Double Bragg diffraction: A tool for optics

E. Giese,1 A. Roura,1 G. Tackmann,2 E. M. Rasel,2 and W. P. Schleich1 1Institut f¨urQuantenphysik and Center for Integrated Quantum Science and Technology IQST , Universit¨atUlm, Albert-Einstein-Allee 11, D-89081, Germany. 2Institut f¨urQuantenoptik, Leibniz Universit¨atHannover, Welfengarten 1, D-30167 Hannover, Germany. The use of retro-reflection in light-pulse atom interferometry under microgravity conditions nat- urally leads to a double-diffraction scheme. The two pairs of counterpropagating beams induce simultaneously transitions with opposite momentum transfer that, when acting on initially at rest, give rise to symmetric interferometer configurations where the total momentum transfer is automatically doubled and where a number of noise sources and systematic effects cancel out. Here we extend earlier implementations for Raman transitions to the case of Bragg diffraction. In contrast with the single-diffraction case, the existence of additional off-resonant transitions between resonantly connected states precludes the use of the adiabatic elimination technique. Nevertheless, we have been able to obtain analytic results even beyond the deep Bragg regime by employing the so-called “method of averaging,” which can be applied to more general situations of this kind. Our results have been validated by comparison to numerical solutions of the basic equations describing the double-diffraction process.

I. INTRODUCTION ability to manipulate light. It should be noted that all cases mentioned so far involve static potentials (or effec- There has recently been a growing interest in the tive potentials) for the dynamics of the matter waves, possibilities that (ultra)cold atoms in microgravity offer which implies that the kinetic energy (the modulus of for long-time interferometry and its application to high- the wave-vector) before and after scattering remains the precision measurements [1–4]. This has contributed to same. Moreover, the duration of the interaction is de- stimulating the development of compact devices [5–8], termined by the transverse velocity and thickness of the which are typically necessary for such environments and (light) crystal/grating. In this respect one should distin- are also important for their use in navigation, geodesy, guish between thick and thin gratings. The former, which and geophysics, to name a few [9–11]. In this article we correspond to the so-called Bragg regime, exhibits high put forth a generalization of conventional Bragg diffrac- momentum selectivity and only one non-trivial diffrac- tion in light-pulse atom interferometry which can be a tion order for resonant momenta [18, 22]. In contrast, for valuable technique for compact set-ups and microgravity thin gratings (corresponding to the Raman-Nath regime, applications. also known as Kapitza-Dirac for light gratings) there is little momentum selectivity and many diffraction orders are populated [17]. An intermediate regime where ve- A. Overview of atom interferometry and Bragg locity selectivity is somewhat relaxed but diffraction still diffraction takes place mainly to a single order, sometimes known as “quasi-Bragg” regime, has been studied in Ref. [23] and The diffraction of X rays by crystals studied a century will also be of particular interest for us. ago by Bragg [12] and Laue [13] is a beautiful manifes- Laser cooling to sub-Doppler temperatures [24, 25] tation of the wave nature of this high-frequency electro- made the advent of light-pulse atom interferometry [26– magnetic radiation as well as of the periodic structure 28] possible. It is based on time-modulated laser pulses of crystals, and it has played a central role in character- that drive Rabi oscillations between different momentum izing the microscopic structure of a wide range of ma- states (possibly entangled to different internal states) and terials, including complex proteins and DNA. A simi- whose duration and intensity can be adjusted to act as lar phenomenon involving the diffraction of neutrons by beam splitters (π/2 pulses) or mirrors (π pulses). The crystals was exploited to build some of the first matter- laser beams employed are long and wide enough so that arXiv:1308.5205v2 [physics.atom-ph] 11 Nov 2013 wave interferometers and gave rise to the rich field of there is a fixed momentum transfer along the longitudi- neutron interferometry [14, 15]. Material gratings have nal direction of the beam and only the along this also been used to diffract beams of atoms and molecules dimension matters, while transverse velocities remain un- [16]. An interesting alternative approach relies on the use changed. Despite such fixed momentum transfer, a more of standing electromagnetic waves as phase gratings [17– or less narrow momentum band around resonant states 19] or absorption gratings [20, 21], depending on whether will be diffracted [29]. This is because for pulses with resonant or nonresonant radiation is employed. These finite duration differences between initial and final ki- set-ups, still based on the interaction between matter netic energies (which can be larger for shorter pulses) and light but with their roles reversed compared to tradi- are allowed, as can be understood from Heisenberg’s un- tional optical interferometers, provide high-quality grat- certainty relation for time and energy. The role of pulse ings with controllable properties as a consequence of our duration is then analogous to the crystal thickness men- 2 tioned above. B. Microgravity environments and double Bragg The first implemented scheme for light-pulse interfer- diffraction ometry [27] and widely used to date relies on two-photon Raman transitions induced by a pair of counterpropa- The use of a retro-reflection geometry, where the two gating lasers with different frequencies. These are transi- laser beams reach one side of the interferometer set-up tions between different internal states entangled to differ- through a common optical fiber and reflect off a mir- ent momenta [26], and the internal-state labeling allows ror at the other side in such a way that the reflected the read-out of the exit ports even when they spatially beams are aligned with the incident ones (giving rise to overlap. Moreover, the velocity selection effect can be two pairs of counterpropagating beams), is beneficial for somewhat relaxed by using shorter pulses with higher a number of reasons. Firstly, most vibration effects on intensity. This together with the ability to deal with the laser phases are common to the two lasers and cancel substantially overlapping clouds for the two exit ports, out in a two-photon process and only mirror vibrations reduces the requirement of very narrow initial momen- have an effect on processes involving pairs of counter- tum distributions (well below recoil velocities) and cold propagating beams. Secondly, this geometry reduces the thermal atoms obtained from optical molasses can be em- undesired effects of wave-front distortions, which become ployed without the need for evaporative cooling. more important for larger effective momentum transfer An alternative scheme for light-pulse interferometry is and longer interferometer times, since those cancel to based on Bragg diffraction and involves transitions be- first order provided that no additional distortions are tween states with different momenta without changing generated while the beams propagate towards the mirror the internal state. These are induced by a pair of coun- and return. Retro-reflection is, therefore, commonly em- terpropagating lasers with wave numbers k1 and −k2, ployed in atomic fountains for high-precision interferome- leading to a total momentum transfer of ~K ≡ ~k1 +~k2, try [35, 36]. In that case the non-vanishing velocity of the and frequencies slightly detuned to account for the recoil atoms along the direction of the lasers selects one of the energy. There is always a frame where the frequencies two pairs of counterpropagating beams, while the other of the two beams are the same and one has a standing remains off-resonant. However, achieving much higher wave. In this frame only a narrow band of momentum sensitivities requires longer interferometer times and suf- states around the two resonant momenta ±~K/2 will be ficiently extended times are only possible in microgravity diffracted and there is a fixed momentum transfer ∓~K environments [2–4]. Using retro-reflection in micrograv- respectively. The process is then analogous to Bragg ity with narrow initial momentum distributions naturally diffraction in crystals (except that the duration of the leads to a double-diffraction scheme where the action of interaction is determined by the pulse duration). It is the first beam-splitter pulse on atoms initially at rest cre- possible to select other resonant momenta by detuning ates a superposition of two states with opposite momenta the frequencies while keeping the total momentum trans- +~K and −~K since each one of the two pairs of coun- fer ~K fixed, which can be understood as changing to a terpropagating beams induces a resonant transition with different frame where the atoms have the desired initial opposite momentum transfer. Besides doubling the total velocity. These kind of interferometers have been widely momentum transfer, this leads to a symmetric interfer- applied to Bose-Einstein condensates (BECs) [3, 30–32], ometer where a number of systematic effects cancel out, whose narrower momentum distribution (significantly be- including those due to laser-phase noise and those involv- low recoil velocity) allows good spatial separation of the ing terms proportional to K2. This kind of symmetric exit ports and a high diffraction efficiency: For a π pulse interferometer based on a double-diffraction scheme was most atoms are diffracted and only one diffraction order first implemented for Raman transitions in Ref. [37]. (See is populated. (Note that contrary to the Raman case, Ref. [38] for a different proposal employing also two pairs trying to relax the effect of velocity selection by using of counterpropagating Raman beams.) In that case, the shorter pulses with higher intensity can only be done to symmetric configuration led to the added benefit of a re- some extent, corresponding to the quasi-Bragg regime duction of AC Stark shift effects or any other effects act- [23], because one will otherwise start to populate other ing differently on internal states since the atoms are at all diffraction orders [33].) Using a condensate is not manda- times in the same internal state for both interferometer tory since one can apply initially a long velocity-selective branches, in contrast with the single-diffraction scheme. Raman π pulse [34], but this reduces significantly the Note also that, although very natural under microgravity number of atoms available. On the other hand, an ad- conditions, double diffraction can also be employed in a vantage of interferometry based on Bragg diffraction is gravitational field and for non-vanishing initial velocities that having the same internal state in both interferometer provided that the laser beams are transverse to the mo- branches reduces a number of systematic effects and noise tion of the (undiffracted) atoms, as in the gyroscope set- sources. Moreover, with sufficient laser intensity one can up of Ref. [37]. Double Raman diffraction has also been increase the effective momentum transfer to a multiple employed in a gravimeter, where three different injected of ~K (and with that the sensitivity of the interferom- laser frequencies are necessary to account (via appropri- eter) by adjusting the frequencies so that the resonant ate frequency chirping) for the changing Doppler shift of condition corresponds to a higher diffraction order [34]. the accelerated atoms [39]. 3

In this article we analyze in detail the extension of the Furthermore, for a double duration of the pulse it acts double-diffraction scheme to the case of Bragg scattering, as a mirror, exchanging states with momentum +n~K to which is particularly well suited for interferometry with −n~K and vice versa. BECs, as already mentioned above. Some efforts in this The existence of two pairs of counterpropagating direction have already been made experimentally [40], beams results in a much wider range of possibilities for but not in connection with atom interferometry nor fo- higher-order processes relating any two given momen- cusing on the special properties of the double-diffraction tum eigenstates. In particular there are additional off- scheme and the much richer dynamics associated with it. resonant transitions between resonantly connected states. The double Bragg diffraction scheme involves a pair of This precludes the use of the standard method of adia- slightly detuned lasers (with frequency difference corre- batic elimination of nonresonant states [43–45]. Never- sponding to the recoil energy) retro-reflected off a mirror, theless, one can still obtain analytical results employing and it is crucial that only two-photon processes involv- the so-called “method of averaging” described in the next ing counterpropagating beams (with unequal frequen- subsection, which can be applied to more general situa- cies) take place, while those associated with copropagat- tions of this kind, provided that one has two sufficiently ing beams should be entirely suppressed. This is achieved different frequency scales, as controlled by an adiabatic- by injecting the two lasers with orthogonal polarizations ity parameter ε. Moreover, because of the additional pair (that can be either circular or linear [41]) and inserting a of counterpropagating beams, it is no longer possible to quarter-wave plate in front of the mirror, which guaran- find a reference frame where one simply has a standing tees that each reflected beam has a polarization orthog- wave. onal to the incoming one. Similarly to double diffraction The aspects mentioned in the previous paragraph give for Raman processes, the beam splitter creates an equal- rise to the following new features, absent in single diffrac- amplitude superposition of +~K and −~K states, which tion: leads to a symmetric interferometer configuration with similar desirable properties. The use of Bragg scatter- • Fast oscillations with smaller amplitude superim- ing has certain advantages compared to set-ups based on posed on the slow generalized Rabi oscillations be- Raman transitions, even for a double-diffraction scheme. tween resonant states. In contrast to single diffrac- Firstly, it is easier to implement experimentally because tion, their amplitude is of order ε (when considering the small frequency detuning that is required (of the or- square pulses) rather than ε2, and results from the der of tens of kHz) can be achieved with a single laser interference between slow and fast contributions to and acousto-optic modulators (AOMs) rather than two the dynamics of resonant states. phase-locked lasers as required for the Raman case, where the frequency difference is typically of several GHz. This • A non-trivial AC Stark shift associated with higher- can be particularly advantageous for compact devices. order processes. For single Bragg diffraction one Secondly, one can use higher-order Bragg diffraction to can easily argue that the two resonant states ex- increase the effective momentum transfer to n~K while perience the same shifts by considering the frame keeping all the advantageous features of double diffrac- where one has a standing wave and where the sit- tion and the associated symmetric interferometer con- uation is symmetrical for both momentum states. figurations. Thus, in contrast to the double-diffraction As mentioned above, this is no longer possible for scheme based on Raman transitions, where one can only two pairs of counterpropagating beams. do so by introducing additional π pulses [37], in the Bragg case one could use an optimized combination of both techniques to maximize the total amount of effective mo- D. The method of averaging mentum transfer, as already done for single diffraction [42]. The dynamics of momentum states differing by mul- tiples of ~K which are resonantly or nonresonantly con- nected by 2n-photon transitions can be treated analyti- C. Special features of double Bragg diffraction cally when there is a large separation of scales between the frequencies associated with the two kinds of processes We highlight here several important aspects that differ (resonant vs. nonresonant). This separation of scales is from single diffraction as well as new features that were controlled by an adiabaticity parameter ε given by the absent in that case. ratio of those two frequencies (in our case, the Rabi First of all, the resonantly connected states form a frequency for two-photon transitions and the recoil fre- three-level system and one has generalized Rabi oscilla- quency). When this parameter is small, one can study tions between them rather than the usual Rabi oscilla- in a controlled manner the slow and fast contributions to tions for two-level systems. Nevertheless, a pulse with an the dynamics as well as their mutual influence following appropriate duration and laser intensity can still act as the “method of averaging” introduced in Ref. [46]. The a beam splitter: It evolves a zero-momentum state into fast contributions, whose dynamics is modulated by the an equal superposition of +n~K and −n~K momenta. slow part, have a small amplitude and can be organized 4 as an expansion in powers of ε. In turn, the fast contribu- uations of Sec. III and Sec. IV. Hence, the derivation and tions can combine resonantly with fast oscillating terms presentation of the formalism makes a physical interpre- in the Hamiltonian and give corrections to the dynamics tation easier than the rather mathematical presentation of the slow parts. Proceeding in this way, one can write in Ref. [46]. down a set of coupled equations for the slow and fast This method is a procedure to approximately solve terms which are equivalent at any desired order in ε to coupled differential equations with time-dependent coef- the original equations for the full solution (involving the ficients that oscillate at different frequencies. Thus, it is sum of slow and fast terms). the method of choice for atomic Bragg diffraction prob- The method just described can deal with situations lems, since they have exactly this form. where both slow and fast terms contribute to the same degree of freedom. This is a new feature of double Bragg diffraction which is not present in the single-diffraction A. General idea case and cannot be dealt with using the standard adia- batic elimination technique. It is, thus, an approach that Let us assume a system of n coupled differential equa- allows a controlled and systematic way of analyzing the tions of the form so-called “quasi-Bragg” regime [23], where ε is sufficiently small but gives rise to non-negligible corrections, and ob- X iνωrt g˙ = iεH g ≡ iεH0 g + iε e Hν g , (1) taining analytic solutions. It should be noted that the ν=0 basic equations describing the dynamics of momentum 6 n states in double Bragg diffraction that we have derived with g ∈ C , time-independent quantities H, H0, Hν ∈ n n are valid in a more general regime (e. g. for ε close to C × , with the dimensions of a frequency, a frequency one), but then they need to be solved by other means, ωr, and a dimensionless parameter ε  1. We refer to for instance numerically. this parameter as the adiabaticity parameter. When we discuss single Bragg diffraction and double Bragg diffrac- tion, the physical meaning and relevance of this parame- E. Outline of the article ter becomes clearer. In the form of Eq. (1), the hierarchy of increasingly faster oscillating terms with frequencies Our article is organized as follows. We start by review- νωr is manifestly laid out. We see later that ν is associ- ing the method of averaging in Sec. II, where we introduce ated with a certain order of a Bragg transition. a notation which is adapted to the one that we use for the We now separate the solution into slow and fast oscil- examples analyzed in the present article. In that section, lations; i. e., we assume a function the method is explained up to second order, whereas ar- m bitrary orders are discussed in Appendix A. This techni- (m) (m) X j (m) cal part is the basis for the following two applications. In g = γ + ε f j(γ ) , (2) Sec. III we first recall the physical process of single atomic j=1 Bragg diffraction and then discuss the connection of the which satisfies the differential equation method of averaging to the ordinary adiabatic elimina- tion [43–45] and the quasi-Bragg regime [23]. Next, we g˙ (m) = iεH g(m) + O εm+1 . (3) turn in Sec. IV to the more sophisticated case of double Bragg diffraction, where the conventional adiabatic elim- (m) The functions f j = f j(t, γ ) depend not only on the ination is not possible and the method of averaging needs (m) to be applied. Various features of this scattering process on the slowly evolving term γ but are explicitly time- are discussed using the insights gained by the method dependent. of averaging. In that section we restrict ourselves to the This scheme is not a perturbative treatment in the con- case of circularly polarized light waves and square pulses. ventional way. In fact, the method of averaging finds a (m) A situation with linearly polarized lasers and an under- solution g that satisfies the original differential equa- m lying magnetic field is discussed in Appendix B and leads tion up the order ε . to the same results. Furthermore, the extension of our At each order, we assume that the slowly evolving part (m) results to time-dependent laser pulses is briefly described γ fulfils the equation in Appendix C. Finally, we conclude in Sec. V with a dis- m cussion of our results. (m) (m) X µ (m) γ˙ = iεH0 γ + i ε pµ(γ ) . (4) µ=2

