Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Excitation of atoms in an optical lattice driven by polychromatic amplitude modulation

Open Access Open Access

Abstract

We investigate the mutiphoton process between different Bloch states in an amplitude modulated optical lattice. In the experiment, we perform the modulation with more than one frequency components, which includes a high degree of freedom and provides a flexible way to coherently control quantum states. Based on the study of single frequency modulation, we investigate the collaborative effect of different frequency components in two aspects. Through double frequency modulations, the spectrums of excitation rates for different lattice depths are measured. Moreover, interference between two separated excitation paths is shown, emphasizing the influence of modulation phases when two modulation frequencies are commensurate. Finally, we demonstrate the application of the double frequency modulation to design a large-momentum-transfer beam splitter. The beam splitter is easy in practice and would not introduce phase shift between two arms.

© 2015 Optical Society of America

1. Introduction

Ultracold atoms in periodically driven optical lattices, including shaken and amplitude modulated systems, have attracted much attention recent years, for they can bring out new interesting phenomena including realization of artificial gauge fields in different lattice geometries [13] and the coherent control of atomic wavefunctions [46]. In a shaken one dimensional optical lattice, the ferromagnetic transition in trapped Bose gas has also been observed [7] by coupling ground band s of the lattice system to the first excited band p. Specific to amplitude modulated lattice systems, although only bands with the same parity can be coupled, there are still researches on a wide range of problems, such as transfer of atoms from the ground band s to the second excited band d [8], detection of superfluid-Mott insulator transition [9] and study of the dynamical tunnelling of ultracold atoms with quantum chaos [10, 11]. The technique can also be applied in areas including realization of a velocity filter [12] and detection of gravity [13,14].

Normally, studies of amplitude modulated lattices are based on single frequency modulations that only involve processes with one photon emission or absorption. During a polychromatic modulation, not only the amplitudes, but also the phases of different frequency components in the modulation can be controlled independently, providing a more flexible way to coherently manipulate quantum states. In this paper, we coherently transfer atoms from the ground state s to the high excited g band via a double frequency modulation. The peaks of transfer rate with different lattice depths are measured experimentally, while influence of the modulation phase is demonstrated by performing an interference between two independent paths of excitations. These experiments completely investigated how frequencies and phases of different modulation frequency components would influence the excitation between Bloch states. Furthermore, we show an application of the double frequency modulation to build a large-momentum-transfer (LMT) beam splitter. Comparing with LMT beam splitters based on Bloch oscillation [8,15] or high order Bragg scattering [16], our method is easy in practise and would not introduce phase shift between two separated atom clouds.

The paper is organized as follows: In Sections 2 and 3 the polychromatic modulation theory based on Floquet method and our experimental system are introduced respectively. In Section 4, we study the effects of the modulation phase for the single frequency modulation. In Section 5 collaborative effect of different frequency components is studied, demonstrate the effect of both modulation frequencies and phases. Section 6 presents a way to realize a LMT beam splitter. Discussion and conclusion are in Section 7.

2. Theory for optical lattice with polychromatic modulation

For an atom in an amplitude modulated lattice system along x^ axis, as schematically shown in Fig. 1, the time dependent Hamiltonian can be written as

H(t)=px22M+V0cos2(kLx)+iVicos(ωit+ϕi)cos2(kLx).

 figure: Fig. 1

Fig. 1 A sketch of lattice depth modulation in our system. The depth of lattice potential V0 cos2(kLx) is driven by a polychromatic modulation ∑iVi cos(ωit +ϕi).

Download Full Size | PDF

The first term in the right hand side is kinetic energy with M the atom’s mass and px its momentum along the x direction. The second term represents optical lattice without the modulation. V0 is the constant part of the lattice depth, and the wave vector is kL = 2π/λ with λ the laser wavelength. The last term expresses the amplitude modulation with the modulation amplitude Vi, the frequency ωi and the phase ϕi of each frequency components.

Typically, time-periodic systems are described by Floquet’s theorem [17]. In our system, each of the modulation frequencies ωi gives a period Ti = 2π/ωi and the Hamiltonian has a period of H(t + T) = H(t), with T the lowest common multiple of Ti. By defining a time-evolution operator for one period as U^(T), solutions to this problem must satisfy

|ψq,α(t+T)=U^(T)|ψq,α(t)=eiεq,αT|ψq,α(t),
and the Floquet states |uq,α are defined as, |ψq,α=eiqxiεq,αt|uq,α with the quasi-momentum q the band index α and quasi-energy εq,α , which leads to
(H(t)it)|uq,α=H0|uq,α=εq,α|uq,α.

