where

    G =√

    3 ωRρ. (1.60)

    This shows that the growth rate of the collective instability described by this Hamil-

    tonian is governed by ωRρ. In general, the exponential gain depends on the detuning

    parameters δ and ∆.

    The collective recoil process of the system produces an atomic longitudinal grat-

    ing on the scale of the wavelength. This can be measured by the behavior of the

    “bunching” parameter . In fact, the bunching parameter is the source for the field in

    (1.51) and its modulus can range from 0, when the particles are randomly distributed

    in phase, to 1, when the particles are confined periodically to regions smaller than

    the wavelength. When the collective instability develops, the emitted radiation cre-

    ates a correlation between the particles, and the beam becomes strongly modulated

    (self-bunching takes place). The time dependence of the bunching is also exponen-

    tial and produces values that are almost unity. The strong bunching is due to the

    fact that in the high-gain exponential regime, the particles are trapped in the closed

    orbits of a pendulum phase space due the beat of the pump and of the self-consistent

    scatter field which appears explicitly in (1.32). We stress that this behavior takes

    place only for long times τ ≥ 1. In general, one must perform a careful analysis ofthe linear solution which has the form

    A(δ, τ) =3∑

    j=1

    Cjeiλj(δ)τ (1.61)

    where λj(δ) are the three roots of the cubic, and Cj are fixed by the initial conditions.

    Here, we simply describe the results.

  • 18 Chapter 1. Classical CARL

    Figure 1.1: Gain as a function of δ for different interaction times τ . (a) τ = 1:

    Small-gain Madey regime where G is an antisymmetric function of δ. (b) τ =

    10: Characteristic symmetric shape of the gain curve in the high-gain exponential

    regime. Note that the gain curve is orders of magnitude larger than for τ = 1.

    For τ < 1, the time behavior is not exponential and one has the well known

    small-gain Madey regime [33, 34] (see Fig. 1.1a). The gain

    G(δ, τ) =[|A(δ, τ)|2]− |A0|2

    |A0|2 (1.62)

    is an oscillating function of τ , and for fixed τ is an antisymmetric function of δ,

    which is positive (gain) for δ > 0, zero for δ = 0 and negative for δ < 0 showing

    an absorptive behaviour. This gain is associated with the small bunching resulting

    from the fact that the particles are not trapped, i.e. move on open orbits of the

    pendulum phase space. However, when τ >> 1 and δ < δc, the particles becomes

    trapped. As a consequence, the exponential behavior takes over and the gain G as a

    function of δ changes shape as τ increases, and acquires a symmetric dependence on

    δ, as shown in Fig. 1.1b. Its maximum at δ = 0 increases exponentially, as stated

    before. Hence, the exponential regime is observable only if the interaction time is

    sufficiently large so that τ >> 1.

    The behavior of the probe field as a function of the normalized time τ is shown

    in Fig. 1.2a. This is a numerical simulation of the exact set (1.31-1.33) under

  • 1.4. Experimental realizations 19

    Figure 1.2: (a): Probe field as a function of τ for the exact CARL equations with

    Γ = 0, ∆ = 600, δ = 0; (b): Bunching parameter as a function of τ for the exact

    CARL equations, using the parameters of a.

    conditions of zero decay rate (Γ = 0). Fig. 1.2b shows the behavior of the bunching

    parameter as a function of τ . Note the very high value of saturation of the modulus

    of the bunching (≈ 0.8).

    1.4 Experimental realizations

    The signature of CARL is an exponential growth of a seeded probe field oriented

    reversely to a strong pump interacting with an active medium. On the other hand,

    atomic bunching and probe gain can also arise spontaneously from fluctuations with

    no seed field applied. The underlying runaway amplification mechanism is particu-

    larly strong, if the reverse probe field is recycled by a ring cavity.

    After the theoretical proposal of CARL, experiments have been performed in

    order to observe its peculiar features. As we saw the most striking effect due to the

    CARL dynamics is the exponential growth of a seeded probe field oriented reversely

    to the pump field. On the other hand, atomic bunching and probe gain can also

    arise spontaneously from fluctuations with no seed field applied, particularly if the

    runaway amplification mechanism is enforced recycling the reverse probe field by

    a ring cavity. The first attempts to observe CARL action have been undertaken

  • 20 Chapter 1. Classical CARL

    Figure 1.3: Scheme of the experimental setup in Tbingen. A Ti-sapphire laser is

    locked to one of the two counterpropagating modes (α+) of a ring cavity. The beam

    αin− can be switched off by means of a mechanical shutter (S). The atomic cloud is

    located in the free space waist of the cavity mode. The evolution of the interference

    signal between the two light fields leaking through one of the cavity mirrors and the

    spatial evolution of the atoms via absorption imaging are observed. Figure taken

    from Ref. [17].

    only in hot atomic vapors [14, 16]. In particular P.R. Hemmer, N.P. Bigelow and

    coworkers [14] performed the first experiment in a strongly pumped atomic sodium

    vapor without the introduction of a counterpropagating probe. These experiments

    led to the identification of a reverse field with some of the expected characteristics.

    However, the gain observed in the reverse field can have other sources [35], which

    are not necessarily related to atomic recoil.

    The first unambiguous experimental proof of the CARL effect has been obtained

    only very recently [17] in a system of cold atoms in a collision-less environment.

    In this experiment (see Fig.1.3) a high-Q ring cavity is pumped by a Ti-Sapphire

    laser locked to one mode (α+) of the cavity. The85Rb atomic cloud is located in

    the free space waist of the cavity mode with a magneto-optical trap, working at a

    temperature of several 100µK. The reverse field α− has been monitored as the beat

    signal between the field α− itself and the pump α+. In this experiment, in contrast

    with the usual CARL model, the atoms are prepared already in a bunched state.

    In Fig.1.4(a) we can see that oscillations appear on the beat signal, showing the

    arising of the reverse field due to recoil effect even in the absence of a seed field .