II. METHOD OF AVERAGING Using this ansatz together with differentiation of Eq. (2), one can equate the coefficients of different orders of ε on We now summarize the method of averaging first in- the left side of Eq. (3) to the right side. This way, one troduced by Ref. [46]. Our notation is quite different can determine the functions f j and pj. We now derive from Ref. [46] and already adapted to the scattering sit- these conditions. 5

Taking the time derivative of Eq. (2) yields When we use the ansatz Eq. (4) for the time derivative of γ(m), we find the equation m m (m) X ∂f j X ∂f j(γ ) ˙ (m) ˙ (m) j j ˙ (m) g =γ + ε + ε (m) γ . ∂t (m) ∂γ j=1 γ j=1

! ! ∂f ∂f (γ(m)) ∂f ˙ (m) H (m) 1 2 (m) 1 H (m) 2 g =ε i 0 γ + + ε i p2(γ ) + i (m) 0 γ + ∂t γ(m) ∂γ ∂t γ(m) m j 2 ! (m) − (m) X j (m) ∂f j 1(γ ) (m) ∂f j X ∂f µ(γ ) (m) m+1 − H O + ε i pj(γ ) + i (m) 0 γ + + i (m) pj µ(γ ) + ε . ∂γ ∂t (m) ∂γ − j=3 γ µ=1

Now we compare this expression to the one on the right- For m = 1, i. e. up to the first approximation, Eq. (4) hand side of Eq. (3) using the ansatz Eq. (2), i. e. to reads

(m) m+1 (m) 2 (m) (1) (1) iεH g + O ε = iεH γ + iε H f 1(γ ) γ˙ = iεH0γ (7) m X j (m) m+1 with a time-independent H . Hence, we find the slowly + i ε Hf j 1(γ ) + O ε , 0 − j=3 evolving solution

(1) (1) we find conditional equations for f j. In this way, we can γ (t) = exp [iεH0t] γ (0) , (8) determine Eq. (2). where we have defined the matrix exponential exp [A] ≡ P n n=0 A /n! for square matrices A. B. First-order solutions With Eq. (6), we can calculate rapidly oscillating cor- rections to this slow solution. The full expression reads Up to the order of ε, we find with the help of H = for m = 1 with Eq. (2) P iνωrt H0 + ν=0 e Hν the condition   6 iνωrt (1) 1 X e (1) ∂f g (t) =  + ε Hν  γ (t) , (9) 1 X iνωrt (m) νωr = i e Hν γ . (5) ν=0 ∂t (m) 6 γ ν=0 6 1 where we used the identity ≡ δn,n0 . Here, we see that We can integrate this equation easily, since due to the the fast terms are suppressed with ε  1. (m) partial derivative with respect to time the function γ To ensure that the initial conditions are fulfilled, we can be treated as a constant, and find assume

Z X X eiνωrt   1 (m) iνωrt H (m) H (m) − f 1(γ ) = dt i e ν γ = ν γ . (1) X Hν (1) νωr 1 ν=0 ν=0 γ (0) =  + ε  g (0) . (10) νωr 6 6 ν=0 (6) 6

When this integration is performed, the initial condition In the following, we refer to choosing the initial condition (1) yields constants of integration, which may be functions γ (0) as the dressed-state formulation. If we choose (1) of γ(m) but are not explicitly time-dependent. Thus, it g (0) as an initial condition, we call this the bare-state would just yield a slowly evolving term. In our formula- formulation. We do not discuss the interpretation of ini- tion we set this constant equal to zero, since we want a tial condition here in more detail but refer to Sec. III D, clear separation of timescales. This approach is valid if where we examine this point using the example of single we ensure that the initial conditions are fulfilled, as we Bragg diffraction. do later in this section. In the integration we see the link to the conventional adiabatic elimination: Just the rapidly oscillating terms C. Second-order solutions are integrated, the slowly evolving term γ(m) is constant in time. This procedure is very close to the adiabatic To get solutions for m ≥ 1, the next step is to compare elimination discussed in Sec. III B in more detail. the coefficients to the order of ε2. With the solution 6

Eq. (6) for f 1 we find the differential equation III. BRAGG

(m) iνωrt ∂f 2(γ ) Xe (m) (m) = i [HHν − Hν H0] γ −i p2(γ ). The method of averaging presented so far can be ap- ∂t νωr ν=0 plied to a large class of problems. In particular, it 6 When defining gives us the opportunity to gain more insight into Bragg diffraction processes. Even though it is necessary for the (2) X Hµ ν Hν HµH0 double-diffraction case, one can also apply it to single Φµ ≡ − − (11) νωr µωr diffraction. In this approach, the distinction between the ν=0 6 deep Bragg regime and the so-called quasi-Bragg regime for µ =6 0 and as investigated in Ref. [23] comes out clearly. Moreover, (2) X H ν Hν the methods and approximations used in this chapter can Φ0 ≡ − , be generalized to the much more complex double Bragg νωr ν=0 6 diffraction case. As an introduction we now discuss the this differential equation takes the form familiar Bragg diffraction from this point of view. (m) ∂f 2(γ ) X = i eiµωrt Φ(2) γ(m) − i p (γ(m)) , (12) ∂t µ 2 A. Model µ

P iνωrt where we have used again H = H0 + ν=0 e Hν . We assume a two-level atom with an energy separa- 6 We emphasize that in the sum in Eq. (12) the term tion ~ωeg interacting with two counterpropagating light µ = 0 is included, i. e. a time-independent term appears waves, one with frequency ωa, and the other with ωb. where the phase factor is equal to unity. On the other Both light fields are aligned parallel to the z-direction. side, the function p2, describing a slowly evolving term This set-up is depicted in Fig. 1. as one can see from Eq. (4), appears in the differential equation and thus has to be also independent of time. So ωb ωa (m) (2) (m) for p2(γ ) ≡ Φ0 γ , Eq. (12) takes exactly the form of Eq. (5), namely z atom (m) ∂f 2(γ ) X iµωrt (2) (m) = i e Φµ γ , ∂t FIG. 1. Schematic set-up of Bragg diffraction. A two-level µ=0 6 atom is interacting with two counterpropagating light fields where just rapidly oscillating terms occur. This approach of frequency ωa and ωb, which are far detuned from the atomic leads, in complete analogy to the previous section, to transition.

iµωrt (m) X e (2) (m) f 2(γ ) = Φµ γ , Thus, the atom is interacting with the electric field µωr µ=0 i(kbzˆ ωbt) i( kazˆ ωat) 6 Eˆ = Eb e − +Ea e − − + h.c. , (14) where again the constant of integration was set to zero. With this choice we get for m = 2 from Eq. (4) the with the amplitudes Ea and Eb and the absolute values differential equation ka and kb of the wave vectors. The sign in front of ka in the second exponential reflects the fact that the light is (2) (2) 2 (2) (2) γ˙ =iεH0γ + iε Φ0 γ , travelling in the opposite direction. ˆ where we have time-independent coefficients and hence The field operator E describes two classical fields, but find accounts for the mechanical action of the light on the atom, i. e. the recoil. Indeed, the operator (2) h 2 (2) i γ (t) = exp iεH0t + iε Φ0 t γ2(0) Z ikzˆ     e± = dp |p ± ~ki hp| X H ν Hν (2) = exp iε H0 + ε −  t γ (0) . νωr shifts the momentum by ± k. Hence, we can change the ν=0 ~ 6 momentum of the atom by applying these lasers. This (13) can be understood in terms of momentum conservation: Since we have already determined f 2, we could also ob- The atom in the ground state |gi absorbs one photon of tain the rapidly oscillating corrections. Because we dis- momentum ~kb. The atom is now in the excited state cuss in the following sections the second approximation |ei and has picked up the recoil ~kb. An emission of without fast corrections, we refrain from presenting this a photon with momentum −~ka may be stimulated by cumbersome expression. the other laser and lead to a total transfer of the atomic The procedure described in this section can be ex- momentum by ~K ≡ ~(kb + ka), where K is called the tended step by step to arbitrary orders of ε, as we show effective wave vector of the Bragg pulse. This process is in Appendix A. shown in Fig. 2. 7