For a special case of single frequency modulation, considering the time and coordinate periodicity of the Floquet states |uq,α, by a Fourier transformation we can study the problem with a set of basis |vlm=ei(2lkLxmω1t) in an extended Hilbert space S=HT. Where H is the Hilbert space and T is the space of all functions with periodic T. By integration in one period, we can turn to a time-independent system.

Alm,lm=1T0T|vlmH0vlm|dt.

To get rid of the phase ϕ1, we perform an unitary transformation Ulm,lm=eimϕ1 to the Hamiltonian by A′ = U1AU. There are three kinds of terms in the matrix of A′. The diagonal terms Alm;lm=(2lh¯kL)2/2M+mh¯ω1 are the energy of momentum states shifted by absorbing or emitting Floquet photons with energy h¯ω1. The terms Alm;l±1,m=V0/4 show stationary component of the lattice would coupling atoms with momentum difference 2h¯kL. In addition, Alm;l±1,m±1=V1/8 terms show the time modulation of the lattice, which would coupling two momentum states separated by 2h¯kL while absorbing or emitting one Floquet photon.

When the modulation frequency is near-resonant with the energy difference between two specific bands and far detuned from the others, we can get an effective Hamiltonian by a rotating wave approximation.

HR=(HRV1eiϕ1Ωαβ*V1eiϕ1Ωαβ*Eβh¯ω),
where Eα and Eβ are the energy of Bloch states , |β〉 in the system without modulation. Coupling constant between two states is Ωαβ =〈α|cos2(kLx) |β〉/4.

The extension to polychromatic driven is straightforward. For example, a double frequency modulation induced two-photon process between s-g band is described by an effective Hamiltonian Hsg as:

Hsg=(Eseiϕ1V1Ωsdeiϕ2V2Ωsd000eiϕ1V1Ωsd*Edh¯ω10eiϕ2V2Ωdgeiϕ1V1Ωdg0eiϕ2V2Ωsd*0Edh¯ω2eiϕ1V1Ωdg0eiϕ2V2Ωdg0eiϕ2V2Ωdg*eiϕ1V1Ωdg*Egh¯(ω1+ω2)000eiϕ1V1Ωdg*00Eg2h¯ω1000eiϕ2V2Ωdg*00Eg2h¯ω2),

The effective Hamiltonian Hsg is constructed by means of nearly degenerate perturbation technique [18], in which we include six nearly degenerate states considering four main processes in the excitation. The six states are |Es〉 the s band, |Edh¯ω1, |Edh¯ω2 the d band dressed by Floquet photon ω1 or ω2 and |Esh¯(ω1+ω2), |Eg2h¯ω1 and |Eg2h¯ω2 the g band dressed by two Floquet photons. Using this basis a general state (v1, v2, v3, v4, v5, v6)T gives complex coefficient of the six dressed states. Population of Bloch states s is |v1|2, while population on g band is |v4ei(ω1+ω2)t+v5e2iω2t|2, given by coherent superposition of all g band states dressed with different Floquet photons. Solution of the model consists with the time dependent Schrödinger equation and the effective model provides us a better understanding of the mutiphoton process. However, in the calculation more states associated with higher order processes could be included to get a more accurate result, especially when the modulation amplitude is large.

3. Experimental system

Our experiment begins with a quasi-pure condensate of typically 1.5×105 87Rb atoms in the |F = 2,mF = 2〉 hyperfine ground state, produced in a combined potential of a single-beam optical dipole trap and a quadruople magnetic trap. The trapping frequencies are ωx = 2π × 28Hz,ωy = 2π × 60Hz,ωz = 2π × 70Hz. The optical lattice is formed by a retro-reflected red detuned laser beam, with lattice constant a = λ/2 = 426nm focused to a waist of 110μm. Density of atoms in our system is less than 5×1013cm3, and the mean-field interaction can be omitted to capture the main physical mechanism in the excitation [19,20].

The lattice depth is calibrated by Kapitza-Dirac scattering and the modulation of lattice depth is controlled by an acousto-optic modulator(AOM). The modulation amplitudes Vi and phases ϕi are generated from a signal generator and the intensity of lattice laser is monitored by a photodetector. Experimental results are absorption images taken after 28ms time-of-flight (TOF). Occupation number at different momentum states |2lh¯kL (l is integral momentum index) can be given from TOF images by nl = Nl/N, with Nl the atom number at momentum state |2lh¯kL and N the total atom number. The initial state for experiment is prepared non-adiabatically with numerically designed sequence of lattice pulses [21, 22]. For experimental convenience, the lattice pulses are carried out with the same depth as the constant part of the modulated lattice potential V0. Typically, each of the pulses and the subsequent intervals are lasting for no more than 25μs, and the whole process can be finished within 60μs, thus the loading time is greatly reduced comparing with traditional adiabatic loading method.