    Notice that the amplitude of the oscillations is rapidly dumped, however they are

    still discernible after more than 1ms. Moreover as the interaction time between the

  • 1.4. Experimental realizations 21

    0 50 100 1500

    2

    4

    6

    t (µs)

    Pbe

    at (

    µW) (a)

    (b)

    0 1 20

    200

    400

    600

    800

    t (ms)

    ∆ω/2

    π (k

    Hz)

    (c)1 mm

    (f)

    (e)

    (d)

    Figure 1.4: (a) Recorded time evolution of the observed beat signal between the

    two cavity modes with N = 106 and P cav± = 2W. At time t = 0 the pumping of

    the probe α− has been interrupted. (b) Numerical simulation with the temperature

    adjusted to 200µK. (c) The symbols (X) trace the evolution of the beat frequency

    after switch-off. The dotted line is based on a numerical simulation. The solid line is

    obtained from numerical simulation with the assumption that the fraction of atoms

    participating in the coherent dynamics is 1/10 to account for imperfect bunching.

    (d) Absorption images of a cloud of 6×106 atoms recorded for high cavity finesse at0ms and (e) 6ms after switching off the probe beam pumping. All images are taken

    after a 1ms free expansion time. (f) This image is obtained by subtracting from

    image (e) an absorption image taken with low cavity finesse 6ms after switch-off.

    The intracavity power has been adjusted to the same value as in the high finesse

    case. Figure taken from Ref. [17].

    pump and the atoms increases the detuning between probe and pump increases too

    (see Fig. 1.4(c)). As a consequence, the collective recoil gives rise to a detectable

    displacement of the atoms which has been indeed observed taking time-of-flight

    absorption images of atomic cloud at various times (see Fig. 1.4(d),(e),(f)). Besides

  • 22 Chapter 1. Classical CARL

    these results, in experiment [17] a second set-up that differs from the original CARL

    proposal has been used. The usual CARL dynamics never reaches a steady state and

    the power of the reverse field decreases in time. Hence, in order to reach a stationary

    regime, a friction force has been introduced through an optical molassa so that a

    steady-state velocity of the atoms is reached when the velocity dependent dumping

    force balances the CARL acceleration. As a consequence the reverse field too reaches

    fixed detuning and amplitude, so that the lasing process becomes stationary. The

    work on this subject to explain the experiment is still in progress [37].

    1.5 Concluding remarks

    By taking into account the translational degrees of freedom of the active medium,

    we have described a mechanism that can lead to the exponential amplification of

    a weak probe. Roughly speaking, we can interpret the process of amplification

    as evolving in two steps: first, the external field creates a weak gain profile in

    the frequency response of a collection of independent driven atoms and begins the

    buildup of a spatial structure with the help of the atomic recoil; next the probe,

    whose carrier frequencies lies within a selected gain region of the active medium,

    undergoes exponential amplification. The role of the atomic recoil is essential to

    this process: not only it is the cause of the emergence of the spatial grating pattern,

    but it also reinforces the coherent growth of the signal to be amplified as energy is

    transferred from the atoms to the probe field.

    An alternative way of interpreting the probe amplification is to view it as the

    reflection of the pump field from the moving grating pattern or as a kind of coherent

    scattering from the bound states of the atoms.

    We stress that even though we have demonstrated the amplification of a probe

    signal, since the saturation value of the intensity and of the bunching is independent

    of the initial value of the probe, the process can be initiated from spontaneous

    emission noise.

  • Chapter 2

    Quantum CARL

    The realization of Bose Einstein condensation in dilute alkali gases [1, 2] opened the

    possibility to study the coherent interaction between light and an ensemble of atoms

    prepared in a single quantum state. For example, Bragg diffraction [38] of a BEC by

    a moving optical standing wave can be used to diffract any fraction of the condensate

    into a selectable momentum state, realizing an atomic beam splitter. In particular,

    collective light scattering and matter-wave amplification caused by coherent center-

    of mass motion of atoms in a condensate illuminated by a far off-resonant laser

    [19, 39, 40] have been interpreted as superradiant Rayleigh scattering and can be

    investigated using a quantum theory based on a quantum multi-mode extention of

    the CARL model [21, 41, 42, 43]. The main drawback of the semiclassical model is

    that, as it considers the center-of-mass motion of the atoms as classical, it cannot

    describe the discreteness of the recoil velocity, as has been observed in the experiment

    of Ref.[19]. The original CARL theory, which treats the atomic center-of-mass

    motion classically, fails when the temperature of the atomic sample is below the

    recoil temperature TR = ~ωR/kB, M is the atomic mass and kB is the Boltzmannconstant. So, to extend the model in the region of ultracold atoms, a quantum

    mechanical description of the center-of-mass motion of the atoms must be included.

    In this chapter we present a way to work out this program simply performing

    a first quantization of the external variables θ and P of atoms [20]. Even if not

    complete this model gives a simple description of all the features of the considered

    system and in particular allows to define the main different regimes.

    In the conservative regime (no radiation losses), the quantum model depends on

    23

  • 24 Chapter 2. Quantum CARL

    a single collective parameter, ρ, that can be interpreted as the average number of

    photons scattered per atom in the classical limit. When ρ À 1, the semiclassicalCARL regime is recovered, with many momentum levels populated at saturation.

    On the contrary, when ρ ≤ 1, the average momentum oscillates between zero and~~q, and a periodic train of 2π hyperbolic secant pulses is emitted.

    In the dissipative regime (large radiation losses) and in a suitable quantum limit

    (ρ <√

    2κ), a sequential superfluorescence scattering occurs, in which after each pro-

    cess atoms emit a π hyperbolic secant pulse and populate a lower momentum state.

    These results describe the regular arrangement of the momentum pattern observed

    in the aforementioned experiments of superradiant Rayleigh scattering from a BEC.

    2.1 First quantization

    The atomic motion is quantized when the average recoil momentum is comparable

    to ~~q where ~q = ~k2 − ~k1 is the difference between the incident and the scatteredwave vectors, i.e. the recoil momentum gained by the atom trading a photon via

    absorbtion and stimulated emission between the incident and scattered waves. The

    starting point of the following model is the classical model of equations (1.49-1.51)

    derived in chapter 1

    dθjdτ

    = Pj (2.1)

    dPjdτ

    = − [Aeiθj + A∗e−iθj] (2.2)

    dA

    dτ= iδA +

    1

    N

    N∑j=1

    e−iθj (2.3)

    for N two-level atoms exposed to an off-resonant pump laser, whose electric field has

    a frequency ω2 = ck2 with a detuning from the atomic resonance, ∆20 = ω2 − ω0,much larger than the natural linewidth of the atomic transition, γ. The ‘probe

    field’ has frequency ω1 = ω2 − ∆21 and electric field with the same polarization ofthe pump field. In the absence of an injected probe field, the emission starts from

    fluctuations and the propagation direction of the scattered field is determined either

    by the geometry of the condensate (as in the case of the MIT experiment [19], where

    the condensate has a cigar shape) or by the presence of an optical resonator tuned

    on a selected longitudinal mode.