|ei B. Three-term recurrence relation in adiabatic approximation ∆ In this section, we derive a three-term recurrence re- ωeg |gi lation for the momentum distribution of atoms in the ground state. For this, we perform the conventional adi- abatic approximation, as e. g. explained in Refs. [43, 44]. 2 4ωr ∼ p We use in rotating wave approximation, e. g. see Ref. [47], 2ωr ωr the Hamiltonian 0 2 ˆ pˆ H = + ~ωeg |ei he| |−2~Ki|−~Ki|0i |~Ki |2~Ki p 2M nh i o i(kbzˆ ωbt) i( kazˆ ωat) + ~ Ωb e − +Ωa e − − |ei hg| + h.c. , FIG. 2. Momentum transfer and in Bragg diffrac- tion. The absorption of a photon from one field and the emis- sion of a photon into the field travelling in the opposite di- where we have assumed a interaction of the ˆ ˆ rection results in a recoil of the value ~K ~(kb + ka). This atomic dipole moment d with the electric field E deter- process is resonant, if the difference between≡ the two lasers mined by Eq. (14). The Rabi frequencies are defined as ∆ω = ωb ωa is equal to the change of kinetic energy, which ≡ − h | ˆ · | i − 2 Ωj e d Ej g /~ and the atom has the mass M. is proportional to p . The dashed transitions are off-resonant We emphasize that we consider throughout this article and suppressed; the gray line denotes the sum of potential and 2 a regime where we have a large detuning and the effects kinetic energy of the virtual excited state. ωr = ~K /(2M) denotes the recoil frequency. of can be neglected. An approach to Bragg diffraction taking the effects of recoil due to spontaneous emission into account can be found e. g. in Since in typical Bragg diffraction a superposition of Refs. [48–50]. different momenta in the ground state is created, the We now allow this Hamiltonian to act on an arbitrary population of the excited state needs to be virtual. As state a result, both light fields have to be far detuned from Z the atomic ωeg. The detuning is defined as |ψi ≡ dp [g(p) |g, pi + e(p) |e, pi] , ∆ ≡ ωeg − ωb ≈ ωeg − ωa, as shown in Fig. 2. This condition is essential for the adiabatic elimination of the apply the Schr¨odingerequation excited state that is performed in the next section and leads to a two-photon process. d Figure 2 shows an additional feature of the Bragg scat- i~ |ψi = Hˆ |ψi (16) dt tering process. Only if the difference ∆ω ≡ ωb − ωa of the two frequencies of the light waves is equal to the to equate the coefficients in front of |g, pi and |e, pi, and difference of the kinetic energy gained by the momen- find the coupled differential equations tum transfer, this process is resonant, i. e. each resonant process has to start and end on the kinetic parabola i(ωeg ωb)t i(ω k +ν k )t ie˙(p) =Ωb e − e− − b − b g(p − ~kb) p2/(2M), where M denotes the mass of the atom and i(ωeg ωa)t i(ωk +νk )t p the momentum. Other processes are of higher order + Ωa e − e− a a g(p + ~ka) (17) and thus suppressed. For example, the dashed lines in i(ωeg ωa)t i(ω k +ν k )t ig˙(p) =Ω∗ e− − e− − a − a e(p − ~ka) the figure represent such diffraction orders. Usually, all a i(ωeg ωb)t i(ωk +νk )t off-resonant momentum states can be adiabatically elim- + Ωb∗ e− − e− b b e(p + ~kb) . (18) inated as well. We now demonstrate that the method of averaging provides a convenient way to calculate the Here, we have already transformed into the interac- populations of these levels. tion picture in which e(p) has to be multiplied by  2   2  The resonance condition can be written in a general exp i(p /(2M~) + ωeg)t and g(p) by exp ip /(2M~)t . form: For the j-th resonance, we have j two-photon pro- In the equations above, we have introduced the frequency 2 cesses, i. e. the gain of energy due to the lasers is j∆ω. If ωk ≡ ~k /(2M), which corresponds to the kinetic en- this energy coincides with the difference of the kinetic en- ergy of an atom with momentum ~k and the frequency ≡ ergy caused by the momentum transfer j~K, the process νk pk/M, which accounts for the Doppler effect. is on resonance. This condition reads In the following, we simplify this set of differential equations by eliminating the excited state adiabatically. 1 ~(jK)2 ~K2 We assume the change of the population amplitude of the ∆ω = = j ≡ jωr , (15) j 2M 2M ground state g(p) and the frequencies ωk +νk to be much smaller than the detuning ∆ = ωeg − ωa,b and thus find where ωr is called the recoil frequency. by only integrating the exponentials exp [i(ωeg − ωa,b)t] 8 in Eq. (17) the approximate expression anymore, but with three levels. Nevertheless, effective Rabi oscillations occur and with our definition of the fre- Ω ∼ b i(ωeg ωb)t i(ω k +ν k )t quency we find here a more intuitive connection. e(p) = − e − e− − b − b g(p − ~kb) ∆ The value of ∆ω now determines the resonances of the

Ωa i(ωeg ωa)t i(ωk +νk )t Bragg diffraction process. According to Eq. (15), if we − e − e− a a g(p + ~ka) (19) ∆ set ∆ω = jωr, the transition from |0i to |j~Ki is on resonance, i. e. the time-dependent phase factors vanish for the excited state. Of course, this approximation is for these states as we see in Eq. (20).  only valid for Ωa,b ∆. We now focus on the first-order Bragg diffraction pro- At this point, we see the connection of the method of cess, which is the process shown in Fig. 2. For that, we set averaging to the adiabatic elimination: In Eq. (6) we also ∆ω = ω and assume Ω to be real. The last assumption perform an integration assuming that the slow solution r (m) is not necessary, but simplifies the notation and calcula- γ is constant in time in comparison to the oscillating tion. To keep track of the laser phases, one can associate terms. In general, the adiabatic elimination is a delicate with ∆ω the respective difference of laser phases. interplay of fast and slow variables. The interpretation With these assumptions we thus get from Eq. (20) the in this case is simple: If the detuning is high, the Rabi system of coupled differential equations oscillations from |gi to |ei are very fast and the popula- i2nω t iν t tion of the excited state is suppressed by Ωa,b/∆. In this g˙(p + n~K) =iΩ e− r e− D g(p + ~(n + 1)K) case the fraction Ω /∆ corresponds to the adiabaticity a,b + iΩ ei2(n 1)ωrt eiνDt g(p + (n − 1)K) parameter. We assume it to be so small that a further − ~ (23) treatment through the method of averaging is not neces- sary. for n ∈ Z, which corresponds to a time-dependent three- We can now use the solution Eq. (19) and substitute term recurrence relation. it into Eq. (18) to find the effective equation

ΩaΩb∗ i∆ωt i(ω +ν )t C. Application of the method of averaging ig˙(p) = − e e− K K g(p + ~K) ∆ Ω∗Ωb a i∆ωt i(ω K +ν K )t In order to cast this equation into the form of Eq. (1), − e− e− − − g(p − ~K) ∆ we first have to introduce the dimensionless adiabatic- i(∆ω ω )t iν t = − Ω e − r e− D g(p + ~K) ity parameter ε. Equation (23) implies that there are i(∆ω+ω )t iν t two timescales: The Rabi frequency Ω and the recoil fre- − Ω∗ e− r e D g(p − ~K), (20) quency ωr. In the spirit of the rotating wave or the adi- where we have defined the effective frequency Ω ≡ abatic approximation, we assume the coupling to higher momentum states, i. e. to higher n, to be suppressed due ΩaΩ∗/∆, the difference ∆ω ≡ ωb − ωa of the laser fre- b to fast oscillating terms. This condition implies that the quencies, and the effective wave vector K = kb +ka. The recoil frequency is defined as recoil frequency has to be large in comparison with the Rabi frequency, and hence we define the adiabaticity pa- ~K2 rameter ωr ≡ (21) 2M Ω ε ≡ (24) and the Doppler frequency as ωr as the comparison of these two timescales. pK ν ≡ . (22) In the Bragg regime, the population of higher momen- D M tum states is suppressed and hence we find Ω  ωr, or The coupling to the same momentum has been removed ε  1, and the method of averaging can be applied.  2 2  Thus, the parameter ε describes the adiabaticity of the by multiplying g(p) with exp i(|Ωa| + |Ωb| )t/∆ . At this point we want to emphasize that we have de- diffraction process. fined our effective frequency Ω in such a way that it is the We now define g(p + ~nK) ≡ gn as components of a frequency of the probability amplitude, rather than the vector and arrive at Rabi frequency ΩRabi at which the population of a state   oscillates. To link our definition to the latter–maybe X iνωrt g˙ = iε H0 + e Hν  g , more familiar–definition, we make use of the relation ν=0 6 Ω Ω ≡ Rabi . which has the form of Eq. (1) and where the matrices are 2 defined as iν t Indeed, a π/2 pulse, i. e. ΩRabi t = π/2, is in our descrip- H D ( ν )n,n = ωr e− δn+1,n0 δ2n, ν tion Ω t = π/4. On the other side, when we face double 0 − iνDt  + e δn 1,n δ2(n 1),ν , diffraction, we are not dealing with a two-level system − 0 − 9 with the Kronecker delta δm,n. The coupling strength exponential as in Eq. (8), we find in Eq. (23) was Ω and it still is, but in order to have a      dimensionless expansion parameter we have introduced 1 0 0 1 γ(1)(t) = cos (Ωt) + i sin (Ωt) γ(1)(0) . Ω ≡ ε ωr, which is why the Hamilton matrix has the 0 1 1 0 dimensions of a frequency. (26) For the moment, we assume p = 0, i. e. νD = 0. In this case, Hν becomes time-independent, namely This result contains the well-known Rabi oscillations [43] between the states |0i and | Ki with the effective Rabi  ~ (H ) =ω δ δ + δ δ , (25) frequency Ω. With this result, we see how Bragg pulses ν n,n0 r n+1,n0 2n, ν n 1,n0 2(n 1),ν − − − can be used to create superpositions of different momen- and the differential equation is exactly of the form of tum states. A beam splitter, or π/2 pulse, is achieved for Eq. (1). However, this equation just describes the time Ω t = π/4, a mirror, or π pulse, for Ω t = π/2. evolution of g(~nK) ≡ gn for the resonant momenta. We discuss the role of a deviation from resonance in the context of double Bragg diffraction in Sec. IV C 2. 2. The quasi-Bragg regime

1. Rabi oscillations between momentum states These Rabi oscillations have been derived many times before, see e. g. [43–45], by adiabatically eliminating all We now apply the method of averaging described in momentum states but |0i and |~Ki. We see now how the Sec. II to find a slow solution γ with the matrices Hν method of averaging corresponds to the technique of adi- from Eq. (25). According to Eq. (7), the differential equa- abatic elimination. Of course this is an approximation, tion for the first approximation (m = 1) of the slowly but our method allows us to easily derive corrections to (1) (1) (1) T these Rabi oscillations, and thus to calculate populations oscillating terms is with γ = (γ0 , γ1 ) , where the superscript T denotes the transpose, of the levels that would have been eliminated if we had pursued the conventional approach. 0 1 We note that this solution is completely independent of γ˙ (1) = iεH γ(1) = iΩ γ(1) . 0 1 0 the adiabaticity parameter ε. Hence, the slowly evolving solution γ(1) is correct to lowest order of ε. We call the Taking more than two states into account is possible, but regime where this description is sufficient the deep Bragg since we chose ∆ω = ωr, just the two neighboring states regime. If we want to calculate corrections up to the are resonant. next higher order, which now leads into the quasi-Bragg Solving this differential equation by taking the matrix regime explored by Ref. [23], we can use Eq. (9) to get

  iνωrt (1) X e (1) g (t) = 1 + ε Hν  γ (t) νωr ν=0 6  ε i2ωrt ε i2ωrt   1 2 e cos (Ωt) i 2 e sin (Ωt) 0 ε i2ωrt     − 2 e− 0 0 0  1 0 0 1  (1) =  ε i2ωrt + cos (Ωt) + i sin (Ωt)  γ (0) .  0 0 0 − 2 e−  0 1 1 0  ε i2ωrt ε i2ωrt 0 i 2 e sin (Ωt) 2 e cos (Ωt) 1 (27)

To be able to calculate these correction terms and the order process deviates from the resonant kinetic energy, population of the states |−~Ki and |2~Ki, we have en- as shown in Fig. 2. For processes of this kind, energy larged the matrices to 4 × 4 and considered the vector conservation is not ensured, if such a dashed off-resonant (1) (1) (1) (1) (1) T transition is made, but due to the energy-time uncer- g = (g 1, g0 , g1 , g2 ) . This correction does not change the− Rabi oscillations between |0i and |~Ki but tainty this is possible. (1) (1) creates populations in the coefficients g 1 and g2 . The The solutions Eq. (27) are depicted in Fig. 3 for the population of these states is suppressed− by a factor ε and initial condition γ(1)(0) = (0, 1, 0, 0)T. The two bot- thus not relevant in the deep Bragg regime. tom lines show the small-amplitude off-resonant states In Eq. (27), the coefficients oscillate with a frequency n = −1, 2 magnified by a factor of 50. The off-resonant of 2ωr which is identical to the frequency that a second- state n = −1 behaves exactly like its adjacent resonant 10

1.0 We note that the population of the state |−~Ki now n = 0 n = 1 oscillates on a fast timescale with the frequency 2ωr. The 0.8 populations of the other states do not change signifi- cantly. When compared to the dotted numerical results,

2 0.6 small-amplitude deviations from the analytical solutions (1) n

g appear. To find these analytically, higher orders of ε 0.4 50, 50, would have to be taken into account. × n = 1 n×= 2 0.2 − 1.0 0.0 n = 0 n = 1 0 π/4 π/2 3π/4 π 0.8 Ω t

2 0.6 50, (1) (1) 2 n × FIG. 3. Populations g of momentum states for single g n = 1 n 0.4 − Bragg diffraction in the| quasi-Bragg| regime. The initial con- 50, × γ(1) T n = 2 dition (0) = (0, 1, 0, 0) corresponds to the dressed-state 0.2 formulation, and the adiabaticity parameter is ε = 0.1. The two bottom lines are magnified by a factor of 50 and show the 0.0 small populations in the off-resonant states n = 1, 2. 0 π/4 π/2 3π/4 π − Ω t state with n = 0, but with a suppressed amplitude. The (1) 2 FIG. 4. Populations gn of momentum states for single same is true for the other two states. Since the fast oscil- Bragg diffraction in the| quasi-Bragg| regime. In contrast to lations of the off-resonant terms are contained in a phase Fig. 3 now the initial condition is g(1)(0) = (0, 1, 0, 0)T, which factor exp [±i2ωrt], we do not see them in the corre- corresponds to the bare-state formulation. The adiabaticity sponding population, given by the modulus square. In parameter is again ε = 0.1. The two bottom lines are mag- Sec. IV we show that in double Bragg diffraction effects nified by a factor of 50 and show the small populations in the off-resonant states n = 1, 2. The dotted lines represent from these fast oscillations do occur even for the resonant − states. numerical results. For Fig. 3, we have chosen an initial condition for γ(1)(0) and are thus in a dressed-state formulation in- For square-shaped Bragg pulses, for which the treat- stead of choosing g(1)(0) and defining γ(1)(0) according ment based on the method of averaging is very well- to Eq. (10) in the bare-state formulation. Therefore, suited, the bare-state formulation is easier to interpret γ(1)(0) cannot be interpreted as the initial momentum since the initial condition corresponds to the physical sit- distribution. For this reason, the initial conditions for uation. Nevertheless, one has to keep in mind that most (1) 2 |gn | in Fig. 3 do not coincide with the intuitive ones. applications involve adiabatically turned-on light waves. We discuss in Sec. III D the bare-state formulation. In this case, the adiabaticity parameter ε becomes time- The discussion of this section predicts features of the dependent, as discussed in Appendix C. In Eq. (10) this diffraction process when leaving the deep Bragg regime translates into ε(0) = 0, and thus the dressed-state cor- and serves as an application of the method of averaging. responds to the bare-state formulation. However, for the case of double Bragg diffraction, this method is necessary. We focus on this point in Sec. IV. IV. DOUBLE BRAGG DIFFRACTION