4. The initial phase effect in single frequency modulation

Single frequency modulation can be seen as the basis of polychromatic driven. In this part, we present the preparation of a Floquet state in the single frequency driven system, which is shown to be highly related to the modulation phase.

Following the discussion in Sec. 2, Fig. 2 depicts a typical quasi-energy spectrum of the single frequency driven system, which is obtained by direct diagonalization of A′ at various quasi-momentum q. The same calculation also gives eigenvectors in the extended Hilbert space. The spectrum exhibits a complex structure as a result of the periodically repetition of high excited bands.

 figure: Fig. 2

Fig. 2 Left side is the calculated Floquet spectra of a single frequency driven system, with parameters V0 = 5Er, V1 = 0.5Er, h¯ω1=5Er. In the figure the first seven bands are presented. The heavy lines depict states maximally overlapping with the s(blue), p (green) and d (red) Bloch bands respectively. Right side shows the details of two Floquet bands most overlapping with s and d bands. The two bands are separated by a band gap EF at q = 0.

Download Full Size | PDF

To connect our calculation with the experiments, we should project the Floquet states |uq,α〉 into momentum space. The occupation number at momentum states 2lh¯kL and its phase can be given by a summation of all the Fourier components with the same l as cl(t)|2lh¯kL=meimω1τνlm|2lh¯kL, where νlm=νlm|uq,α is coefficient of the Floquet state. τ is related to the modulation phase ϕ1 and holding time t as τ=ω1t+ϕ12πT. It is also useful to define the overlapping between the Floquet state |uq,α 〉 and a Bloch state |nq〉 of the undriven potential as P=0T|lcl(t)|nq|2dt. Property of the Floquet band is typically characterized by its most overlapping Bloch band [23]. Without loss of generality, our experiments are restricted to quasimomentum q = 0, and energy gap between Bloch bands α and β are written as h¯ωαβ. The technique can also be applied in systems with acceleration [12], which brings out phenomena different from our study.

When modulation phase ϕ1 is given, a Floquet state can be projected into the momentum space, and such a state can be prepared by carrying out two lattice pulses with numerically designed pulse sequence [21, 22]. In the fast loading process, lattice depth of the two pulses is kept constant. For a target state, the fidelity of a prepared state can be given numerically for different pulse durations and time intervals, and the optimized pulse sequence is obtained by finding the maximum loading fidelity. Throughout this method we can get a loading fidelity of more than 95% experimentally.

With the initial modulation phase ϕ1 = −π/2, a pure Floquet state is loaded by lattice pulses calculated with τ=ϕ12πT=T/4. For the same state, we also present a modulation with ϕ1 = π/2 to show the influence of modulation phase. The population on momentum components |0h¯kL and |±2h¯kL for different initial phases ϕ1 = −π/2 and ϕ1 = −π/2 are shown in Fig. 3(a1) and 3(a2) respectively. The parameters are V0 = 5.0Er,V1 = 0.5Er and driven frequency is h¯ω1=5.0Er, red detuned from the band gap h¯ωsd=5.2Er. Figure 3(a1) is a Floquet state, which shows a time period of T = 2π/ω1 = 62.75μs. The time evolution in momentum space would be quite different when the modulation phase is changing by π, as shown in Fig. 3(a2). Besides the oscillation with a frequency near ω1, we observe a slow growth of n1. Meanwhile, the amplitude of oscillations are also increasing following increase of the population on d band.

 figure: Fig. 3

Fig. 3 Time evolution of nl measured from the experiments with initial modulation phase (a1) ϕ = −π/2 and (a2) ϕ = π/2. Time averaged fraction 〈nl〉 are also shown for (b1) ϕ = −π/2 and (b2) ϕ = π/2 respectively. n0 is shown with black dots comparing to the numerical simulation in solid lines. n1 and n1 are shown in average with red circles the corresponding numerical result is shown in dashed lines. Each point is averaged by three experiments and the error bars indicate the standard deviation.