  • 2.1. First quantization 25

    In order to quantize both the radiation field and the center-of-mass motion of the

    atoms, we consider θj, pj = (ρ/2)Pj = Mvzj/~q and a = (Nρ/2)1/2A as quantumoperators satisfying the canonical commutation relations

    [θ̂j, p̂j′

    ]= iδjj′

    [â, â†

    ]= 1. (2.4)

    With these definitions, Eqs.(2.1)-(2.3) are transformed into the Heisenberg equations

    of motion

    dθ̂jdτ

    =2

    ρp̂j (2.5)

    dp̂jdτ

    = −√

    ρ

    2N

    [âeiθ̂j + â∗e−iθ̂j

    ](2.6)

    dâ

    dτ= iδâ +

    √ρ

    2N

    N∑j=1

    e−iθ̂j (2.7)

    associated with the Hamiltonian:

    Ĥ =1

    ρ

    N∑j=1

    p̂2j + i

    √ρ

    2N

    (N∑

    j=1

    â†e−iθ̂j − h.c.)− δâ†â =

    N∑j=1

    Hj(θ̂j, p̂j), (2.8)

    where

    Ĥj(θ̂j, p̂j) =1

    ρp̂2j + i

    √ρ

    2N(â†e−iθ̂j − aeiθ̂j)− δ

    Nâ†â (2.9)

    We note that [Ĥ, Q̂] = 0, where Q̂ = â†â +∑

    j p̂j is the total momentum in units

    of ~q. In order to obtain a simplified description of a BEC as a system of Nnoninteracting atoms in the ground state, we use the Schrödinger picture for the

    atoms (instead of the usual Heisenberg picture [44]), i.e.

    |ψ(θ1, . . . , θN)〉 = |ψ(θ1)〉 . . . |ψ(θN)〉, (2.10)

    where |ψ(θj)〉 obeys the single-particle Schrödinger equation,

    i∂

    ∂τ|ψ(θj)〉 = Hj(θj, pj)|ψ(θj)〉. (2.11)

    In this model we describe the scattered radiation field classically. Hence, considering

    the corresponding c-number a of the field operator â (i.e. its expectation value),

    eq.(2.7) yields:

    da

    dτ= iδa + g

    N∑j=1

    〈ψ(θj)|e−iθj |ψ(θj)〉. (2.12)

  • 26 Chapter 2. Quantum CARL

    Let now expand the single-atom wavefunction on the momentum basis, |ψ(θj)〉 =∑n cj(n)|n〉j, where p̂j|n〉j = n|n〉j, n = −∞, . . .∞ and cj(n) is the probability

    amplitude of the j-th atom having momentum −n~~q. Remembering that

    [e±iθ̂i , p̂j ] = −δi,j e±iθ̂i and e±iθ̂j |n〉j = |n± 1〉j (2.13)

    we obtainda

    dτ= iδa + g

    N∑j=1

    c∗j(n + 1)cj(n). (2.14)

    Introducing the collective density %̂ with matrix elements on the base {|n〉}

    %m,n =1

    N

    N∑j=1

    cj(m)∗cj(n)ei(m−n)δτ , (2.15)

    a straightforward calculation yields, from Eqs.(2.12) and (2.15) , the following closed

    set of equations:

    d%m,ndτ

    = i(m− n)δm,n%m,n+

    ρ

    2[A (%m+1,n − %m,n−1) + A∗ (%m,n+1 − %m−1,n)] (2.16)

    dA

    dτ=

    ∞∑n=−∞

    %n,n+1 − κA, (2.17)

    where δm,n = δ + (m + n)/ρ and we have redefined the field as A =√

    2/ρNae−iδτ .

    We have also introduced a damping term −κA in the field equation, whereκ = κc/ωRρ, κc = c/2L and L is the sample length along the probe propagation,

    which provides an approximated model describing the escape of photons from the

    atomic medium. In the presence of a ring cavity of length Lcav and reflectivity R,

    κc = −(c/Lcav)lnR, as shown in the usual “mean-field” approximation [44].Eqs.(2.16) and (2.17) determine the temporal evolution of the density matrix

    elements for the momentum levels. In particular, Pn = %n,n is the probability of

    finding the atom in momentum level |n〉, 〈p̂〉 = ∑n n%n,n is the average momentumand

    B =∑

    n

    %n,n+1 (2.18)

    is the bunching parameter. Eqs.(2.16) and (2.17), as we will show in the next

    chapter, are the equations for expectation values correspondent to those derived

  • 2.1. First quantization 27

    Figure 2.1: Classical limit of CARL for ρ À 1 in the case κ = 0. (a): |A|2 vs. τ asobtained from the classical eqs.(1)-(3) (dashed line) and from the quantum eqs.(7)

    and (8) for ρ = 10 (solid line); (b): population level pn vs. n at the occurring

    of the first maximum of |A|2, at τ = 12.4. The other parameters are δ = 0 andA(0) = 10−4. Figure taken from Ref. [20].

    in a complete second quantized treatment which introduces bosonic creation and

    annihilation operators of a given center-of-mass momentum [18].

    For a constant field A, Eq.(2.16) describes a Bragg scattering process, in which

    m− n photons are absorbed from the pump and scattered into the probe, changingthe initial and final momentum states of the atom from m to n. Conservation of

    energy and momentum require that during this process ω1 − ω2 = (m + n) ωR, i.e.δm,n = 0. Eqs.(2.16) and (2.17) conserve the norm, i.e.