D. Dressed versus bare states To introduce symmetric Bragg pulses analogously to the symmetric Raman pulses in the double-diffraction In the previous section, we have discussed the dressed- scheme by Ref. [37], we now extend our model to an in- state formulation of single Bragg diffraction in the quasi- teraction with four light waves. The method of averaging Bragg regime. We have used Eq. (27) and displayed outlined in Sec. II and applied in Sec. III is needed to find the time evolution of the dressed state, i. e. γ(1)(0) = a theoretical description of this process, since the elim- (0, 1, 0, 0)T, in Fig. 3. For square-shaped light pulses, ination of nonresonant momentum states is much more the dressed state might seem unphysical. So in Fig. 4 subtle. we depict the dynamics of a bare state with the initial condition g(1)(0) = (0, 1, 0, 0)T. This initial condition translates with Eq. (10) into a sophisticated superposi- A. Model tion of initial states γ(1)(0). In Fig. 4 we just display terms where the amplitudes scale up to the order of ε to In contrast to single Bragg diffraction, where two be consistent with the first-order treatment. counterpropagating light waves interact with a two-level 11 atom, we consider the following problem: One atom in- mz = 1 mz = 0 |e+i teracts with two pairs of light fields. Each pair induces a mz = −1 Bragg diffraction process, but in opposite directions. A |e−i possible set-up is shown in Fig. 5. ω+ λ Ω− Ω+ /4 plate ω−

ωa, σ1 pair 1 ωb, σ1 atom |gi mz = 1 mz = 0 ωb, σ2 pair 2 ωa, σ2 mz = −1 z mirror FIG. 6. Transitions between magnetic sublevels due to cir- cular polarizations. Laser pair 1 with polarization σ drives − the transitions between g and e , which have the energy FIG. 5. Schematic set-up of double Bragg diffraction. Two | i | −i difference ~ω , with a Rabi frequency Ω ; laser pair 2 with pairs of counterpropagating light waves (pair 1 and pair 2) − − polarization σ causes the transitions between g and e , induce a Bragg scattering process in opposite directions. In + + which have the energy difference ω , with a Rabi| i frequency| i order to distinguish both pairs, they have orthogonal polar- ~ + Ω . The kinetic energy is neglected in this figure and m izations σ and σ . This orthogonality can be achieved by a + z 1 2 denotes the magnetic . retroreflecting geometry using a mirror and a λ/4 wave plate.

To suppress stimulated emission induced by pair 1 if |e±i the atom was excited by pair 2, the polarizations of both pairs are chosen to be orthogonal, i. e. σi ·σj = δi,j. Oth- ∆ erwise, spurious scattering processes might occur: The atom in Fig. 5 could, for example, absorb a ‘red’ photon from pair 1 and thus gain the momentum ~ka. If now an ω± emission of a ‘blue’ photon from pair 2 into the same di- rection occurs, the atom loses the momentum ~kb, which |gi yields the total momentum transfer ~(ka − kb). In addi- tion, the desired processes with momentum transfers of ±~(ka + kb) take place. Taking into account all possible p processes, one momentum state is therefore coupled to 2~K ~K 0 ~K 2~K eight different momenta. |− i|− i| i | i | i In the main part of this article, we perform the calcula- tion for circularly polarized light fields, i. e., σ = σ and FIG. 7. Momentum transfer and resonances in double Bragg 1 + diffraction. The Bragg scattering processes induced by each σ2 = σ . However, other choices of orthogonal polariza- − pair of lasers correspond to the one of Fig. 2, but are of op- tions are possible as well. In Appendix B we extend our posite directions. The frequencies of the excited states are model to orthogonal linear polarizations with a magnetic ω+ and ω . The dashed transitions are off-resonant and get − field in an arbitrary direction causing a Zeeman splitting. suppressed.

We assume that the atom can be excited to two dif- ferent states |e+i and |e i, corresponding to each polar- ization. Depending on the− specific species of atoms used duces independently a Bragg scattering process; pair 1 in the experiment, one can identify magnetic sub-levels to the left and pair 2 to the right. The dashed lines are as in Fig. 6. Although even atoms with ground state the off-resonant processes of each pair. We discuss their magnetic quantum numbers mz = ±2 are possible, it is meaning and influence in Sec. IV C 1. sufficient to consider just these three states if a magnetic field is applied such that the transitions to the mz = ±2 levels are far detuned. For now, we use the configuration of Fig. 6. B. Three-term recurrence relation For this specific configuration, it is not necessary to make any restrictions on the Zeeman splitting ω+ − ω . However, for different polarizations (e. g., linear polar-− We now derive analogously to Sec. III B a three-term izations as discussed in Appendix B) it is necessary that recurrence relation for the double Bragg diffraction pro- ∼ |ω+ − ω |  ∆. So we assume ω+ = ω ≡ ωeg. cess of the model described in the preceding section. The We show− the momentum transfer for− double Bragg interaction of a two-level atom with four light waves diffraction in Fig. 7. Here, each pair of light fields in- shown in Fig. 5 yields in rotating wave approximation 12

[47] the Hamiltonian state, we find a system of coupled differential equations for g(p) and e (p). The latter ones can be eliminated adi- 2 ± ˆ pˆ abatically in the case of large detuning ∆ ≡ ωeg − ωa,b in H = + ~ωeg (|e+i he+| + |e i he |) 2M − − comparison to the Rabi frequencies Ω , i. e. |Ω /∆|  1, n h i analogous to the approach of Sec. III± B. In this± way, we + Ω ei(kazˆ ωat) + ei( kbzˆ ωbt) |e i hg| ~ + − − − + find the system of differential equations h i o i( kazˆ ωat) i(kbzˆ ωbt) + Ω e − − + e − |e i hg| + h.c. , i pK t − − ig˙(p) = − g(p + ~K) e− m (28)  2 2  |Ω+| i(∆ω+ω )t |Ω | i(∆ω ω )t × e− r + − e − r where we have introduced the Rabi frequencies Ω ≡ ∆ ∆ ± ˆ ˆ pK −|E| he | d · σ |gi /~ and used he | d · σ |gi = 0. For i m t ± ± ± ∓ − g(p − ~K) e the sake of simplicity we assume all the light fields to have  2 2  |Ω+| i(∆ω ω )t |Ω | i(∆ω+ω )t the same amplitude |E|, even though this assumption is × e − r + − e− r not necessary for the further treatment. In particular, ∆ ∆ different amplitudes are assumed in Appendix B. The Hamiltonian Hˆ given by Eq. (28) describes exactly for the probability amplitude of the state |gi. As in the situation shown in Fig. 6 and Fig. 7. Acting with it Sec. III B, we have transformed again into the interac- tion picture. Here, ∆ω ≡ ωb − ωa denotes the difference on the state 2 of the light frequencies and ωr ≡ ~K /(2M) the recoil Z frequency, respectively, where the effective wave vector is |ψi = dp [g(p) |p, gi + e+(p) |e+, pi + e (p) |e , pi] − − again defined as K ≡ ka + kb. With the notation gn ≡ g(p + ~nK) and defining the and applying the Schr¨odingerEq. (16) to equate the co- effective transition frequency Ω ≡ |Ω |2/∆, the time- efficients of |g, pi and |e , pi to the time derivative of this dependent three-term recurrence relation± reads ±

pK h i pK h i i t i[∆ω+(2n+1)ωr]t i[∆ω (2n+1)ωr]t i t i[∆ω+(2n 1)ωr]t i[∆ω (2n 1)ωr]t ig˙n = − Ω e− m gn+1 e− + e − − Ω e m gn 1 e − + e− − − . − (29)

When we compare this expression to the conventional C. First-order diffraction Bragg diffraction case, Eq. (23), we identify two terms instead of one within each of the brackets. The first In this section, we concentrate on first-order diffraction term comes from the light fields of pair 1, the second where ∆ω = ωr. For this case, Eq. (29) reduces to one from pair 2. Since each term oscillates at a different pK h i i t i2(n+1)ωrt i2nωrt frequency, there are always two couplings between neigh- ig˙n = − Ω e− m gn+1 e− + e− boring momentum levels. Hence, we cannot perform the pK h i i t i2nωrt i2(n 1)ωrt adiabatic elimination of off-resonant momentum levels in − Ω e m gn 1 e + e − . − the familiar way, since some levels are resonantly and off-resonantly coupled at the same time. For example, With the notation from Sec. III B, the matrices have the form Fig. 7 shows that the state |−~Ki is coupled to |0i reso- nantly by pair 1 with a solid green line, and off-resonantly H  iνDt ( ν )n,n =ωr e δn 1,n0 (δ2n,ν + δ2(n 1),ν ) 0 − − by pair 2 with a dashed brown line. Nevertheless, the iνDt  + e− δn+1,n (δ2(n+1), ν + δ2n, ν ) , method of averaging described in Sec. II can be applied 0 − − and automatically deals with this problem. (30) where we have recalled the definition of the Doppler fre- quency νD from Eq. (22). Note that we have not yet chosen p = 0 as in Eq. (25). In Sec. IV C 1 we consider p = 0, whereas in Sec. IV C 2 we allow for p =6 0.

1. Beam splitters and mirrors

To see if we can realize beam splitters and mirrors based on double Bragg diffraction, we turn to the deep 13

Bragg regime, where the slow solution γ(1) is sufficient. 1.0 n = 0 We consider first the case of p = 0, which also a) 0.8 n = 1 means that νD = 0, and the matrices Hν become time- ± independent. According to Eq. (7) we have to solve the

2 0.6 differential equation (1) n γ 0 1 0 0.4 (1) (1) (1) γ˙ = iεH0γ = iΩ 1 0 1 γ , 0 1 0 0.2

(1) (1) (1) (1) T 0.0 π with γ = (γ 1 , γ0 , γ1 ) . 0 π/2 3π/2 2π According to− Eq. (8), the solution of this system is √ given by the matrix exponential which can be carried 2Ω t out by any computer algebra system. In this case, we 1.0 n = 1 find b) − 0.8 n = 0      1 0 −1 √ 1 0 1 n = +1 (1) 1 1   2 0.6 γ (t) =   0 0 0  + cos 2Ωt 0 2 0 2 2 (1) −1 0 1 1 0 1 n γ   0.4 √ 0 1 0 i   (1) +√ sin 2Ωt 1 0 1 γ (0) . (31) 0.2 2 0 1 0 0.0 which describes quasi-Rabi oscillations of a three-level 0 π/2 π 3π/2 2π (1) 2 system. The populations |γn | of the momentum states √2Ω t are shown in Fig. 8 for the initial conditions (0, 1, 0)T and T (1, 0, 0) . Indeed, π/2 and π pulses√ can be performed FIG. 8. Rabi oscillations to realize a beam splitter (a) and and the effective Rabi frequency is 2Ω. Thus, beam mirror (b) in the deep Bragg regime. For the adiabaticity splitters and mirrors can be realized. (1) 2 parameter ε 1 the quantity γn corresponds to the pop-  | | This result is not obvious, since these oscillations oc- ulation of a momentum state n~K . For the demonstration cur in an effective three-level system. In contrast to sin- of a beam splitter (a) the initial| conditioni is (0, 1, 0)T. Start- gle Bragg diffraction, where a π/2 pulse was generat- ing in 0 , a superposition ( ~K + ~K )/√2 is generated ing an equal superposition of the initial state and the for √2Ω| it = π/2. For the demonstration|− i | i of a mirror (b) the T other state, we now use the term “π/2 pulse” when an initial condition is (1, 0, 0) . Starting in ~K , the whole pop- | i ulation is transferred to ~K at the time √2Ω t = π. equal superposition of the two states |±~Ki is created |− i and the initial state is completely depopulated. In the same spirit, a “π pulse” now exchanges the population of the two states |−~Ki and |~Ki. plotted in Fig. 9 and compared to a numerical simula- The definition of the matrices Hν in Eq. (30) shows tion. For the choice of ε = 0.1, we see excellent agree- that there are adjustments to the momentum states |0i ment. For the plot, we used the bare state initial condi- (1) T and |±~Ki if we leave the deep Bragg regime and take tion g (0) = (0, 1, 0) from Eq. (10). rapidly oscillating corrections of order ε into account. In These small-amplitude fast oscillations at the order of addition to that, fast oscillations of the momentum states ε for dressed states are a new feature of double Bragg |±2~Ki with small amplitudes can be found analogously diffraction, namely that we have simultaneously a reso- to the single-diffraction case discussed in Sec. III C. But nant and off-resonant coupling of a momentum state. since we are only interested in the change of the dynamics of the states |0i and |±~Ki we just look at the innermost 3 × 3 matrix, even though we have used a larger matrix 2. Small deviations from resonant momenta: Velocity to calculate it. In this case Eq. (9) reads selectivity   0 ei2ωrt 0 (1) (1) ε i2ω t i2ω t (1) g = γ + − e− r 0 − e− r  γ . So far, we have just considered first-order double Bragg 2 0 ei2ωrt 0 diffraction with corrections of order ε for the case p = 0. In this case, the momenta |0i and |±~Ki are resonant, The influence of off-resonant processes is taken into ac- but all other states |n~Ki are suppressed. But if we count by a fast oscillation on top of the slow Rabi os- now allow p =6 0, we can still find approximate analytic cillation. The fast frequency is again 2ωr, which corre- solutions which are a double Bragg generalization of the sponds to the energy that the off-resonant process de- single Bragg case as discussed in Ref. [33]. We dedicate (1) 2 viates from the resonance. The populations |gn | are the current section to investigate this feature in more 14