Download Full Size | PDF

The influence of modulation phase is shown more clearly when taking time average in one modulation period T. In Fig. 3(b1) with ϕ1 = −π/2 the Floquet state shows a constant population in different momentum states by taking time average. While for ϕ1 = π/2, Fig. 3(b2) shows a Rabi oscillation between s-d bands with Rabi frequency EF/h¯, where EF is the gap between two Floquet bands. The observation can also be explained by Eq. (5). The phase of coupling terms would change with ϕ1, thus the eigenstate is superposition of |s〉 and |d〉 with a relative phase determined by ϕ1. When the modulation phase is changed, the loaded state is no longer an eignstate, and we can observe the Rabi oscillation.

5. Excitation of ground state via a double frequency modulation

The study of lattice modulation can be extended from single frequency to a more general form as described in Eq. (6). Specific to two-photon excitation between s-g band, the time dependent lattice is described as VL(t) = V0 +V1 cos(ω1t +ϕ1) +V2 cos(ω2t +ϕ2), which includes seven parameters V0,V1,V21212. During the experiments, we mainly focus on the influence of modulation frequencies and phases, while keeping other parameters constant.

5.1. Spectrum of two-photon excitation

The single frequency modulation shows that ϕ1 is related to relative phase between s and d band components of the Floquet state. However, with a nearly pure s band, the phase of single modulation can be neglected, only the phase difference between two modulations is important. Therefore, we chose ϕ1 = π while leaving ϕ2 variable to control the relative phase between two modulations.

For s-g band coupling through a two-photon absorption process, the sum of the two modulation frequencies is chosen as ω1 +ω2 = ωsg. This process is well described by Eq. (6), in which we have considered different pumping paths, as schematically shown in Fig. 4(a). There are two cases which would benefit the excitation process.

 figure: Fig. 4

Fig. 4 (a) Two special cases in detecting the transfer population spectrum. In case 1, absorption of photons with ω1(orange) or ω2(red) is resonant with d band. In case 2, two frequencies are equal. (b) For s-g coupling ω1 provides a two-photon process while ω2 = 2ω1 provides a one-photon process. Phases of two paths are controlled independently by modulation phases of ω1 and ω2.

Download Full Size | PDF

Case 1: Resonant two-photon process. When ω1 = ωsd or ω2 = ωsd, atoms are transferred from |s〉 to |g〉 with the assistance of d band as an intermediate band.

Case 2: Equal frequency two-photon process. When ω1 =ω2 =ωsg/2, two modulations with the same frequency can be added together, and the coupling strength of the process is doubled.

With resonance condition ω1 = ωsd in case 1, when two modulation amplitudes are chosen as V1Ωsd = V2Ωdg the transfer rate would show a maximum resonant peak. And we keep this modulation amplitudes while sweeping ω1. Modulation phase is chosen from numerical simulation to get a maximum transfer.

In the experiments, we sweep the frequency ω1 for different lattice depths V0 =5Er, 10Er and 14Er. Population on momentum states ±4h¯kL measured from the experiment are shown with rectangles in Fig. 5, comparing with the theoretical calculation shown in solid curves. Within the lattice depth we considered, g band is greatly concentrated on |±4h¯kL momentum states, thus n±2 can reflect transfer rate to g band.

 figure: Fig. 5

Fig. 5 Spectrum for the population on ±4h¯kL states with increasing of modulation frequency ω1. Population detected on ±4h¯kL after a double frequency modulation for (a) V0 = 5Er(black) with V1 = 1.4Er, ±V2 = 1.6Er t = 300μs, (b) V0 = 10Er(blue) with V1 = 2.8Er, V2 = 2.2Er and t = 200μs, (c) V0 = 14Er(red) with V1 = V2 = 2.5Er and t = 150μs are shown in rectangles with error bars. Solid lines are corresponding numerical simulation.

Download Full Size | PDF

Figure 5(a) shows the case of V0 = 5Er. For the lattice depth we have Ωsddg = 1.11, correspondingly the modulation amplitudes are chosen as V1 = 1.4Er, V2 = 1.6Er, and t = 300μs. The holding time t may be chosen shorter if the maximum of numerical simulation is reached at an earlier time. In the figure, two peaks appear at ω1 = ωsd and ω2 = ωsd which follows case 1 we have discussed. And there the central peak at frequency ω1 = ω2 following case 2 is much lower than two peaks for case 1.

Figure 5(b) shows the spectrum with V0 = 10Er. With the increasing of V0, the energy difference ωsd is getting closer to ωdg, and three peaks are overlapping. Comparing with V0 = 5Er the central peak is much higher, because the process of case 2 is also near resonance with d band.

For V0 = 14Er only one peak would be measured in the spectrum as shown in Fig. 5(c), which means case1 and case2 are fulfilled simultaneously. Under this condition, the coupling between s and g band is also greatly enhanced. Modulation amplitudes are V1 = V2 = 2.5Er, for the two modulation frequencies are the same at case1 and can’t be distinguished.