    ∑m %m,m = 1, and, when

    κ = 0, also the total momentum 〈Q̂〉 = (ρ/2)|A|2 + 〈p̂〉.Fig. 2.1a shows |A|2 vs. τ , for κ = 0, δ = 0 and A(0) = 10−4, comparing the

    semiclassical solution with the quantum solution in the semiclassical limit that

    corresponds to ρ À 1: the dashed line is the numerical solution of Eqs.(2.1)-(2.3), fora classical system of N = 200 cold atoms, with initial momentum pj(0) = 0 (where

    j = 1, . . . , N) and phase θj(0) uniformly distributed over 2π, i.e. unbunched; the

    continuous line is the numerical solution of Eqs.(2.16) and (2.17) for ρ = 10 and

    a quantum system of atoms initially in the ground state n = 0, i.e. with %n,m =

    δn0δm0. We can notice that the quantum system behaves, with good approximation,

    classically. Because the maximum dimensionless intensity is |A|2 ≈ 1.4, the constantof motion 〈Q̂〉 gives 〈p̂〉 ≈ −0.7ρ and the maximum average number of emitted

  • 28 Chapter 2. Quantum CARL

    photons is about 〈â†â〉 ∼ Nρ. Hence in this limit the CARL parameter ρ canbe interpreted as the maximum average number of photons emitted per atom (or

    equivalently, as the maximum average momentum recoil, in units of ~q, acquired bythe atom) in the classical limit. Fig.2.1b shows the distribution of the population

    level Pn at the first peak of the intensity of Fig. 2.1a, for τ = 12.4. We observe

    that, at saturation, twenty-five momentum levels are occupied, with an induced

    momentum spread comparable to the average momentum.

    2.2 Linear regime

    Let us now consider the equilibrium state with no probe field, A = 0, and all the

    atoms in the same momentum state n, i.e. with %n,n = 1 and the other matrix

    elements zero. This is equivalent to assume the temperature of the system equal to

    zero and all the atoms moving with the same velocity −n~~q, without spread. Thisequilibrium state is unstable for certain values of the detuning. In fact, by linearizing

    Eqs.(2.16) and (2.17) around the equilibrium state, the only matrix elements giving

    linear contributions are %n−1,n and %n,n+1, showing that in the linear regime the

    only transitions allowed from the state n are those towards the levels n − 1 andn+1. Introducing the new variables Bn = %n,n+1 + %n−1,n and Dn = %n,n+1− %n−1,n,Eqs.(2.16) and (2.17) reduce to the linearized equations:

    dBndτ

    = −iδnBn − iρDn (2.19)

    dDndτ

    = −iδnDn − iρBn − ρA (2.20)

    dA

    dτ= Dn − κA, (2.21)

    where δn = δ + 2n/ρ. Seeking solutions proportional to ei(λ−δn)τ , we obtain the

    following cubic dispersion relation:

    (λ− δn − iκ)(λ2 − 1/ρ2) + 1 = 0. (2.22)

    In the exponential regime, when the unstable (complex) root λ dominates,

    B(τ) ∼ ei(λ−δn)τ and, from Eq.(2.19), Dn = −ρλBn. The classical limit is recoveredfor ρ À 1 when κ = 0 or ρ À √κ when κ > 1 and δn ≈ δ, i.e. neglecting theshift due to the recoil frequency ωR. In this limit, maximum gain occurs for δ = 0,

  • 2.2. Linear regime 29

    with λ = (1− i√3)/2 when κ = 0 or λ = −(1 + i)/√2κ when κ > 1. Furthermore,|%n,n+1| ∼ |%n−1,n|, so that the atoms may experience both emission and absorbtion.This result can be interpreted in terms of single-photon emission and absorption by

    an atom with initial momentum −n~~q. In fact, energy and momentum conservationimpose ω1 − ω2 = (2n ∓ 1)ωR (i.e. δn = ±1/ρ) when a probe photon is emittedor absorbed, respectively. Because in the semiclassical limit the gain bandwidth is

    ∆ω ∼ ωRρ À ωR when κ = 0 (or ∆ω ∼ κc À ωR when κ > 1) the atom can bothemit or absorbe a probe photon.

    On the contrary, in the quantum limit the recoil energy ~ωR can not be ne-glected, and there is emission without absorbtion if |%n,n+1| ¿ |%n−1,n|, i.e.

    Bn ≈ −Dn , λ ≈ 1ρ. (2.23)

    This is true for ρ < 1 when κ = 0 with the unstable root

    λ ≈ 1ρ

    +δ′n

    2− 1

    2

    √(δ′n)

    2 − 2ρ (2.24)

    (where δ′n = δn − 1/ρ), and for ρ <

    √2κ when κ > 1 with

    1),which are both less than the frequency difference 2ωR between the emission and

    absorbtion lines. Hence, in the quantum limit the optical gain is due exclusively to

    emission of photons, whereas in the semiclassical limit gain results from a positive

    difference between the average emission and absorbtion rates. When κ = 0, the

    resonant gain in the limit ρ < 1 is

    GS = ωRρ

    √ρ

    2=

    √3

    Ω02∆20

    γ√

    Neff , (2.26)

    where γ = µ2k3/3π~²0 is the natural decay rate of the atomic transition, Ω0 isthe Rabi frequency of the pump and Neff = (λ

    2/A)(c/γL)N is the effective atomic

    number in the volume V = ΣL, where Σ and L are the cross section and the length

    of the sample. When κ > 1, the resonant superfluorence gain in the limit ρ <√

    is

    GSF =ωRρ

    2

    2κ=

    3

    4πγ

    (Ω0

    2∆20

    )2λ2

    AN. (2.27)

  • 30 Chapter 2. Quantum CARL

    Figure 2.2: Quantum limit of CARL for ρ < 1 in the case κ = 0. (a) |A|2 and (b)〈p〉 vs. τ , for ρ = 0.2, δ = 5, A(0) = 10−5 and the atoms initially in the state n = 0.We note that 〈p〉 = −(ρ/2)(|A|2 − |A(0)|2). Figure taken from Ref. [20].

    The above results show that the combined effect of the probe and pump fields on a

    collection of cold atoms in a pure momentum state n is responsible of a collective

    instability that leads the atoms to populate the adjacent momentum levels n − 1and n + 1. However, in the quantum limit ρ < 1 when κ = 0 (or ρ <

    √2κ when

    κ > 1) conservation of energy and momentum of the photon constrains the atoms

    to populate only the lower momentum level n− 1. This holds also in the nonlinearregime, as we have verified solving numerically Eqs.(2.16) and (2.17).