1.0 factors are of importance. n = 0 In Sec. II, we have integrated in Eq. (6) by neglect- n = 1 0.8 ± ing any time-dependence of Hν . This approximation is only true if these matrices vary slower than the exponents 2 0.6 exp [iνωrt]. This requirement is fulfilled for |νD|  ωr or (1) n

g |p|  K/2. Our solution is only a good approxima- 0.4 ~ tion if we fulfill this condition, so we just consider small 0.2 deviations from the resonant momenta. In this case γ(1) can be found easily. If we multiply 0.0 γ 1 by exp [∓iνDt], we find the new system 0 π/2 π 3π/2 2π ± √2Ω t   νD Ω 0 γ˙ (1) = i  Ω 0 Ω  γ(1) (32) FIG. 9. Beam splitter in the quasi-Bragg regime for the adi- 0 Ω −νD abaticity parameter ε = 0.1 represented by the population (1) 2 gn of the momentum state n~K . The initial condition | | | i of differential equations for these coefficients. is g(1)(0) = (0, 1, 0)T. The dashed lines are the result of a numerical simulation, the solid lines are the approximate an- This equation shows what happens if the momentum alytical solutions. In this regime, fast oscillations with small deviates from a multiple of ~K: Two elements ±νD show amplitude modify the Rabi oscillations from Fig. 8. up on the main diagonal of the matrix, which differ from the element in the center of the matrix. This is why they act as a detuning or AC Stark shift [51] on the Rabi detail. oscillations. For increasing deviations from the resonant For p =6 0, the matrices Hν are, according to Eq. (30), momenta the interaction is therefore suppressed. This proportional to exp [±iνDt] and thus time-dependent. effect is well-known and called velocity selectivity [30, 33]. Since the method of averaging delicately plays with the This feature becomes more obvious when we look at the combination of fast and slowly oscillating terms, these solution

  2 2  2 2 2    Ω −νDΩ −Ω νD+Ω νDΩΩ νD Ω 0 (1) 1 2 cos (Ωefft) 2 i sin (Ωefft) (1) γ (t) =  2 −νDΩ νD νDΩ+ 2  νDΩ 2Ω −νDΩ +  Ω 0 Ω  γ (0) Ωeff 2 2 Ωeff 2 2 2 Ωeff −Ω νDΩΩ Ω −νDΩ νD+Ω 0 Ω −νD (33)

of Eq. (32), where we have defined the effective Rabi In the bottom part of the figure the difference between p 2 2 this analytical solution and a numerical simulation is frequency Ωeff ≡ 2Ω + νD. The matrix exponential defined by Eq. (8) was calculated with the help of a shown. As expected, they coincide only in the imme- computer algebra system. Indeed, Eq. (33) reduces to diate neighborhood of p = 0 but deviate for larger values Eq. (31) for p = 0, i. e. if the Doppler frequency νD van- of p. ishes. The effective Rabi frequency clearly demonstrates Even though the exact frequencies and amplitudes are that the Doppler frequency νD acts as a detuning. not very well approximated for increasing momenta, the Figure 10 shows in comparison with Fig. 8 the effect of width of the resonance is similar and thus the approxi- detuning and suppression for a momentum p = 0.01~K, mation made above surprisingly good. which corresponds to νD = 0.02ωr. We note that the Since in experiments the atoms do have a distribution frequency of the oscillation as well as the amplitude are over many momenta, one has to average the Rabi oscil- changed. lations shown in Fig. 11 over the momentum and weigh Figure 11 shows a three-dimensional plot of the Rabi them with their initial momentum distribution. In our oscillations for different momenta. The analytical solu- simulation we use an initial Gaussian distribution tion, i. e. Eq. (33), for the diffracted state populations (1) |γ |2 is shown in the top part of the figure. As already  2  1 (1) (p/~K) ± γ (0, p) = N exp − , discussed above, for large deviations from the resonance 0 4σ2 p = 0 the interaction is suppressed, and almost no popu- p lation occurs, which very clearly demonstrates the veloc- ity selectivity. with a momentum width of σp in units of ~K, and a 15

1.0 n = 0 n = 1 0.8 ±

2 0.6 a) 0.4 2 (1) n

γ 1 0.4 π (1) ± 0.2 γ

0.2 0 π/2 0.1 − 0 0.0 0.1 0 √2Ω t 0 π/2 π 3π/2 2π p/(~K) √2Ω t

(1) 2 FIG. 10. Rabi oscillations of the populations γn of the

| | 2 momentum states (n + 0.01)~K in the deep Bragg regime b)

| i 1 0.2

(ε = 0.01) for the off-resonant momentum p = 0.01 K. The num ~ ± g

Rabi oscillations are detuned and suppressed, which leads to π velocity selectivity. The initial condition is (0, 1, 0)T and the − 2

time is in units of the resonant frequency √2Ω. 1 (1) ± 0 π/2 γ

0.1 − 0 0.1 0 √2Ω t normalization constant p/(~K)

1 √  2 − FIG. 11. Velocity selectivity of double Bragg diffractionex- N ≡ 2π σp K , ~ (1) 2 pressed by the Rabi oscillation of the population γ 1 in | ± | and calculate numerically the integrated population their dependence on the initial momentum (a). For p = 0, i. e. for the resonant momenta ~K, we get full Rabi oscilla- tions, but as we leave this resonance± this process is suppressed ~K/2 Z 2 and detuned. The difference between analytical and numeri- (1) N 1 = dp γ 1 (t, p) , cal solutions (b) shows a large deviation between both solu- ± ± tions. This deviation is mainly due to different frequencies, ~K/2 − not to different amplitudes. The width of the resonance is of atoms with momenta around ±~K. In time-of-flight comparable in both solutions. Here we have chosen ε = 0.01. experiments these can be identified by measuring the clouds drifting away from the center. The results of this numerical integration for different single Bragg diffraction has its origin in the three-level momentum widths shown in Fig. 12 display a damping behavior of double diffraction. An extensive numerical of the Rabi oscillation. This feature is only an artefact of discussion of the importance of the momentum width for the interaction times; for very long times the oscillation single Bragg diffraction for different diffraction orders is revives. Since in atom interferometry the interest lies on given in Ref. [33]. π/2 or π pulses, the interaction time is limited and re- Of course, the velocity selectivity decreases with in- vivals are not observable. The damping can be explained creasing light field intensity, i. e. in the end with increas- very easily: As we already discussed, the Rabi frequency ing adiabaticity parameter ε. This feature can be seen depends on the initial momentum and thus gets washed from Eq. (33), where the states are suppressed by the p 2 2 p 2 2 out when averaged over a momentum distribution. This fraction Ω/ 2Ω + νD = ε/ 2ε + [2p/(~K)] . For in- dephasing effect is an intrinsic problem of the diffrac- creasing ε, the significance of p =6 0 decreases. tion process, which can be improved by using narrower On the other side, an increasing ε leads to the quasi- momentum distributions, generated, e. g., by techniques Bragg regime. So higher-order excitations become more such as delta-kick cooling [3]. and more significant. √ The frequency of the effective Rabi oscillation is now 2Ω instead of Ω. This property might lead to the con- clusion that the interaction times in double diffraction D. Second-order double Bragg diffraction would be shorter. But to compare the velocity selectivity of single and double Bragg diffraction, it is important to The formalism introduced in our article makes it pos- look not at the interaction times, but at the arguments of sible to also describe second-order double Bragg diffrac- the trigonometric functions in Eq. (33). From this point tion. For that, we choose according to Eq. (15) the res- of view,√ a beam splitter and a mirror take twice as long, onance condition ∆ω = 2ωr. The resulting second-order e. g. at 2Ω t = π/2 instead of Ω t = π/4 for a π/2 pulse. process is depicted in Fig. 13. Hence, the higher velocity selectivity in comparison with The first scattering process, a two-photon transition, is 16

3 be written as σp = 10− 2 h i 0.4 σp = 10− iνDt i(2n+3)ωrt i(2n 1)ωrt ig˙n = − Ω e− gn+1 e− + e− −

iν t h i(2n+1)ω t i(2n 3)ω ti

1 D r r − Ω e gn 1 e + e − . ± − N 0.2

1. Emergence of the asymmetric AC Stark effect

0.0 For the sake of simplicity we look at p = 0, i. e. ν = 0, 0 π/2 π 3π/2 D which leads us to the matrices √2Ω t (Hν ) = ωr [δn 1,n (δ2n+1,ν + δ2n 3,ν ) n,n0 − 0 − FIG. 12. Damped Rabi oscillations in the numerically in- +δn+1,n (δ2n+3, ν + δ2n 1, ν )] . (34) 0 − − − tegrated population N 1 when the adiabaticity parameter is ± ε = 0.01. For a broad momentum distribution, the oscillations Hence, Hν is zero for even ν; in particular, H0 = 0. get damped out due to the average over different frequencies Since the differential equation for the first slow ap- and amplitudes. This effect is less pronounced for smaller proximation reads γ˙ (1) = 0 we find the trivial solution values of the momentum width σp. γ(1)(t) = γ(1)(0). This result is not surprising, since in this approximation we can only describe first-order pro- cesses and they are all suppressed, in agreement with |e±i Fig. 13. Next, we turn to the second-order approximation; i. e., we go to the next higher order in ε, that is m = 2. Ac- cording to Sec. II, the solution given by Eq. (13) reads   (2) 2 X H ν Hν (2) γ (t) = exp iε − t γ (0). νωr ν=0 |gi 6

With the specific form of Hν given by Eq. (34) we find p  76  − 105 0 −1 0 0 28 2 |−2~Ki|−~Ki|0i |~Ki |2~Ki  0 15 0 3 0  X H ν Hν  4  − = ωr  −1 0 − 3 0 −1  , (35) νωr  2 28  FIG. 13. Two processes in second-order double Bragg diffrac- ν=0  0 3 0 15 0  6 76 tion (∆ω = 2ωr). A four-photon transition in each direction 0 0 −1 0 − 105 occurs. For an initial population of the momentum states where we have focused on the inner 5×5 matrix and used 0 and 2~K , the states ~K are just virtually occu- | i |± i |± i (2) (2) (2) (2) (2) (2) T pied. Different initial conditions are sketched in Fig. 14. The γ ≡ (γ 2 , γ 1 , γ0 , γ1 , γ2 ) . − − dashed lines show off-resonant processes. We can already recognize the structure of the resulting oscillations. Similar to the case of p =6 0 in Eq. (32) for first-order diffraction in the previous section, we now see entries on the main diagonal of the matrix. As already off-resonant by ωr. But adding a subsequent two-photon analyzed in Sec. IV C 2, values on the main diagonal that transition, one reaches resonantly the states |±2~Ki. For differ from the central element account for an AC Stark certain parameters, the intermediate state is just occu- shift. For a second-order process, this effect is not sur- pied virtually. This is a process similar to the adiabatic prising. But in contrast to the case in the context of elimination of the excited state described in Sec. III B. velocity selectivity, the deviations do have the same sign. Now, the transition frequency Ω has to be considerably We come back to this point after we have discussed the smaller than the detuning, i. e. the deviation from reso- solutions nance, which is ωr.  (2)   (2)  γ 2 (t) γ 2 (0) On the other hand, in the deep Bragg regime, where −(2) −(2) γ (t) = M 2(t) · γ (0) (36) off-resonant processes are strongly suppressed, nothing  0  ±  0  γ(2)(t) γ(2)(0) would happen. Therefore, we have to leave the deep +2 +2 Bragg regime in order to allow this process, as we shall and see now. (2) ! (2) ! γ 1 (t) γ 1 (0) When we use the resonance condition ∆ω = 2ωr for −(2) = M 1(t) · −(2) (37) ± the second-order Bragg scattering process, Eq. (29) can γ+1 (t) γ+1 (0) 17 for the slowly evolving part in second approximation. structure of the differential equation, we boldfaced those Here we have separated the oscillations between the entries in Eq. (35) that are responsible for the transitions states |−2~Ki , |0i , and |2~Ki from the ones between between |−~Ki and |~Ki. The dynamics of the states is |−~Ki and |~Ki. determined by the matrices To show that this separation follows directly from the

 1 0 −1  1 0 1 −16 105 −16 76 2 108 2 i ε ωrt 1 i ε ωrt cos (Ω 2t) i sin (Ω 2t) M 2 = e− 108  0 0 0  + e− 105  ± 0 2 0 − √ ±  105 32 105  ± 2 −1 0 1 2 1 0 1 23074 −16 105 −16

and 1.0      n = 0 28 2 i ε ωrt 1 0 0 1 n = 2 M 1= e− 15 cos Ω 1t + i sin Ω 1t , 0.8 ± ± ± 0 1 ± 1 0