The experimentally detected peaks are governed by two cases, which are within the description of Eq. (6). Thus in the numerical simulation we can neglect higher order processes of emission and absorption of Floquet photons. The discrepancy between experimental result and the theoretical simulation is probably due to the influence of interaction and initial momentum distribution of the condensate. These effects would destroy coherency during the modulation, and the measured population of excited state would be lower than the maximum value in the theoretical simulation.

5.2. The role of modulation phases

When ω1 and ω2 are incommensurate, the influence of relative phase is not prominent because there is only one path for the pumping process and no states could interfere with each other. However, the effect of modulation phases would be more pronounced when two modulation frequencies are commensurate.

As shown in Fig. 4(b), when 2ω1 = ω2, relative phase of modulations can be shown by performing a one-photon process simultaneously with a two-photon process. We choose the frequency ω1 = ωsg/2 at central peak of V0 = 10Er lattice and the second modulation is performed with ω2 = ωsg. Modulation amplitudes are V1 = V2 = 2.5Er and t = 500μs. Similar to the process described by Eq. (6), in this problem we consider the interference between two states |Eg2h¯ω1 and |Egh¯ω2. During the two-photon process with ϕ1 = π, phase of state |Eg2h¯ω1 remains zero, while ϕ2 determine phase of state |Egh¯ω2 as ϕ2π. Relative phase of two modulations is defined as (ϕ2π)ω2ω1(ϕ1π)=ϕ2π, and the interference can be changed from constructive to destructive with different ϕ2.

Figure 6(a) depicts the population transferred to g band with the phase of one-photon process changing by 2π. Figure 6(b) shows how the depth of lattice is varying with time for four different modulation phases. With ϕ2 = π, the population transferred to |g〉 through one-photon process and through two-photon process are in phase and the transfer rate is enhanced. When ϕ2 is increasing, relative phase of two processes are deviating from zero, and the population on |g〉 would decrease. In the case of ϕ2 = 0 the two process are out of phase, and only few atoms can be transferred to g band. Further increasing of ϕ2 would increase the measured population, and when ϕ2 is changed by 2π the population on g band reaches the maximum value again.

The single atom Hamiltonian in Eq. (1) can well explain the excitation in our system. However, for the 500μs modulation, decoherence from mean-field interaction and the momentum distribution becomes significant, thus it is necessary to carry out a simulation based on the time-dependent Gross-Pitaevskii equation,

ih¯ψt=[h¯22m2x2+VL(x,t)+12mωx2x2+g|ψ|2]ψ,
where g is the parameter of interaction. In the simulation, we consider both the distribution of initial momentum and the mean-field interaction which would deduce the measured excitation rate. Dashed line in Fig. 6(a) gives the result of simulation which fits well with the experiment. Similarly, the relative phase can also be changed by ϕ1 of the two-photon process, which gives a period of π.

 figure: Fig. 6

Fig. 6 The excited population on g band shows the interference between two paths. (a) Population transferred to n±2 is shown in black dots with error bars. The dashed line shows theoretical simulation for comparison. (b1)-(b4) VL for different phases.

Download Full Size | PDF

The collaborative effect of different frequency components is studied in two aspects, demonstrate the effect of both modulation frequencies and phases of different frequency components. The usefulness of this study is not limited to double frequency modulation, within these two aspects, collaborate effect of more frequency components can also be well understood.

6. Application in LMT beam splitter

The double frequency modulation can also be applied to realize a LMT beam splitter by pumping atoms to higher momentum states, which is useful in experiments of atomic interferometry [24]. According to the resonant condition we have discussed, a preferred choice is V0 = 14Er with the modulation frequency ω1 = ωsd = ωdg. The other modulation frequency is chosen as ω2 = ωgi (with i the 6th excited Bloch band) to get a distribution at ±6h¯kL.

We begin with a condensate, the lattice is suddenly turn on, modulated with V1 = V2 = 2.5Er and the preferable phases are found numerically. A TOF image of experimental result is shown in Fig. 7(a). Figure 7(b) shows the population of atoms on momentum states |±6h¯kL where we have subtracted the thermal gas. Near 80% of the atoms are coherently transferred into ±6h¯kL momentum states within 160μs, and the maximum lattice depth needed is below V0 +V1 +V2 = 24Er. Comparing with a LMT beam splitter based on high order Bragg scattering [16], our method needs a much lower lattice depth and is easy in practice. Furthermore, the process is symmetric for both sides, and would not introduce phase shift between two separated atom clouds.

 figure: Fig. 7

Fig. 7 A LMT beam splitter with a separation of 12h¯kL. (a) TOF image of the LMT beam splitter. (b) Experimentally measured population of atom on momentum states |±6h¯kL are shown in black dots with error bars.