    In the quantum limit above, the exact equations reduce to those for only three

    matrix elements, %n,n, %n−1,n−1 and %n−1,n, with %n−1,n−1 +%n,n = 1. Introducing the

    new variables Sn = Sn−1,n and Wn = %n,n − %n−1,n−1, Eqs.(2.16) and (2.17) reduceto the well-known Maxwell-Bloch equations [45]:

    dSndτ

    = −iδ′nSn +ρ

    2AWn (2.28)

    dWndτ

    = −ρ(A∗Sn + h.c.) (2.29)dA

    dτ= Sn − κA. (2.30)

    When κ = 0 and δ′n = 0, if the system starts radiating incoherently by pure quantum-

    mechanical spontaneous emission, the solution of Eqs.(2.28)-(2.30) is a periodic train

    of 2π hyperbolic secant pulses [46] with

    |A|2 = (2/ρ) Sech2[√

    ρ

    2(τ − τn)

    ], (2.31)

  • 2.2. Linear regime 31

    Figure 2.3: Sequential superfluorescent (SF) regime of CARL. (a) |A|2 and (b) 〈p〉vs. τ , for ρ = 2, δ = 0.5, κ = 10, and the same initial conditions of fig.2.2. Figure

    taken from Ref. [20].

    where τn = (2n + 1)ln(ρ/2)/√

    ρ/2. Furthermore, the average momentum

    〈p̂〉 = n + Th2[√

    ρ

    2(τ − τn)

    ]− 1 (2.32)

    oscillates between n and n−1 with period τn. We observe that the maximum numberof photons emitted is 〈â†â〉peak = (ρN/2)|A|2peak = N , as expected. Fig. 2.2 showsthe results of a numerical integration of Eqs.(2.16) and (2.17), for κ = 0, ρ = 0.2 and

    δ = 5, with the atoms initially in the momentum level n = 0 and the field starting

    from the seed value A0 = 10−5. The intensity |A|2 and the average momentum 〈p̂〉

    vs. τ are in agreement with the predictions of the reduced Eqs.(2.28)-(2.30).

    In the superradiant regime, κ > 1, Eqs.(2.28)-(2.30) describe a single SF

    scattering process in which the atoms, initially in the momentum state n, ‘decay’

    to the lower level n − 1 emitting a π hyperbolic secant pulse, with intensity andaverage momentum

    |A|2 = 14[κ2 + (δ′n)2]

    Sech2[(τ − τD)

    τSF

    ], 〈p̂〉 = n− 1

    2

    {1 + Th

    [(τ − τD)

    τSF

    ]}(2.33)

    where τSF = 2(κ2 + δ′2n )/ρκ is the ‘superfluorescence time’ [28], the delay time

    is τD = τSF Arcsech(2|Sn(0)|) ≈ −τSF ln√

    2|Sn(0)| and |Sn(0)| ¿ 1 is the initialpolarization.

    Figures 2.3a and b shows |A|2 and 〈p̂〉 vs. τ calculated solving Eqs.(2.16) and(2.17) numerically with κ = 10, ρ = 2, δ = 0.5 and the same initial conditions of Fig.

  • 32 Chapter 2. Quantum CARL

    2.2. We observe a sequential SF scattering, in which the atoms, initially in the level

    n = 0, change their momentum by discrete steps of ~~q and emit a SF pulse duringeach scattering process. We observe that for δ = 1/ρ the field is resonant only with

    the first transition, from n = 0 to n = −1; for a generic initial state n, resonanceoccurs when δ = (1 − 2n)/ρ, so that in the case of Fig. 2.3a the peak intensityof the successive SF pulses is reduced (by the factor 1/[κ2 + (2n/ρ)2]) whereas the

    duration and the delay of the pulse are increased. However, the pulse retains the

    characteristic Sech2 shape and the area remains equal to π, inducing the atoms to

    decrease their momentum by a finite value ~~q. We note that, although the SF timein the quantum limit (τSF = 2κ/ρ at resonance) can be considerable longer than

    the characteristic superradiant time obtained in the classical limit, τSR =√

    2κ, the

    peak intensity of the pulse in the quantum limit is always approximately half of the

    value obtained in the semiclassical limit (see Ref.[47] for details).

    2.3 Concluding remarks

    We have shown that the CARL model describing a system of atoms in their momen-

    tum ground state (as those obtained in a BEC) and properly extended to include a

    quantum-mechanical description of the center-of-mass motion, allows for a quantum

    limit in which the average atomic momentum changes in discrete units of the photon

    recoil momentum ~~q and reduce to the Maxwell-Bloch equations for two momen-tum levels. The behavior of the system is different for conservative and dissipative

    regimes. The regular arrangement of momentum pattern observed in the superradi-

    ant Rayleigh scattering experiments with BECs (see also chapter 4 for details) can

    be interpreted as being due to the sequential superfluorescence scattering.

  • Chapter 3

    Quantum field theory

    In this chapter we derive a fully quantized model of a gas of bosonic two-level atoms

    which interact with a strong, classical, undepleted pump laser and a weak, quantized

    optical ring cavity mode, both of which are as usual assumed to be tuned far away

    from atomic resonances. Starting from the second-quantized hamiltonian of the

    system, we will write an effective model for the time evolution of the ground state

    atomic field operator and of the probe field operator, adiabatically eliminating the

    excited state atomic field operator and including effects of atom-atom collisions [48].

    3.1 The CARL-BEC model

    The second-quantized Hamiltonian of the system is

    Ĥ = Ĥatom + Ĥprobe + Ĥatom−probe + Ĥatom−pump + Ĥatom−atom, (3.1)

    where Ĥatom and Ĥprobe give the free evolution of the atomic field and the probemode respectively, Ĥatom−probe and Ĥatom−pump describe the dipole coupling betweenthe atomic field and the probe mode and pump laser, respectively, and Ĥatom−atomcontains the two-body s-wave scattering collisions between ground state atoms.

    The free atomic Hamiltonian is given by

    Ĥatom =∫

    d3z

    [Ψ̂ †g (z)

    (− ~

    2

    2m∇2 + Vg(z)

    )Ψ̂g(z)

    + Ψ̂ †e (z)(− ~

    2

    2m∇2 + ~ω0 + Ve(z)

    )Ψ̂e(z)

    ], (3.2)

    33

  • 34 Chapter 3. Quantum field theory

    where m is the atomic mass, ωa is the atomic resonance frequency, Ψ̂e(z) and Ψ̂g(z)

    are the atomic field operators for excited and ground state atoms respectively, and

    Vg(z) and Ve(z) are their respective trap potentials. The atomic field operators obey

    the usual bosonic equal time commutation relations[Ψ̂j(z), Ψ̂

    †j′(z

    ′)]

    = δj,j′δ3(z− z′) (3.3)

    [Ψ̂j(z), Ψ̂j′(z

    ′)]

    = [Ψ̂ †j (z), Ψ̂†j′(z

    ′)] = 0, (3.4)

    where j, j′ = {e, g}. The free evolution of the probe mode is governed by theHamiltonian

    Ĥprobe = ~ck1†Â, (3.5)where c is the speed of light, k1 is the magnitude of the probe wave number k1, and

     and † are the probe photon annihilation and creation operators, satisfying the

    boson commutation relation [Â, †] = 1. The probe wavenumber k1 must satisfy

    the periodic boundary condition of the ring cavity, k1 = 2π`/L, where the integer `

    is the longitudinal mode index, and L is the length of the cavity.