(38) 2 0.6

(2) n where the Rabi oscillation between the states |± Ki has γ

~ 2 0.4 the frequency Ω 1 ≡ (2/3) ε ωr = (2/3) εΩ, determined ± by one of the boldfaced numbers in Eq. (35), and the 0.2 quasi-Rabi oscillation√ between the states√|±2~Ki has the ≡ 2 frequency Ω 2 ( 23074/105)ε ωr = ( 23074/105)εΩ. 0.0 π Analogously± to the result of the adiabatic elimination 0 π/2 3π/2 2π of the excited state Eq. (20), where we found the Rabi Ω 2 t ± frequency Ω = |Ω |2/∆, the frequency of this solution ± 2 is now proportional to Ω /ωr = Ωε. The same structure FIG. 15. Rabi oscillations in second-order double Bragg (2) 2 makes the connection to the adiabatic approximation ob- diffraction where γn corresponds to the population of the | | vious. The adiabaticity parameter ε now decreases the momentum state n~K for ε 1 .The initial condition frequency and makes the second-order oscillation slower (0, 0, 1, 0, 0)T was chosen| i to focus on the diffraction process in comparison to a first-order process. 0 2~K where the states ~K are not involved. The |oscillationsi ↔ |± arei slightly suppressed|± duei to an asymmetric AC |e±i Stark shift. The dashed lines show the numerical solution for ε = 0.01.

and no involvement of other states. The resonant second- order process that connects these two states is depicted in Fig. 14. |gi However, we are more interested in the oscillations from Eq. (36). They correspond to the processes shown in Fig. 13. Indeed, these oscillations are the quasi-Rabi p oscillations of a three-level system between the relevant momentum states |±2 Ki and |0i. The oscillations are |−2~Ki|−~Ki|0i |~Ki |2~Ki ~ suppressed by a factor smaller than unity in front of the sine-function and plotted in Fig. 15. The states cannot FIG. 14. Scattering process between the states ~K ac- be completely depopulated with a π/2 or π pulse, i. e. cording to Eq. (37). If the atoms are initially|± in onei of these states, a resonant Rabi oscillation between them takes perfect beam splitters and mirrors are not possible. This place in second-order Bragg diffraction. This phenomenon feature is very different from single second-order Bragg can be explained by the resonant four-photon process shown scattering. Qualitatively, this difference is a consequence above. The transitions for other initial conditions are shown of the form of the matrix Eq. (35) where we have already in Fig. 13. highlighted the origin of this AC Stark shift due to the ‘asymmetry’ of the coupling to off-resonant states. We first discuss the oscillation between the states We also want to emphasize that the two-photon light |±~Ki. According to Eq. (38) they do have the same shift occurring in Raman transitions [52, 53] is related form as the ones in Eq. (26), which are just Rabi oscilla- to this asymmetric AC Stark shift, but there are sub- tions between these two states. There is no suppression tle differences. First of all, the two-photon light shift is 18 a frequency shift due to off-resonant states occurring in study the effect of a slightly violated resonance condition; a double Raman set-up, where an asymmetry is intro- i. e., we choose ∆ω = (2 + δ)ωr with δ  1. duced by giving the atom an initial velocity, in contrast When we include the phase factor exp [i2δωrt] in the to the double Raman diffraction of Ref. [37]. However, states γ 2, the matrix for the new system of differential frequency shifts due to off-resonant transitions occur even equations± reads in single Bragg and Raman diffraction. The asymmetric  76 δ  AC Stark shift in second-order double Bragg diffraction − 105 + 2 ε2 0 −1 0 0 28 2 is a consequence of the special form of Eq. (35) and not  0 15 0 3 0  XH ν Hν  4  − just due to higher off-resonant states. =ωr −1 0 − 3 0 1 . νωr  2 28  Since the possibility of higher-order diffraction is one of ν=0 0 0 0 6  3 15  the advantages of Bragg diffraction over Raman diffrac- 76 δ 0 0 −1 0 − 105 + 2 ε2 tion, it seems as if this advantage is lost in double diffrac- tion, since no perfect beam splitters and mirrors in higher With the choice δ ≡ −(32/105) ε2 this matrix reduces to orders can be realized. However, we show in the following  4  section that this effect can be compensated. In addition − 3 0 −1 0 0 to that, we believe that the asymmetric AC Stark shift 28 2  0 15 0 3 0  X H ν Hν  4  is a small effect in comparison to the damping and the − = ωr −1 0 − 3 0 −1  , imperfections due to the width of the momentum distri- νωr  2 28  ν=0  0 3 0 15 0  bution discussed above. 6 4 0 0 −1 0 − 3 where now the elements on the main diagonal that be- 2. Compensation for the asymmetric AC Stark shift long to the states |±2~Ki do not differ from the central element, and thus no asymmetric AC Stark shift occurs. Fortunately, the asymmetric AC Stark shift can be The Rabi oscillation Eq. (37) between the states |±~Ki compensated for. To understand this claim analytically, does not change at all, but the matrix of Eq. (36) reads we recall from Sec. IV C 2 the case of momenta p =6 0 and now

  1 0 −1 √  1 0 1 √  0 1 0 i 4 Ωεt 1 cos 2Ωεt i sin 2Ωεt M 2 = e− 3   0 0 0  + 0 2 0 − √ 1 0 1 ± 2 −1 0 1 2 1 0 1 2 0 1 0

and a resonant effective√ Rabi oscillation as in Eq. (31) insert it into Eq. (29) which yields via the relation with a frequency 2Ωε occurs. Hence, the frequency of h i i(2n+3)ωrt i(2n+1)ωrt the first-order diffraction is just multiplied by ε. ig˙n = − Ω gn+1 e− + e− In general, the effect of the asymmetric AC Stark shift h i i(2n+1)ωrt i(2n 1)ωrt − Ω gn 1 e + e − leading to suppressed Rabi oscillations in second-order − double diffraction can be eliminated by just changing the frequencies of the lasers. This result is numerically veri- the matrices fied in Fig. 16. (Hν ) = ωr [δn 1,n (δ2n+1,ν + δ2n 1,ν ) n,n0 − 0 − +δn+1,n (δ2n+3, ν + δ2n+1, ν )] . 0 − −

E. Quasi-resonances Since H0 ≡ 0 we directly turn to the second approxima- tion and find In second-order approximation a feature for the first-  4  − 5 0 −1 0 order Bragg condition occurs which we call quasi- H H 4 X ν ν  0 3 0 −1 resonances. In order to demonstrate this phenomenon − = ωr  4  , νωr −1 0 0  ν=0 3 we again choose ∆ω = ωr, but now concentrate on mo- 6 4 0 −1 0 − 5 menta close to half-integers of ~K, e. g. of ±~K/2 and so on. As Fig. 17 shows, a second-order process now (2) (2) (2) (2) (2) T for γ = (γ 3 , γ 1 , γ 1 , γ 3 ) , where we have defined becomes resonant. − 2 − 2 2 2 (2) (2) (2) For this reason we set p = K/2, i. e., ν = ω , and γ (p + n K) = γ ( K/2 + n K) ≡ γ 1 . ~ D r ~ ~ ~ n+ 2 19

1.0 1.0 n = 1 − 2 3 0.8 0.8 n = 2 2 0.6 2 0.6

(2) (2) 0 n γ γ

0.4 0.4

0.2 0.2

0.0 0.0 0 π/2 π 3π/2 2π 0 π/2 π 3π/2 2π

√2Ωε t Ω 1 t ± 2

FIG. 16. Rabi oscillations in second-order double Bragg FIG. 18. Rabi oscillations between the quasi-resonances diffraction after elimination of the asymmetric AC Stark ~K/2 3~K/2 for ∆ω = ωr (solid lines). The oscilla- shift. The elimination was achieved by choosing ∆ω = tions|− arei suppressed↔ | i by approximately one half. The dashed 2 (2 32/105 ε )ωr (solid line). The initial condition was again lines show a numerical solution of the Rabi equations for − T (0, 0, 1, 0, 0) . The oscillations are not detuned anymore. ε = 0.01 and confirm our analytical result. The dashed line which represents the numerical solution for ε = 0.01 and the same choice of ∆ω is in agreement with the approximate analytical result. elements corresponding to the latter are boldfaced in the matrices above. These oscillations are shown in Fig. 18 |e±i and coincide with the dotted numerical solution as well. In fact, these quasi-resonances are no special feature of double Bragg diffraction, as one sees easily from Fig. 17. We included them for completeness and want to empha- size that this processes are in double diffraction asymmet- ric, too. So, in general the symmetry of double diffraction is lost if there are non-vanishing initial momenta. On the |gi other side, the width of these resonances in momentum space is much narrower and the oscillation occurs on a much larger timescale than the first-order resonance, as p one sees from the discussion above. |−2~Ki|−~Ki|0i |~Ki |2~Ki

FIG. 17. Quasi-resonances in first-order double Bragg diffrac- V. CONCLUSIONS tion for ∆ω = ωr. Even though we consider first-order diffrac- tion, a second-order process between momenta ~K/2 and |± i In this paper we have introduced and analyzed double ~3K/2 is resonant. |± i Bragg diffraction in the context of light-pulse atom inter- ferometry, which has a number of appealing properties Since there are deviations on the main diagonal, the for precision interferometry, particularly in micrograv- matrix exponential yields again detuned oscillations. The ity environments, as summarized in the introduction. In solution reads the first place, we have derived the basic equations gov- erning the dynamics of momentum eigenstates as driven 4 2 (2) i ε ωr (2) γ (t) = e− 15 M 1 (t) γ (0) by two-photon processes that result from eliminating an 2 ± excited state largely detuned from single-photon transi- with tions. This system of coupled linear ordinary differen- 16 0 15 0  tial equations with time-dependent coefficients can be i sin(Ω 1 t) 2 0 −16 0 15 solved numerically, but it is, of course, convenient to M 1 =cos(Ω 1 t) 1 − √ ±   ± 2 ± 2 481 15 0 −16 0  have simple analytical solutions, from which deeper in- 0 15 0 16 sight can be gained, whenever possible. The existence √ of additional nonresonant transitions between resonantly and the effective Rabi frequency Ω 1 ≡ ( 481/15) εΩ. connected states precludes the use of the standard adi- 1 ± 2 Here, ≡ δn,n0 denotes the identity operator. abatic elimination procedure. Nevertheless, the method The suppressed Rabi oscillations are separately oscil- of averaging introduced in Ref. [46] provides a systematic lating between the states |−~K/2i and |3~K/2i as well treatment for this kind of situations in terms of slow and as between |~K/2i and |−3~K/2i. To show the emer- fast contributions (corresponding in our case to the Rabi gence of these two separated diffraction processes, the frequency associated with two-photon processes and the 20 recoil frequency) as long as the adiabaticity parameter ε Our treatment demonstrates that higher-order Bragg given by the ratio of those two frequency scales is small. processes are also possible for double diffraction. More- For the deep Bragg regime, with ε  1, we find gener- over, the existence of an asymmetric AC Stark shift for alized Rabi oscillations within an effective three-level sys- the two resonant states (which can be easily compensated tem comprising states with momenta 0, ~K and −~K. in practice) has been established. This feature is absent By an appropriate choice of pulse duration and laser in- for single diffraction. In that case there is a frame where tensity one can generate the analogue of π/2 and π pulses, the two counterpropagating beams give rise to a standing acting, respectively, as beam splitters and mirrors. In ad- wave and where the two resonant states are symmetric. dition, we have also considered nonvanishing initial mo- This cannot be done simultaneously for the two pairs menta, for which the resonance condition is no longer of counterpropagating beams employed in double diffrac- fulfilled exactly. Analytical results have been obtained tion. for small momenta, whose effect is analogous to that of Having orthogonal polarizations for the copropagating a detuning and which lead to damped Rabi oscillations beams and for the equal-frequency counterpropagating for initial states with nonvanishing momentum width. It ones is crucial in order to avoid spurious transitions, as is, however, worth pointing out that for larger initial emphasized in the introduction. While we have restricted momenta the states with opposite momentum transfer our attention to circularly polarized light beams in the (+n~K and −n~K) are no longer equally populated, as main body of the paper, the case of linear polarizations illustrated by the exact result on “quasi-resonances” of and an arbitrary direction of the magnetic field defining Sec. IV E. This can be intuitively understood by consid- the quantization axis and determining the Zeeman split- ering the transformation to the reference frame where the ting is analyzed in detail in Appendix B. initial momentum vanishes and noticing that in this case Throughout most of the article we have focused on the the frequency detuning for the two pairs of counterprop- case of square pulses, but many of our results can be agating beams is not the same. extended to smooth time-dependent pulses, as explained With our approach we have also been able to study in Appendix C. The advantages of using smooth pulses analytically the quasi-Bragg regime, with ε < 1 but not were already pointed out for single diffraction in Ref. [23]. too small (e. g. ε ∼ 0.1). This is a regime of particular Similar advantages apply to the case of double diffrac- interest because it relaxes somewhat the effect of veloc- tion. Furthermore, the dependence on the laser phases in ity selectivity and it corresponds to a parameter range the contributions that give rise to the fast superimposed typically accessible in experiments; in particular it makes oscillations is potentially rather problematic for precision possible Bragg scattering of relatively high order without interferometry, and for double diffraction they appear al- 2 excessively long pulse durations. This regime was stud- ready at order ε rather than ε . Fortunately, for smooth ied for single Bragg diffraction in Ref. [23]. For square pulses, such as a Gaussian profile, these contributions are 2 pulses our analysis of the double diffraction case reveals instead exponentially suppressed like exp(−1/ε ) (which the existence of fast oscillations with smaller amplitude falls off faster then any power as ε → 0) as shown in superimposed on the slow generalized Rabi oscillations Appendix C 2 b. We plan to address these issues in more between resonant states. The amplitude of these oscilla- detail in a future analysis of the response of an interfer- tions for double diffraction is of order ε, in contrast with ometer based on double diffraction. the single diffraction case, where the amplitude is of order ε2 and, hence, much more suppressed. Our analytical re- sults for these oscillations are in excellent agreement with ACKNOWLEDGMENTS numerical solutions, as shown in Fig. 9. It should be stressed that besides being well suited We thank H. Ahlers, S. Kleinert, M. Meister, to problems where the conventional adiabatic elimina- V. Tamma, and W. Zeller for their support and many tion procedure cannot be applied, such as double Bragg fruitful discussions. A. R., E. M. R. and W. P. S. diffraction, the method of averaging can also be valuable acknowledge financial support by the German Space for studying certain aspects of single diffraction since it Agency (DLR) with funds provided by the Federal Min- provides a simpler and mathematically more transparent istry of Economics and Technology (BMWi) due to an en- description. An example of that are the superimposed actment of the German Bundestag under Grant No. DLR fast oscillations for square pulses in single Bragg diffrac- 50WM1131-1137 (project QUANTUS-III). G. T. thanks tion. These have been found [23] when solving exactly the the Max-Planck-Gesellschaft for financial support. Mathieu equation for a pulse of constant amplitude [54] and matching the solution to the free solution (in absence of external electromagnetic field) before and after the square pulse. A satisfactory description when employing Appendix A: Method of averaging to arbitrary adiabatic expansions requires a proper understanding of orders the subtle connection between “dressed” states and the “bare” initial states typically accessible in experiments, In Sec. II, we introduced the method of averaging [46] as discussed in Sec. III D. and the two lowest approximations. In this appendix, we 21 now show that this method is valid to arbitrary orders differential equation and we derive the corresponding expressions. (m) (m) m+1 For this purpose, we recall from Sec. II that we have g˙ = iεH g + O ε to solve the equation up to the m-th order in ε. X g˙ = iεH g + iε eiνωrt H g . At each order, we assume that the slowly evolving part 0 ν (m) ν=0 γ fulfills the equation 6 m The ansatz (m) (m) X µ (m) γ˙ = iεH0 γ + i ε pµ(γ ) . (A2) m µ=2 (m) (m) X j (m) g = γ + ε f j(γ ) , (A1) j=1 The task is now to find the f j and pj. For this rea- son, we use Eqs. (A1) and (A2) above and equate the separating the slow from the fast oscillations, satisfies the coefficients