Download Full Size | PDF

A momentum splitting of 12h¯kL is not the limit of the amplitude modulation. More frequency components or another subsequent modulation can be introduced to reach a larger momentum splitting. For example, after the double frequency modulation, by preforming another single frequency modulation resonance with the energy difference between |±6h¯kL and |±8h¯kL momentum states, the atoms can be transferred to |±8h¯kL coherently.

7. Discussion and conclusion

Following Floquet-Bloch theory we have presented an experimental preparation of a Floquet state and study its property. The non-adiabatic loading method would take much less time and greatly reduce the heating problem. The loading method could also be extended to systems with a shaken lattice [7] or a combined modulation [6, 23], which provides a several millisecond longer lifetime for condensate in the experimental study on areas including the detection of Floquet topological states [17, 25, 26].

In conclusion, based on the study of single frequency modulation, we investigate double frequency modulation in detail. Experimental observations show that different modulation frequency components would influence each other in two ways. When two frequencies are resonant with subsequent excitation processes, the modulation can induce a resonant two photon process which can effectively transfer atoms to higher excited bands. With specific modulation frequencies, interference between a one-photon path and a two-photon path is observed, revealing influence of modulation phases when two frequencies are commensurate. The quantum interference can be used to enhance the excitation rate or destruct unwanted excitations. Using the technique of double frequency modulation, we also demonstrate an efficient way to realize a LMT beam splitter with low lattice depth. Our study provides a more flexible way to coherently control quantum states through an optical lattice.

Acknowledgments

This work is partially supported by the state Key Development Program for Basic Research of China No. 2011CB921501, NSFC (Grants No. 61475007, No. 11334001 and No. 91336103), RFDP (Grants No. 20120001110091).

References and links

1. J. Struck, C. Ölschläger, M. Weinberg, P. Hauke, J. Simonet, A. Eckardt, M. Lewenstein, K. Sengstock, and P. Windpassinger, “Tunable gauge potential for neutral and spinless particles in driven optical lattices,” Phys. Rev. Lett. 108, 225304 (2012). [CrossRef]   [PubMed]  

2. P. Hauke, O. Tieleman, A. Celi, C. Ölschläger, J. Simonet, J. Struck, M. Weinberg, P. Windpassinger, K. Sengstock, M. Lewenstein, and A. Eckardt, “Non-Abelian gauge fields and topological insulators in shaken optical lattices,” Phys. Rev. Lett. 109, 145301 (2012). [CrossRef]   [PubMed]  

3. J. Struck, M. Weinberg, C. Ölschläger, P. Windpassinger, J. Simonet, K. Sengstock, R. Höppner, P. Hauke, A. Eckardt, M. Lewenstein, and L. Mathey, “Engineering Ising-XY spin-models in a triangular lattice using tunable artificial gauge fields,” Nature Phys. 9, 738–743 (2013). [CrossRef]  

4. A. Alberti, V. V. Ivanov, G. M. Tino, and G. Ferrari, “Engineering the quantum transport of atomic wavefunctions over macroscopic distances,” Nature Phys. 5, 547–550 (2009). [CrossRef]  

5. A. Zenesini, H. Lignier, D. Ciampini, O. Morsch, and E. Arimonodo, “Coherent control of dressed matter waves,” Phys. Rev. Lett. 102, 100403 (2009). [CrossRef]   [PubMed]  

6. K. Hai, Y. Luo, G. Lu, and W. Hai, “Phase-controlled localization and directed transport in an optical bipartite lattice,” Opt. Express 22, 4277–4289 (2014). [CrossRef]   [PubMed]  

7. C. V. Parker, L. C. Ha, and C. Chin, “Direct observation of effective ferromagnetic domains of cold atoms in a shaken optical lattice,” Nat. Phys. 9, 769–774 (2013). [CrossRef]  

8. J. H. Denschlag, J. E. Simsarian, H. Häffner, C. McKenzie, A. Browaeys, D. Cho, K. Helmerson, S. L. Rolston, and W. D. Phillips, “A Bose-Einstein condensate in an optical lattice,” J. Phys. B 35, 3095–3110 (2002). [CrossRef]  