    The atomic and probe fields interact in the dipole approximation via the Hamil-

    tonian

    Ĥatom−probe = −i~g1Â∫

    d3zΨ̂ †e (z)eiks·zΨ̂g(z) + H.c., (3.6)

    where g1 = µ[ck1/(2~²0LS)]1/2 is the atom-probe coupling constant. Here µ isthe magnitude of the atomic dipole moment, and S is the cross-sectional area of

    the probe mode in the vicinity of the atomic sample (where it is assumed to be

    approximately constant across the length of the atomic sample).

    In addition, the atoms are driven by a strong pump laser, which is treated classi-

    cally and assumed to remain undepleted. The atom-pump interaction Hamiltonian

    is given in the dipole approximation by

    Ĥatom−pump = ~Ω2

    e−iω2t∫

    d3zΨ̂ †e (z)eik2·zΨ̂g(z) + H.c., (3.7)

    where Ω is the Rabi frequency of the pump laser, related to the pump intensity I by

    Ω2 = 2µ2I/~2²0c, ω2 is the pump frequency, and k2 ≈ ω2/c is the pump wavenumber.The approximation indicates that we are neglecting the index of refraction inside

    the atomic gas, as we assume a very large detuning ∆20 = ω2 − ω0 between thepump frequency and the atomic resonance frequency.

  • 3.1. The CARL-BEC model 35

    Finally, the collision Hamiltonian is taken to be

    Ĥatom−atom = 2π~2σ

    m

    ∫d3zΨ̂ †g (z)Ψ̂

    †g (z)Ψ̂g(z)Ψ̂g(z), (3.8)

    where σ is the atomic s-wave scattering length. This corresponds to the usual s-wave

    scattering approximation, and leads in the Hartree approximation to the standard

    Gross-Pitaevskii equation for the ground state wavefunction (in the absence of the

    driving optical fields).

    We limit ourselves to the case where the pump laser is detuned far enough away

    from the atomic resonance that the excited state population remains negligible,

    a condition which requires that ∆ À γa. In this regime the atomic polarizationadiabatically follows the ground state population, allowing the formal elimination

    of the excited state atomic field operator.

    First we write the Heisenberg equation of motion for the field operators. The

    commutation relation with the Hamiltonian are[Ψ̂g(z), Ĥprobe

    ]=

    [Ψ̂e(z), Ĥprobe

    ]=

    [Ψ̂e(z), Ĥatom−atom

    ]= 0 (3.9)

    [Â, Ĥatom

    ]=

    [Â, Ĥatom−pump

    ]=

    [Â, Ĥatom−atom

    ]= 0 (3.10)

    [Ψ̂g(z), Ĥatom

    ]=

    (− ~

    2

    2m∇2 + Vg(z)

    )Ψ̂g(z) (3.11)

    [Ψ̂g(z), Ĥatom−probe

    ]= i~g1†Ψ̂e(z)e−ik1·z (3.12)

    [Ψ̂g(z), Ĥatom−pump

    ]=

    ~Ω2

    eiω2tΨ̂e(z)e−ik2·z (3.13)

    [Ψ̂g(z), Ĥatom−atom

    ]=

    4π~2σm

    Ψ̂ †g (z)Ψ̂g(z)Ψ̂g(z) (3.14)[Ψ̂e(z), Ĥatom

    ]=

    (− ~

    2

    2m∇2 + ~ω0 + Ve(z)

    )Ψ̂e(z) (3.15)

    [Ψ̂e(z), Ĥatom−probe

    ]= −i~g1Âeik1·zΨ̂g(z) (3.16)

    [Ψ̂e(z), Ĥatom−pump

    ]=

    ~Ω2

    e−iω2teik2·zΨ̂g(z) (3.17)[Â, Ĥprobe

    ]= ~ck1Â (3.18)

    [Â, Ĥatom−probe

    ]= i~g1

    ∫d3zΨ̂ †g (z)e

    −ik1·zΨ̂e(z) (3.19)

    so the equation of motions read

    i~dΨ̂g(z)

    dt=

    (− ~

    2

    2m∇2 + Vg(z) + 4π~

    mΨ̂ †g (z)Ψ̂g(z)

    )Ψ̂g(z)

  • 36 Chapter 3. Quantum field theory

    +

    (i~gs†e−iks·z +

    ~Ω2

    eiωte−ik·z)

    Ψ̂e(z)

    i~dΨ̂e(z)

    dt= ~ω0Ψ̂e(z) +

    (−i~g1Âeik1·z + ~Ω

    2e−iω2teik2·z

    )Ψ̂g(z) (3.20)

    i~dÂ

    dt= ~ck1Â + i~g1

    ∫d3rΨ̂e(z)e

    −ik1·zΨ̂ †g (z) (3.21)

    where we have dropped the kinetic energy and trap potential terms under the as-

    sumption that the lifetime of the excited atom, which is of the order 1/∆, is so small

    that the atomic center-of-mass motion may be safely neglected during this period.

    For the same reason, we are justified in neglecting collisions between excited atoms,

    or between excited and ground state atoms in the collision Hamiltonian (3.8).