! ! ∂f ∂f (γ(m)) ∂f ˙ (m) H (m) 1 2 (m) 1 H (m) 2 g =ε i 0 γ + + ε ip2(γ ) + i (m) 0 γ + ∂t γ(m) ∂γ ∂t γ(m)

m  j 2  (m) − (m) X j (m) ∂f j 1(γ ) (m) ∂f j X ∂f µ(γ ) (m) m+1 − H O + ε ipj(γ ) + i (m) 0 γ + + i (m) pj µ(γ ) + ε . ∂γ ∂t (m) ∂γ − j=3 γ µ=1

to the ones on the right-hand side of Eq. (3), i. e. with iteratively to arbitrary orders of ε. In order to illustrate this technique we now consider (m) m+1 (m) 2 (m) k iεH g + O ε = iεH γ + iε H f 1(γ ) the case of ε . The functions f j can be written as m iµω t X j (m) m+1 X e r + i ε Hf j 1(γ ) + O Ω . (m) (j) (m) − f j(γ ) = Φµ γ µωr j=3 µ=0 6

The terms proportional to ε and ε2 were dealt with in for j < k and each of the slow corrections pj is chosen to (m) (j) (m) Sec. II, where we have determined f 1 and f 2 for a spe- be of the from pj(γ ) = Φ0 γ . Then, the differen- cific choice of p1 and p2. We can extend this treatment tial equation for f k reads

k 2 iωrµt iωrµt iµωrt ∂f k X e (k 1) (m) X e (k 1) (m) X− X e (l) (k l) (m) (m) =iH Φµ − γ − i Φµ − H0 γ − i Φµ Φ0 − γ − ipk(γ ) . ∂t (m) µωr µωr µωr γ µ=0 µ=0 l=1 µ=0 6 6 6

We now define equation

(k 1) (k 1) k 2 (l) (k l) (k) X Hµ ν Φν − Φµ − H0 X− Φµ Φ0 − Φµ ≡ − − − νωr µωr µωr ν=0 l=1 6 ∂f k X iµωrt (k) (m) = i e Φµ γ for µ =6 0 and ∂t (m) γ µ=0 6 (k 1) (k) X H ν Φν − Φ0 ≡ − νωr ν=0 6 which can again be integrated. In this way, all orders up (m) (k) (m) so that with pk(γ ) = Φ0 γ we get the differential to m can be iteratively determined. 22

Appendix B: Double Bragg diffraction with This orthogonality relation will prove to be crucial later. orthogonal linear polarizations We now assume the atomic level structure shown in Fig. 20. Since the magnetic field in z0-direction is re- In Sec. IV A we considered circularly polarized laser sponsible for the Zeeman splitting, we consider all polar- beams. In this appendix we show that even for different izations with respect to this quantization axis. orthogonal polarizations and an arbitrary magnetic field we arrive at the same differential equations used through- = 1 out this article, provided certain requirements are satis- mz |e+i fied. mz = 0 mz = −1 |ezi λ/4 plate |e−i B ω , σ ω , σ ω+ a y b y ℘− ωz ℘z ℘+ pair 1 ω− pair 2 ω , σ ω , σ b x atom a x ′ |gi mz = 1 z z mirror mz = 0 mz = −1 FIG. 19. Linear-linear configuration for double Bragg diffrac- tion. The magnetic field B is not in the z-direction of the FIG. 20. Atomic level structure. We consider three excited 0 pair of light fields, but in a different one denoted by z . The states e and ez with the frequencies ω and ωz, respec- ± ± polarizations of the two pairs are linear and orthogonal to the tively,| andi one ground| i state g . The dipole moments ℘ and σ σ | i ± z-direction of propagation, i. e., x and y. ℘z initiate transitions between sublevels and correspond to the polarizations σ0 and σz0 in the primed frame. ± We now assume an experimental set-up as shown in Fig. 19. The linear polarizations σx and σy are orthogo- nal and named with respect to the direction of propaga- The dipole transitions induced by the electric field tion z of the light fields. The magnetic field B responsible from Eq. (B1), exciting the states |e i and |ezi from ± for a Zeeman splitting and determining the quantization the ground state |gi, that is from the magnetic quantum axis points in a different direction z0. number mz = 0, correspond to each of the three polariza- With respect to the propagation direction of the light, tions in the primed frame. The frequencies of the excited the electric field reads states are ω and ωz, and the dipole moments are defined ± h i as ˆ i(kbzˆ ωbt) i( kbzˆ ωbt) E =Eb σx e − +σy e − − h i 1 i( kazˆ ωat) i(kazˆ ωat) ˆ + Ea σx e − − +σy e − + h.c. ℘ ,z ≡ − he ,z| d · σ0 ,z |gi , ± ~ ± ± (B1)

Writing the polarizations as a linear combination of po- where dˆ denotes again the dipole operator. Note that the larizations σ0 and σz0 , which are now with respect to the amplitudes of the electric fields are not included in this ± z0-direction of the quantization axis of the atom, we get definition.

(k) (k) (k) X (k) Electric field and atom couple through dipole interac- σk = a+ σ+0 + a σ0 + az σz0 = aj σj0 − − tion, which leads to the interaction j= ,z ± for k = x, y. Here, σ0 are circular polarizations orthog- h   X iωbt (x) ikbzˆ (y) ikbzˆ ± HˆI = ℘∗ e− Eb a e +a e− onal to the z0-direction and σz0 is the unit vector along ~ j j j j= ,z the z0-direction. ±  i Since in the unprimed frame the polarizations were iωat (x) ikazˆ (y) ikazˆ + e− Ea a e− +a e |eji hg| orthogonal, we find with the use of the orthogonality in j j the primed frame the relation + h.c.

X (k) (l) δk,l =σk · σl∗ = aj ai∗ σj0 · σi0 ∗ j,i= ,z which is much more complicated than Eq. (28). Never- ± X (k) (l) X (k) (l) theless, we can derive the differential equations for the = aj ai∗ δj,i = aj aj∗ . (B2) coefficients in the same way as in Sec. III B and find, j,i= ,z j= ,z ± ± when changing into the , 23

 (x) (y)  i(ωb ωj )t i(ω kb +ν kb )t i(ωkb +νkb )t ie˙j(p) =℘j∗ e− − Eb e− − − aj g(p − ~kb) + e− aj g(p + ~kb)

 (x) (y)  i(ωa ωj )t i(ωka +νka )t i(ω ka +ν ka )t + ℘j∗ e− − Ea e− aj g(p + ~ka) + e− − − aj g(p − ~ka) and

X h  (x) (y)  i(ωb ωj )t i(ωkb +νkb )t i(ω kb +ν kb )t ig˙(p) = ℘j e − Eb∗ e− aj∗ ej(p + ~kb) + e− − − aj∗ ej(p − ~kb) j= ,z ±  (x) (y) i i(ωa ωj )t i(ω ka +ν ka )t i(ωka +νka )t + e − Ea∗ e− − − aj∗ ej(p − ~ka) + e− aj∗ ej(p + ~ka) ,

2 where we have recalled the definitions ωk ≡ ~k /(2M) and νk ≡ pk/M from Sec. III B. We can again eliminate the excited states ej(p) adiabatically and find the resulting differential equations for the ground state

 2  2 2 2  2 2 X |℘jEb| |℘jEa| ig˙(p) =g(p) a(x) + a(y) + a(x) + a(y) ω − ω j j ω − ω j j j= ,z b j a j ±  i(ωb ωa)t 2 i(ωa ωb)t 2 X 2 i(ωK +νK )t e − (x) e − (y) +g(p + ~K) |℘j| e− EaE∗ a + EbE∗ a ω − ω b j ω − ω a j j= ,z a j b j ±  i(ωb ωa)t 2 i(ωa ωb)t 2 X 2 i(ω K +ν K )t e − (y) e − (x) +g(p − ~K) |℘j| e− − − EaE∗ a + EbE∗ a ω − ω b j ω − ω a j j= ,z a j b j ± | |2 | |2 X ℘jEb i(ω2k +ν2k )t (x) (y) X ℘jEb i(ω 2k +ν 2k )t (y) (x) +g(p + 2~kb) e− b b a∗ a + g(p − 2~kb) e− − b − b a∗ a ω − ω j j ω − ω j j j= ,z b j j= ,z b j ± ± | |2 | |2 X ℘jEa i(ω2k +ν2k )t (y) (x) X ℘jEa i(ω 2k +ν 2k )t (x) (y) +g(p + 2~ka) e− a a a∗ a + g(p − 2~ka) e− − a − a a∗ a ω − ω j j ω − ω j j j= ,z a j j= ,z b j ± ±  i(ωb ωa)t i(ωa ωb)t  X 2 i(ω∆k+ν∆k)t e − e − (y) (x) +g(p + ~∆k) |℘j| e− EaE∗ + EbE∗ a a∗ ω − ω b ω − ω a j j j= ,z a j b j ±  i(ωb ωa)t i(ωa ωb)t  X 2 i(ω ∆k+ν ∆k)t e − e − (x) (y) +g(p − ~∆k) |℘j| e− − − EaE∗ + EbE∗ a a∗ , (B3) ω − ω b ω − ω a j j j= ,z a j b j ±

where we defined ∆k ≡ kb − ka. spurious terms vanish. Equation (B3) is quite cumbersome and, in general, Hence, we can conclude that for a linear-linear config- cannot be solved with our method of averaging since it uration we can use Eq. (29) , as long as the magnetic couples to eight different momenta of different timescales. field is not too strong, i. e. the Zeeman splitting ω+ − ω These are the spurious couplings mentioned in Sec. IV A. is not too large. Very similar arguments hold true for− Fortunately, we can neglect all terms except the coupling circular polarizations with a magnetic field not aligned to g(p ± ~K) under the following conditions: with the light propagation. If we assume all ℘j to be identical, which is a reason- able argument since the light fields Ea,b are not included in the definition, and demand that the detuning ∆ is Appendix C: Pulse shapes much larger than the Zeeman splitting, i. e. ∼ ωj − ωa,b = ∆ Throughout this paper, we have focused on box-shaped pulses, where we have a time-dependent pulse in the form for j = ±, z, which is fulfilled under the condition of a step function. We refer at certain stages of the dis- ω+ − ω  ∆, we can use the orthogonality relation of cussion of the time-evolution to π/2 and π pulses, de- Eq. (B2)− since all other terms are independent of j and pending on the duration of the pulses. They create an perform the sums. The only surviving terms of Eq. (B3) equal superposition of two momentum states, or an ex- lead to Eq. (29). Indeed, the sums proportional to the change of their populations, respectively. 24