9. T. Stöferle, H. Moritz, C. Schori, M. Köhl, and T. Esslinger, “Transition from a strongly interacting 1D superfluid to a Mott insulator,” Phys. Rev. Lett. 92, 130403 (2004). [CrossRef]   [PubMed]  

10. W. K. Hensinger, H. Häffner, A. Browaeys, N. R. Heckenberg, K. Helmerson, C. McKenzie, G. J. Milburn, W. D. Phillips, S. L. Rolston, H. Rubinsztein-Dunlop, and B. Upcroft, “Dynamical tunnelling of ultracold atoms,” Nature 412, 52–55 (2001). [CrossRef]   [PubMed]  

11. D. A. Steck, W. H. Oskay, and M. G. Raizen, “Observation of chaos-assisted tunneling between islands of stability,” Science 293, 274–278 (2001). [CrossRef]   [PubMed]  

12. P. Cheiney, C. M. Fabre, F. Vermersch, G. L. Gattobigio, R. Mathevet, T. Lahaye, and D. Guéry-Odelin, “Matter-wave scattering on an amplitude-modulated optical lattice,” Phys. Rev. A 87, 013623 (2013). [CrossRef]  

13. A. Alberti, G. Ferrari, V. V. Ivanov, M. L. Chiofalo, and G. M. Tino, “Atomic wave packets in amplitude-modulated vertical optical lattices,” New J. Phys. 12, 065037 (2010). [CrossRef]  

14. N. Poli, F.-Y. Wang, M. G. Tarallo, A. Alberti, M. Prevedelli, and G. M. Tino, “Precision measurement of gravity with cold atoms in an optical lattice and comparison with a classical gravimeter,” Phys. Rev. Lett. 106, 038501 (2011). [CrossRef]   [PubMed]  

15. P. Cladé, S. Guellati-Khélifa, F. Nez, and F. Biraben, “Large momentum beam splitter using Bloch oscillations,” Phys. Rev. Lett. 102, 240402 (2009). [CrossRef]   [PubMed]  

16. M. Kozuma, L. Deng, E. W. Hagley, J. Wen, R. Lutwak, K. Helmerson, S. L. Rolston, and W. D. Phillips, “Coherent splitting of Bose-Einstein condensed atoms with optically induced Bragg diffraction,” Phys. Rev. Lett. 82, 871–875 (1999). [CrossRef]  

17. A. Gómez-León and G. Platero, “Floquet-Bloch theory and topology in periodically driven lattices,” Phys. Rev. Lett. 110, 200403 (2013). [CrossRef]   [PubMed]  

18. S.-I. Chu and D. A. Telnov, “Beyond the Floquet theorem: generalized Floquet formalisms and quasienergy methods for atomic and molecular multiphoton processes in intense laser fields,” Phys. Rep. 390, 1–131 (2004). [CrossRef]  

19. O. Morsch, J. H. Müller, M. Cristiani, D. Ciampini, and E. Arimondo, “Bloch oscillations and mean-field effects of Bose-Einstein condensates in 1D optical lattices,” Phys. Rev. Lett. 87, 140402 (2001). [CrossRef]   [PubMed]  

20. D. Choi and Q. Niu, “Bose-Einstein condensates in an optical lattice,” Phys. Rev. Lett. 82, 2022 (1999). [CrossRef]  

21. X. X. Liu, X. J. Zhou, W. Xiong, T. Vogt, and X. Z. Chen, “Rapid nonadiabatic loading in an optical lattice,” Phys. Rev. A 83, 063402 (2011). [CrossRef]  

22. Y. Y. Zhai, X. G. Yue, Y. J. Wu, X. Z. Chen, P. Zhang, and X. J. Zhou, “Effective preparation and collisional decay of atomic condensates in excited bands of an optical lattice,” Phys. Rev. A 87, 063638 (2013). [CrossRef]  

23. T. Salger, S. Kling, S. Denisov, A. V. Ponomarev, P. Hänggi, and M. Weitz, “Tuning the mobility of a driven Bose-Einstein condensate via diabatic Floquet bands,” Phys. Rev. Lett. 110, 135302 (2013). [CrossRef]   [PubMed]  

24. D. Döring, J. E. Debs, N. P. Robins, C. Figl, P. A. Altin, and J. D. Close, “Ramsey interferometry with an atom laser,” Opt. Express 17, 20661–20668 (2009). [CrossRef]   [PubMed]  

25. W. Zheng and H. Zhai, “Floquet topological states in shaking optical lattices,” Phys. Rev. A 89, 061603 (2014). [CrossRef]  

26. S.-L. Zhang and Q. Zhou, “Shaping topological properties of the band structures in a shaken optical lattice,” Phys. Rev. A 90, 051601 (2014). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (7)