    We proceed by introducing the operators Ψ̂ ′e(z) = Ψ̂e(z)eiω2t and â = Âeiω2t,

    which are slowly varying relative to the optical driving frequency. The new excited

    state atomic field operator obeys then the Heisenberg equation of motion

    i~d

    dtΨ̂ ′e(z) = −~∆20Ψ̂ ′e(z) +

    [~Ω2

    eik2·z − i~g1âeik1·z]

    Ψ̂g(z), (3.22)

    We now adiabatically solve for Ψ̂ ′e(z) by formally integrating Eq. (3.22) under the

    assumption that Ψ̂g(z) varies on a time scale which is much longer than 1/∆20. This

    yields

    Ψ̂ ′e(z, t) ≈1

    ∆20

    [Ω

    2eik2·z − ig1â(t)eik1·z

    ]Ψ̂g(z, t)

    − 1∆20

    [Ω

    2eik2·z − ig1â(0)eik1·z

    ]Ψ̂g(z, 0)e

    i∆t

    + Ψ̂ ′e(z, 0)ei∆20t. (3.23)

    The third term on the r.h.s. of Eq. (3.23) can be neglected for most considerations

    if we assume that there are no excited atoms at t = 0, so that this term acting on

    the initial state gives zero. The second term may also be neglected, as it is rapidly

    oscillating at frequency ∆20, and thus its effect on the ground state field operator

    is negligible when compared to that of the first term, which is non-rotating. It is

    useful to keep them temporarily to demonstrate that the commutation relation for

    Ψ̂e(z) is preserved (to order 1/∆20) by the procedure of adiabatic elimination.

    Dropping the unimportant terms, and then substituting Eq. (3.23) into the

    equations of motion for Ψ̂g(z) and for Â, we arrive at the effective Heisenberg equa-

    tions of motion for the ground state field operator and for the probe field operator

  • 3.2. Coupled-modes equations 37

    (CARL-BEC model)

    i~d

    dtΨ̂g(z) =

    [−~22m

    ∇2 + Vg(z) + 4π~2σ

    mΨ̂ †g (z)Ψ̂g(z)

    +i~g(â†e−iq·z − âeiq·z)

    +~(

    Ω2

    4∆20+

    4∆20g2

    Ω2â†â

    )]Ψ̂g(z), (3.24)

    i~d

    dtâ = −~δ̃â + i~g

    ∫d3zΨ̂ †g (z)e

    −iq·zΨ̂g(z), (3.25)

    where we have introduced the new coupling constant g = g1Ω/∆20 that contains the

    parameters of the pump field. The recoil momentum kick the atom acquires from

    the two-photon transition is q = k1−k2, the detuning between the pump and probefields is δ21 = ω2−ω1 and the probe frequency is given by ω1 ≈ ck1, again assumingthat the index of refraction inside the condensate is negligible. The second term in

    Eq. (3.24) is simply the optical potential formed from the counterpropagating pump

    and probe light fields, and the last term gives the spatially independent light shift

    potential, which can be thought of as cross-phase modulation between the atomic

    and optical fields.

    3.2 Coupled-modes equations

    We assume that the atomic field is initially in a BEC with mean number of condensed

    atoms N . Furthermore, we assume that N is very large and that the condensate

    temperature is small compared to the critical temperature. These assumptions allow

    us to neglect the non-condensed fraction of the atomic field. Thus this model does

    not include any effect of condensate number fluctuations. We introduce creation

    and annihilation operators for the atoms of a definite momentum p = n~q. So wesuppose we can write

    Ψ̂(z) =+∞∑n=0

    ĉnΦn(z) (3.26)

    where Φ0(z) is the condensate ground state that satisfies the time independent

    Gross-Pitaevskii equation

    (~2

    2m∇2 − Vg(z)− 4π~

    mN |Φ0(z)|2

    )Φ0(z) = 0. (3.27)

  • 38 Chapter 3. Quantum field theory

    Φn(z) for n 6= 0 are the n-th side modes with momentum n~qΦn(z) = Φ0(z)e

    inq·z. (3.28)

    and ĉm are bosonic operators obeying the commutation relations

    [ĉn, ĉ†n′ ] =

    ∫d3zΦ∗n(z)Φn′(z) = δnn′ [ĉn, ĉn′ ] = 0 (3.29)

    We are assuming that the states Φn(z) form a complete orthonormal system. In

    general this is non true, as the overlap integrals

    〈Φn|Φm〉 =∫

    d3zΦ∗n(z)Φm(z) =∫

    d3z|Φ0(z)|2ei(n−m)q·z (3.30)are not zero for n 6= m and are not 1 for n = m. For most condensate sizes and trapconfigurations, however, these integrals are many orders of magnitude smaller than

    unity. As a result, for typical condensate, the orthogonality approximation yields ac-

    curate results. By properly taking into account the non-orthogonality of the atomic

    field modes, it can be shown that the only surviving effect in the linearized theory

    (see next section) is the modification of the atomic polarization term in the equa-

    tion of motion for the probe field (3.25) to include a second scattering mechanism

    in which a condensate scatters a photon without changing its center of mass state.

    As a consequence of momentum conservation, this process is suppressed by a factor

    〈Φn0|Φn0−1〉 relative to the process which transfers the atom from the condensate inthe state n0 to the side mode state n0− 1. Bose enhancement, on the other hand, isstronger for this transition by a factor

    √N , because we now have N identical bosons

    in both the initial and final states. Thus it is the product√

    N〈Φn0|Φn0−1〉 whichmust be negligible if we have to make the orthogonality approximation.

    From Eq. (3.26) the atomic field operator which annihilates an atom in the n-th

    condensate side mode is defined

    ĉn =

    ∫d3zΦ∗n(z)Ψ̂(z) (3.31)

    Taking the derivative with respect to time and substituting Eq.(3.24) we obtain

    i~d

    dtĉn =

    n2(~q)2

    2mĉn + ~

    ( |Ω|24∆20

    +g21

    ∆20â†â

    )ĉn + i~g

    (â†ĉn+1 − âĉn−1

    )

    +∑m

    ĉm

    ∫d3zΦ∗n(z)e

    inq·z(−~2∇2

    2m+ V (z)

    )Φ0(z)

    +β∑

    m,k,l

    ĉ†mĉkĉl

    ∫d3zΦ∗nΦ

    ∗m(z)Φk(z)Φl(z) (3.32)

  • 3.2. Coupled-modes equations 39

    and inserting the Gross-Pitaevskii Eq. (3.27) finally

    i~d

    dtĉn =

    n2(~q)2

    2mĉn + ~

    ( |Ω|24∆20

    +g21

    ∆20â†â

    )ĉn + i~g

    (â†ĉn+1 − âĉn−1

    )

    −βN∑m

    ĉm

    ∫d3zΦ∗n(z) |Φ0(z)|2 Φm(z)