This approximation does not correspond to most ex- the pulse shape of this Rabi frequency reads κ(t) =κ ˜2(t). perimental situations, where the lasers are turned on and Hence, ε has to be replaced by εκ(t). According to off according to a certain time-dependent function that Eq. (4), the differential equations for the slowly oscil- differs from a step function. In fact, a smooth pulse lating terms take, in general, the form results in more convenient properties of the populated states. Of course, this has subtle implications for the γ˙ = iA(t)γ , method of averaging. This section of the appendix deals where A(t) is now a time-dependent square matrix. with the modification of our method for smooth time- dependent pulse shapes. In this case, the Rabi frequen- cies are proportional to the amplitudes of the electric a. Leading-order approximation fields; i. e., we get a time-dependent frequency Ω˜ κ ˜(t), where the functionκ ˜ =κ ˜(t) corresponds to the pulse shape. To first order, we find A(t) ≡ εκ(t)H0 for the first- order Bragg resonance (∆ω = ωr), where the time- dependence can be factored out and the solution of this 1. Adiabaticity condition differential equation reads   Zt For the adiabatic elimination of the excited state, we (1) (1) γ (t) = exp iε dt0κ(t0)H0γ (t0) . integrate a differential equation of the same form as Eq. (17), namely t0 Hence, in the solution of the slowly oscillating term we ˜ i∆t e˙(t) = −iΩκ ˜(t) e g(t) , just have to replace Ω with Ω R dtκ(t). This substitution coincides with other discussions, e. g., the one of Ref. [23]. where we now have included the pulse shape. Integrating For higher orders, the solution is not that elementary, this equation by parts we find since the time-dependence of A(t) cannot be factored out anymore. So, in general, the quasi-Bragg regime Zt with time-dependent pulses cannot be treated with our e(t) = − i dt Ω˜ κ ˜(t ) ei∆t0 g(t ) 0 0 0 method up to higher than leading order in the adiabatic- t0 ity parameter ε. t ˜ Z ˜ ˙ On the bright side, we showed in Sec. IV D that for Ωκ ˜(t) i∆t Ω κ˜(t0) i∆t = − e g(t) + dt0 e g(t0) second-order Bragg diffraction the matrices simplify and ∆ ∆ we again can factor out the complete time-dependence. t0 t In this sense, finding slowly oscillating solutions with ˜ ˜ ˙ Z Ωκ ˜(t) i∆t Ω κ˜(t) i∆t higher-order accuracy than the leading order is not pos- = − e g(t) + e g(t) − dt0 ..., ∆ i∆2 sible for time-dependent pulses, but the leading order t0 with rapidly oscillating corrections can be calculated. In (C1) second-order double Bragg diffraction we thus can replace Ω2 with Ω2 R dt κ2(t), which is in complete agreement where we have again assumed g(t) to be slowly varying in with Ref. [23]. comparison to ∆, as in Sec. III B and used the fact that But unfortunately, an analytic treatment of p =6 0 is ˙ 2 κ˜(t0) = 0. If now |κ˜(t)/∆ |  |κ˜(t)/∆|, which can be much more complicated, since time-independent terms ˙ ∼ written as |κ/˜ κ˜| = 1/T  ∆ with a characteristic pulse νD in the matrices appear and hence even in leading or- time T , we can neglect the contribution coming from the der the time-dependence cannot be factored out. In the first integration by parts. This condition can be seen as treatment of Ref. [23] the quasi-Bragg regime for time- an adiabaticity condition for the pulse shape compared dependent pulses is also described just for vanishing mo- to the high-frequency oscillation of the highly detuned mentum width. excited state. Neglecting also the higher-order terms, we recover the already known expression for the adiabatic elimination. In this way, we get the very same equation b. Adiabatic pulse shapes in the method of averaging for the ground state, but now with a time-dependent ef- fective Rabi frequency. When solving the partial differential equation for the rapidly oscillating terms f j, one can perform an integra- tion by parts analogously to Eq. (C1). In the same way, 2. Method of averaging with time-dependent pulses ∼ we find the condition |κ/κ˙ | = 1/T  ωr, which clearly shows that the role of the detuning is now played by the We can incorporate time-dependent pules shapes in our recoil frequency. equations just by replacing H by κ(t) H. After the adia- This condition corresponds in the limit to the case batic elimination we get a Rabi frequency Ω = Ω˜ 2/∆, and where the lasers are turned on and off adiabatically. On 25 the other hand, the pulse duration cannot be arbitrarily where H˜ν ≡ Hν /ωr are now the dimensionless matrices. long if we want to perform π/2 or π pulses, so the char- We see now that in the deep Bragg regime these correc- acteristic time is of the order of Ω T ∼ π which limits the tions are suppressed by a Gaussian exp −1/ε2, which pulse duration. Together with this condition above we falls off faster than any power of ε as ε approaches 0. find Thus, the scaling is better in comparison with the box- shaped pulses where the correction is suppressed by ε. 1 Ω ∼  1 , ωrT ωrπ which is fulfilled only in the Bragg regime Ω/ωr  1. This requirement is a consequence of the differential equation satisfied by f j, for example Eq. (5) for f 1

∂f 1 X iνωrt (m) 3. Dressed states = κ(t) i e Hν γ , (C2) ∂t (m) γ ν=0 6 where we have now included the time dependence κ(t) Furthermore, time-dependent pulses make the pictures of the pulse. As one can see from Eq. (C2), γ(m) is of dressed or bare states obsolete, because in this case treated as a constant. This makes it possible to perform both coincide, which can be seen from the initial con- the integration for certain pulse shapes even analytically. dition Eq. (10). Indeed, for time-dependent pulses we For instance, if we assume a Gaussian pulse get

κ(t) = exp −t2/T 2 , √   1 where the characteristic time√ T = π/(2Ω) corresponds − (1) X Hν (1) to a π/2 pulse and T = π/Ω to a π pulse. In the γ (t0) = 1 + ε κ(t0)  g (t0) . ν following we discuss π pulses, but the difference in the ν=0 π/2 pulse is just a factor 1/2. 6 We can now integrate Eq. (C2) easily over all times (i. e., neglecting the experimental truncation of the Gaus- sian pulse) and find Since the light fields are initially switched off, κ(t0) van- ishes and thus √ 2 X (νωrT ) /4 (m) f 1 = πT e− Hν γ . ν=0 6 (1) (1) Since Hν is proportional to ωr, we find for π pulses the γ (t0) = g (t0) . correction

√ 2 Ω X (νωrT ) /4 (m) εf 1 = πT e− ωrH˜ν γ ωr As a consequence the fast oscillatory behavior in single ν=0 6 Bragg diffraction found in Fig. 4, which is an artefact of " √ 2# X  πν  the bare-state formulation, vanishes for time-dependent =π exp − H˜ γ(m) , 2ε ν pulses and the population of the states is better described ν=0 6 by the smooth oscillations seen in Fig. 3.

[1] T. van Zoest, N. Gaaloul, Y. Singh, H. Ahlers, W. Herr, [6] M. Hauth, C. Freier, V. Schkolnik, A. Senger, S. Seidel, W. Ertmer, E. Rasel, M. Eckart, E. Kajari, M. Schmidt, and A. Peters, Appl. Phys. B 113, 49 et al., Science 328, 1540 (2010). (2013). [2] R. Geiger, V. M´enoret,G. Stern, N. Zahzam, P. Cheinet, [7] µQuanS Precision Quantum Sensors, http://www. B. Battelier, A. Villing, F. Moron, M. Lours, Y. Bidel, muquans.com. et al., Nature communications 2, 474 (2011). [8] W. Herr et al., in preparation. [3] H. M¨untinga et al., Phys. Rev. Lett. 110, 093602 (2013). [9] M. A. Kasevich and B. Dubetsky, Kinematic sensors [4] STE-QUEST space mission in ESA’s Cosmic Vision employing atom interferometer phases (United States 2020-22, http://sci.esa.int/ste-quest. Patent 7317184). [5] Q. Bodart, S. Merlet, N. Malossi, F. Pereira dos Santos, [10] D. S. Durfee, Y. K. Shaham, and M. A. Kasevich, Phys. P. Bouyer, and A. Landragin, Applied Physics Letters Rev. Lett. 97, 240801 (2006). 96, 134101 (2010). 26

[11] M. Schmidt, A. Senger, M. Hauth, C. Freier, V. Schkol- [36] S. M. Dickerson, J. M. Hogan, A. Sugarbaker, D. M. S. nik, and A. Peters, Gyroscopy and Navigation 2, 170 Johnson, and M. A. Kasevich, Phys. Rev. Lett. 111, (2011). 083001 (2013). [12] W. L. Bragg, Proc. Cambridge Philos. Soc. 17, 43 (1912). [37] T. L´ev`eque,A. Gauguet, F. Michaud, F. Pereira Dos San- [13] W. Friedrich, P. Knipping, and M. Laue, Annalen der tos, and A. Landragin, Phys. Rev. Lett. 103, 080405 Physik 346, 971 (1913). (2009). [14] H. Rauch, W. Treimer, and U. Bonse, Physics Letters A [38] B. Dubetsky and P. R. Berman, Phys. Rev. A 66, 045402 47, 369 (1974). (2002). [15] H. Rauch and S. A. Werner, Neutron interferometry: [39] N. Malossi, Q. Bodart, S. Merlet, T. L´ev`eque,A. Lan- Lessons in experimental quantum (Oxford dragin, and F. P. D. Santos, Phys. Rev. A 81, 013617 University Press, Oxford, 2000). (2010). [16] A. D. Cronin, J. Schmiedmayer, and D. E. Pritchard, [40] Y. Le Coq, J. Retter, S. Richard, A. Aspect, and Rev. Mod. Phys. 81, 1051 (2009). P. Bouyer, Applied Physics B 84, 627 (2006). [17] P. L. Gould, G. A. Ruff, and D. E. Pritchard, Phys. Rev. [41] The injected polarizations also need to be orthogonal for Lett. 56, 827 (1986). the case of linear polarizations. In contrast, the linear po- [18] P. J. Martin, B. G. Oldaker, A. H. Miklich, and D. E. larizations must be parallel for Raman transitions. This Pritchard, Phys. Rev. Lett. 60, 515 (1988). difference is a consequence of the change of total angular [19] E. M. Rasel, M. K. Oberthaler, H. Batelaan, J. Schmied- momentum by ∆F = 1 in the latter case, whereas the mayer, and A. Zeilinger, Phys. Rev. Lett. 75, 2633 internal state remains± unaltered for Bragg diffraction. (1995). [42] S.-w. Chiow, T. Kovachy, H.-C. Chien, and M. A. Kase- [20] S. Fray, C. A. Diez, T. W. H¨ansch, and M. Weitz, Phys. vich, Phys. Rev. Lett. 107, 130403 (2011). Rev. Lett. 93, 240404 (2004). [43] A. F. Bernhardt and B. W. Shore, Phys. Rev. A 23, 1290 [21] P. Haslinger, N. D¨orre,P. Geyer, J. Rodewald, S. Nimm- (1981). richter, and M. Arndt, Nature Phys. 9, 144 (2013). [44] M. Marte and S. Stenholm, Applied Physics B 54, 443 [22] D. M. Giltner, R. W. McGowan, and S. A. Lee, Phys. (1992). Rev. A 52, 3966 (1995). [45] E. Brion, L. H. Pedersen, and K. Mølmer, Journal [23] H. M¨uller,S.-W. Chiow, and S. Chu, Phys. Rev. A 77, of Physics A: Mathematical and Theoretical 40, 1033 023609 (2008). (2007). [24] P. D. Lett, R. N. Watts, C. I. Westbrook, W. D. Phillips, [46] N. N. Bogoliubov and Y. A. Mitropolsky, Asymptotic P. L. Gould, and H. J. Metcalf, Phys. Rev. Lett. 61, 169 methods in the theory of non-linear oscillations (Hindus- (1988). tan Publishing Corpn., Delhi, 1961). [25] J. Dalibard and C. Cohen-Tannoudji, J. Opt. Soc. Am. [47] W. P. Schleich, Quantum Optics in Phase Space (Wiley- B 6, 2023 (1989). VCH, Weinheim, 2001). [26] C. Bord´e,Physics Letters A 140, 10 (1989). [48] S. Meneghini, I. Jex, K. A. H. van Leeuwen, M. R. Kasi- [27] M. Kasevich and S. Chu, Phys. Rev. Lett. 67, 181 (1991). mov, W. P. Schleich, and V. P. Yakovlev, Laser Physics [28] A. Peters, K. Y. Chung, and S. Chu, Nature (London) 10, 116 (2000). 400, 849 (1999). [49] D. Chudesnikov and V. Yakovlev, Laser Phys 1, 110 [29] K. Moler, D. S. Weiss, M. Kasevich, and S. Chu, Phys. (1991). Rev. A 45, 342 (1992). [50] A. Kazantsev, G. Surdutovich, and V. Yakovlev, Me- [30] M. Kozuma, L. Deng, E. W. Hagley, J. Wen, R. Lutwak, chanical Action of Light on Atoms (World Scientific Pub- K. Helmerson, S. L. Rolston, and W. D. Phillips, Phys. lishing Company Incorporated, 1989). Rev. Lett. 82, 871 (1999). [51] D. F. James and J. Jerke, Canadian Journal of Physics [31] Y. Torii, Y. Suzuki, M. Kozuma, T. Sugiura, T. Kuga, 85, 625 (2007). L. Deng, and E. W. Hagley, Phys. Rev. A 61, 041602 [52] P. Clad´e, E. de Mirandes, M. Cadoret, S. Guellati- (2000). Kh´elifa,C. Schwob, F. Nez, L. Julien, and F. Biraben, [32] J. E. Debs, P. A. Altin, T. H. Barter, D. D¨oring,G. R. Phys. Rev. A 74, 052109 (2006). Dennis, G. McDonald, R. P. Anderson, J. D. Close, and [53] A. Gauguet, T. E. Mehlst¨aubler, T. L´ev`eque, N. P. Robins, Phys. Rev. A 84, 033610 (2011). J. Le Gou¨et, W. Chaibi, B. Canuel, A. Clairon, [33] S. S. Szigeti, J. E. Debs, J. J. Hope, N. P. Robins, and F. P. Dos Santos, and A. Landragin, Phys. Rev. A 78, J. D. Close, New Journal of Physics 14, 023009 (2012). 043615 (2008). [34] H. M¨uller,S.-w. Chiow, Q. Long, S. Herrmann, and [54] C. L¨ammerzahland C. J. Bord´e,Physics Letters A 203, S. Chu, Phys. Rev. Lett. 100, 180405 (2008). 59 (1995). [35] A. Peters, K. Y. Chung, and S. Chu, Metrologia 38, 25 (2001).