Fig. 1
Fig. 1 A sketch of lattice depth modulation in our system. The depth of lattice potential V0 cos2(kLx) is driven by a polychromatic modulation ∑iVi cos(ωit +ϕi).
Fig. 2
Fig. 2 Left side is the calculated Floquet spectra of a single frequency driven system, with parameters V0 = 5Er, V1 = 0.5Er, h ¯ ω 1 = 5 E r. In the figure the first seven bands are presented. The heavy lines depict states maximally overlapping with the s(blue), p (green) and d (red) Bloch bands respectively. Right side shows the details of two Floquet bands most overlapping with s and d bands. The two bands are separated by a band gap EF at q = 0.
Fig. 3
Fig. 3 Time evolution of nl measured from the experiments with initial modulation phase (a1) ϕ = −π/2 and (a2) ϕ = π/2. Time averaged fraction 〈nl〉 are also shown for (b1) ϕ = −π/2 and (b2) ϕ = π/2 respectively. n0 is shown with black dots comparing to the numerical simulation in solid lines. n1 and n1 are shown in average with red circles the corresponding numerical result is shown in dashed lines. Each point is averaged by three experiments and the error bars indicate the standard deviation.
Fig. 4
Fig. 4 (a) Two special cases in detecting the transfer population spectrum. In case 1, absorption of photons with ω1(orange) or ω2(red) is resonant with d band. In case 2, two frequencies are equal. (b) For s-g coupling ω1 provides a two-photon process while ω2 = 2ω1 provides a one-photon process. Phases of two paths are controlled independently by modulation phases of ω1 and ω2.
Fig. 5
Fig. 5 Spectrum for the population on ± 4 h ¯ k L states with increasing of modulation frequency ω1. Population detected on ± 4 h ¯ k L after a double frequency modulation for (a) V0 = 5Er(black) with V1 = 1.4Er, ±V2 = 1.6Er t = 300μs, (b) V0 = 10Er(blue) with V1 = 2.8Er, V2 = 2.2Er and t = 200μs, (c) V0 = 14Er(red) with V1 = V2 = 2.5Er and t = 150μs are shown in rectangles with error bars. Solid lines are corresponding numerical simulation.
Fig. 6
Fig. 6 The excited population on g band shows the interference between two paths. (a) Population transferred to n±2 is shown in black dots with error bars. The dashed line shows theoretical simulation for comparison. (b1)-(b4) VL for different phases.
Fig. 7
Fig. 7 A LMT beam splitter with a separation of 12 h ¯ k L. (a) TOF image of the LMT beam splitter. (b) Experimentally measured population of atom on momentum states | ± 6 h ¯ k L are shown in black dots with error bars.

Equations (7)

Equations on this page are rendered with MathJax. Learn more.

H ( t ) = p x 2 2 M + V 0 cos 2 ( k L x ) + i V i cos ( ω i t + ϕ i ) cos 2 ( k L x ) .
| ψ q , α ( t + T ) = U ^ ( T ) | ψ q , α ( t ) = e i ε q , α T | ψ q , α ( t ) ,
( H ( t ) i t ) | u q , α = H 0 | u q , α = ε q , α | u q , α .
A l m , l m = 1 T 0 T | v l m H 0 v l m | d t .
H R = ( H R V 1 e i ϕ 1 Ω α β * V 1 e i ϕ 1 Ω α β * E β h ¯ ω ) ,
H s g = ( E s e i ϕ 1 V 1 Ω s d e i ϕ 2 V 2 Ω s d 0 0 0 e i ϕ 1 V 1 Ω s d * E d h ¯ ω 1 0 e i ϕ 2 V 2 Ω d g e i ϕ 1 V 1 Ω d g 0 e i ϕ 2 V 2 Ω s d * 0 E d h ¯ ω 2 e i ϕ 1 V 1 Ω d g 0 e i ϕ 2 V 2 Ω d g 0 e i ϕ 2 V 2 Ω d g * e i ϕ 1 V 1 Ω d g * E g h ¯ ( ω 1 + ω 2 ) 0 0 0 e i ϕ 1 V 1 Ω d g * 0 0 E g 2 h ¯ ω 1 0 0 0 e i ϕ 2 V 2 Ω d g * 0 0 E g 2 h ¯ ω 2 ) ,
i h ¯ ψ t = [ h ¯ 2 2 m 2 x 2 + V L ( x , t ) + 1 2 m ω x 2 x 2 + g | ψ | 2 ] ψ ,
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.