    +β∑

    m,k,l

    ĉ†mĉkĉl

    ∫d3zΦ∗n+m |Φ0(z)|2 Φl+k(z) (3.33)

    Substituting Eq. (3.26) in Eqs. (3.25) we obtain for the probe field operator

    dâ

    dt= i∆21â + g

    ∫d3zΨ̂ †g (z)e

    −iq·z Ψ̂g(z). (3.34)

    The source of the field equation (3.34) is the bunching operator

    B̂ =

    ∫d3zΨ̂ †g (z)e

    −iq·z Ψ̂g(z) (3.35)

    If we consider an ideal condensate with a constant atomic density the ground state

    is independent from position variables Φ0(z) = Φ0 = 1/√

    V and eq.(3.33) takes the

    simpler form

    i~d

    dtĉn =

    n2(~q)2

    2mĉn + ~

    ( |Ω|24∆20

    +g21

    ∆20â†â

    )ĉn + i~g

    (â†ĉn+1 − âĉn−1

    )

    −βNV

    ĉn +β

    V

    m,k

    ĉ†mĉkĉn+m−k (3.36)

    and now the bunching operator is given by

    B̂ =∞∑

    n=−∞ĉ†nĉn+1 (3.37)

    Generally in experiments that involves the CARL mechanism, like for example

    superradiant Rayleigh scattering, the laser pulse is applied when the trap is com-

    pletely switched off and the condensate is in expansion, so it can be interesting to

    study the model when the effects of trap potential and of collisions are negligible.

    In this regime we get to the following model for the coupled modes

    d

    dtĉn = −iωRn2ĉn + g

    (â†ĉn+1 − âĉn−1

    )(3.38)

    dâ

    dτ= i∆21â + g

    ∞∑n=−∞

    c†ncn+1. (3.39)

  • 40 Chapter 3. Quantum field theory

    We note that Eqs.(3.38) and (3.39) conserve the number of atoms, i.e.∑

    n ĉ†nĉn = N ,

    and the total momentum, Q̂ = â†â +∑

    n nĉ†nĉn. Defining the operators %̂m,n = ĉ

    †mĉn

    from Eq.(3.38) we derive

    d

    dt%̂mn = iωR(m

    2 − n2)%̂mn+g

    {â (%̂m+1,n − %̂m,n−1) + ↠(%̂m,n+1 − %̂m−1,n)

    }(3.40)

    Taking the expectation values for the operators, taking scaled variables and with

    the substitution A =√

    2/ρN〈â〉e−iδτ , Eqs. (3.40) and (3.39) are equivalent to Eqs.(2.16) and (2.17) introduced with first quantization in chapter 2. A more realistic

    and complete model should take into account even effects of atomic decoherence

    and cavity losses. We will see possible approaches to this problem in some of the

    next chapters (see chapters 5 and 8), modifying the model in the proper way for the

    considered situation.

    3.3 Linearized three-mode model

    From Eq. (3.40), we see that the first-order side modes are optically coupled to both

    the condensate mode and to second-order side modes. For times short enough that

    the condensate is not significantly depleted, the coupling back into the condensate

    is subject to Bose enhancement due to the presence of ∼ N identical bosons in thismode. The coupling to the second-order side mode, in contrast, is not enhanced.

    Hence for these time scales, the higher-order side modes are not expected to play a

    significant role. These arguments suggest developing an approach where, assuming

    that all N atoms are initially in the condensate mode n0 with momentum n0~q, thethree atomic field operators ĉn0 , ĉn0−1, and ĉn0+1 play a predominant role. Therefore,

    we expand the atomic field operator as

    Ψ̂g(z) = 〈z|Φn0〉ĉn0 + 〈z|Φn0−1〉ĉn0−1 + 〈z|Φn0+1〉ĉn0+1 + ψ̂(z), (3.41)

    where the field operator ψ̂(z) acts only on the orthogonal complement to the sub-

    space spanned by the state vectors |Φn0〉, |Φn0−1〉, and |Φn0+1〉. As a result, ψ̂(z)commutes with the creation operators for the three central modes.

    With the assumption of negligible condensate depletion we can simply drop the

    operator ĉn0 substituting it with its mean value 〈ĉn0〉 ≈√

    Ne−in2τ/ρ. The system is

  • 3.4. Concluding remarks 41

    unstable for certain values of the detuning ∆. In fact, by linearizing Eqs.(3.38) and

    (3.39) around the equilibrium state, the only equations depending linearly on the

    radiation field are those for ĉn0−1 and ĉn0+1. Hence, in the linear regime, the only

    transitions involved are those from the state n0 towards the levels n0−1 and n0 +1.With respect to scaled variables and introducing the operators

    â1 = ĉn0−1ei(n20τ/ρ+∆τ) (3.42)

    â2 = ĉn0+1ei(n20τ/ρ−∆τ) (3.43)

    â3 = âe−i∆τ , (3.44)

    Eqs.(3.38) and (3.39) reduce to the linear equations for three coupled harmonic

    oscillator operators:

    dâ†1dτ

    = −iδ−â†1 +√

    ρ/2â3 (3.45)

    dâ2dτ

    = −iδ+â2 −√

    ρ/2â3 (3.46)

    dâ3dτ

    =√

    ρ/2(â†1 + â2), (3.47)

    with Hamiltonian

    Ĥ = δ+â†2â2 − δ−â†1â1 + i

    √ρ

    2[(â†1 + â2)â

    †3 − (â1 + â†2)â3], (3.48)

    where δ± = δ ± 1/ρ and δ = ∆ + 2n0/ρ = (ω2 − ω1 + 2n0ωR)/ρωR. Hence, thedynamics of the system is that of three parametrically coupled harmonic oscillators

    â1, â2 and â3 [49], which obey the commutation rules [âi, âj] = 0 and [âi, â†j] = δij

    for i, j = 1, 2, 3. Note that the Hamiltonian (3.48) commutates with the constant of

    motion

    C = â†2â2 − â†1â1 + â†3â3. (3.49)We will solve exactly this model in chapter 6.

    3.4 Concluding remarks

    We have deduced an appropriate quantum field theory that extends into the ultra-

    cold regime of BEC the CARL model, so that the unique coherence properties of

    the condensates might be further understood and exploited by the interaction with

  • 42 Chapter 3. Quantum field theory

    dynamical light fields. In the limit of no collisions and consider