arxiv:1506.00558v1 [cond-mat.quant-gas] 1 jun 2015dalibard/publications/2015-1506.00558v1.pdf · 2...

9
Dynamic optical lattices of sub-wavelength spacing for ultracold atoms Sylvain Nascimbene, 1, Nathan Goldman, 1, 2 Nigel Cooper, 3 and Jean Dalibard 1 1 Laboratoire Kastler Brossel, Coll` ege de France, ENS-PSL Research University, CNRS, UPMC-Sorbonne Universit´ es, 11 place Marcelin Berthelot, 75005 Paris, France 2 CENOLI, Faculte des Sciences, Universit´ e Libre de Bruxelles (U.L.B.), B-1050 Brussels, Belgium 3 T.C.M. Group, Cavendish Laboratory, J.J. Thomson Avenue, Cambridge CB3 0HE, United Kingdom (Dated: June 2, 2015) We propose a scheme to realize lattice potentials of sub-wavelength spacing for ultracold atoms. It is based on spin-dependent optical lattices with a time-periodic modulation. We show that the atomic motion is well described by the combined action of an eective, time-independent, lattice of small spacing, together with a micro-motion associated with the time-modulation. A numerical simulation shows that an atomic gas can be adiabatically loaded into the eective lattice ground state, for timescales comparable to the ones required for adiabatic loading of standard optical lattices. We generalize our scheme to a two-dimensional geometry, leading to Bloch bands with non-zero Chern numbers. The realization of lattices of sub-wavelength spacing allows for the enhancement of energy scales, which could facilitate the achievment of strongly-correlated (topological) states. Optical lattices have allowed experiments on ultracold atomic gases to investigate a large range of lattice models of quantum many-body physics [1]. Their development led to the realization of strongly-correlated states of mat- ter, such as bosonic and fermionic Mott insulators, and low-dimensional gases [2]. In its simplest form, an optical lattice consists of the optical dipole potential associated with a standing wave of retro-reflected laser light. It can be described as a periodic potential V (x)= U 0 cos 2 (kx), of spatial period d = λ/2, where λ is the laser wavelength and k =2/λ. More complex optical lattices, such as su- perlattices [3, 4] or two-dimensional honeycomb lattices [5, 6], can be generated with suitable laser configurations. The recoil energy E r = h 2 /(8md 2 ), where h is Planck’s constant and m is the atom mass, sets the natural energy scale for elementary processes, such as atom tunneling between neighboring lattice sites, as well as the temper- ature range T . E r /k B 100 nK, typically required for quantum degeneracy. For a large class of models, the physical behavior is dictated by processes associated with even much smaller energies, such as super-exchange or magnetic dipole in- teractions [1]. The associated temperature scales remain out of reach in current experiments. In order to circum- vent this limitation, it is desirable to find novel schemes for generating optical lattices with spacing d eλ, in order to enhance the associated energy scale E er = h 2 /(8md 2 e) [7]. Schemes have been proposed to gener- ate lattices of sub-wavelength spacing, based on multi- photon optical transitions [8] or on adiabatic dressing of state-dependent optical lattices [7]; the realization of lat- tices with spacing d e= λ/4 was reported in Ref. [9]. An interesting alternative would be to trap atomic gases in the electromagnetic fields of nano-structured condensed- matter systems [10–12]. In this letter, we propose a novel scheme leading to lattices of spacing d e= d/N , N being an arbitrary inte- ger, based on spin-dependent lattices with time-periodic 0 1 0 t<T/4 V (x, t) 0 1 T/4 t<T/2 0 1 T/2 t< 3T/4 0 1 3T/4 t<T 0 1 2 0 0.25 x[d] V e(x) FIG. 1. Stroboscopic scheme for engineering short-spacing lattices, illustrated on the case N = 4. We make use of a periodic potential V (x, t) of spatial period d, that is shifted of the distance d/N after every time step of duration T /N (blue curves). The eective potential V e(x) (red curve), resulting from time averaging, exhibits a spatial period d e= d/N . modulation. In the regime of large modulation frequency [13–16], the atom dynamics is governed by an eec- tive static periodic potential of spacing d e, with an additional micro-motion. This description is confirmed by a numerical simulation, which shows the possibility to load adiabatically the ground state of the eective lattice and to perform Bloch oscillations. We discuss the extension of the scheme to two-dimensional lattices with non-trivial topology. Lattices with artificial mag- netic fields, generally leading to topological bands, were recently realized in experiments, with standard lattice spacing [17]. For those systems, increasing the energy scale using short-spacing lattices could prove important arXiv:1506.00558v1 [cond-mat.quant-gas] 1 Jun 2015

Upload: vulien

Post on 16-Jun-2019

217 views

Category:

Documents


0 download

TRANSCRIPT

Dynamic optical lattices of sub-wavelength spacing for ultracold atoms

Sylvain Nascimbene,1, ⇤ Nathan Goldman,1, 2 Nigel Cooper,3 and Jean Dalibard1

1Laboratoire Kastler Brossel, College de France, ENS-PSL Research University,CNRS, UPMC-Sorbonne Universites, 11 place Marcelin Berthelot, 75005 Paris, France

2CENOLI, Faculte des Sciences, Universite Libre de Bruxelles (U.L.B.), B-1050 Brussels, Belgium3T.C.M. Group, Cavendish Laboratory, J.J. Thomson Avenue, Cambridge CB3 0HE, United Kingdom

(Dated: June 2, 2015)

We propose a scheme to realize lattice potentials of sub-wavelength spacing for ultracold atoms.It is based on spin-dependent optical lattices with a time-periodic modulation. We show that theatomic motion is well described by the combined action of an e↵ective, time-independent, latticeof small spacing, together with a micro-motion associated with the time-modulation. A numericalsimulation shows that an atomic gas can be adiabatically loaded into the e↵ective lattice groundstate, for timescales comparable to the ones required for adiabatic loading of standard optical lattices.We generalize our scheme to a two-dimensional geometry, leading to Bloch bands with non-zeroChern numbers. The realization of lattices of sub-wavelength spacing allows for the enhancementof energy scales, which could facilitate the achievment of strongly-correlated (topological) states.

Optical lattices have allowed experiments on ultracoldatomic gases to investigate a large range of lattice modelsof quantum many-body physics [1]. Their developmentled to the realization of strongly-correlated states of mat-ter, such as bosonic and fermionic Mott insulators, andlow-dimensional gases [2]. In its simplest form, an opticallattice consists of the optical dipole potential associatedwith a standing wave of retro-reflected laser light. It canbe described as a periodic potential V (x) = U

0

cos2(kx),of spatial period d = �/2, where � is the laser wavelengthand k = 2⇡/�. More complex optical lattices, such as su-perlattices [3, 4] or two-dimensional honeycomb lattices[5, 6], can be generated with suitable laser configurations.The recoil energy E

r

= h2/(8md2), where h is Planck’sconstant and m is the atom mass, sets the natural energyscale for elementary processes, such as atom tunnelingbetween neighboring lattice sites, as well as the temper-ature range T . E

r

/kB

⇠ 100 nK, typically required forquantum degeneracy.

For a large class of models, the physical behavior isdictated by processes associated with even much smallerenergies, such as super-exchange or magnetic dipole in-teractions [1]. The associated temperature scales remainout of reach in current experiments. In order to circum-vent this limitation, it is desirable to find novel schemesfor generating optical lattices with spacing d

e↵

⌧ �,in order to enhance the associated energy scale Ee↵

r

=h2/(8md2

e↵

) [7]. Schemes have been proposed to gener-ate lattices of sub-wavelength spacing, based on multi-photon optical transitions [8] or on adiabatic dressing ofstate-dependent optical lattices [7]; the realization of lat-tices with spacing d

e↵

= �/4 was reported in Ref. [9]. Aninteresting alternative would be to trap atomic gases inthe electromagnetic fields of nano-structured condensed-matter systems [10–12].

In this letter, we propose a novel scheme leading tolattices of spacing d

e↵

= d/N , N being an arbitrary inte-ger, based on spin-dependent lattices with time-periodic

0

1

0 t < T/4

V(x,t)

0

1

T/4 t < T/2

0

1

T/2 t < 3T/4

0

1

3T/4 t < T

0 1 20

0.25

x[d]

Ve↵(x)

FIG. 1. Stroboscopic scheme for engineering short-spacinglattices, illustrated on the case N = 4. We make use of aperiodic potential V (x, t) of spatial period d, that is shifted ofthe distance d/N after every time step of duration T/N (bluecurves). The e↵ective potential Ve↵(x) (red curve), resultingfrom time averaging, exhibits a spatial period de↵ = d/N .

modulation. In the regime of large modulation frequency[13–16], the atom dynamics is governed by an e↵ec-tive static periodic potential of spacing d

e↵

, with anadditional micro-motion. This description is confirmedby a numerical simulation, which shows the possibilityto load adiabatically the ground state of the e↵ectivelattice and to perform Bloch oscillations. We discussthe extension of the scheme to two-dimensional latticeswith non-trivial topology. Lattices with artificial mag-netic fields, generally leading to topological bands, wererecently realized in experiments, with standard latticespacing [17]. For those systems, increasing the energyscale using short-spacing lattices could prove important

arX

iv:1

506.

0055

8v1

[con

d-m

at.q

uant

-gas

] 1

Jun

2015

2

for creating strongly-correlated states such as fractionalChern insulators [18, 19].

A basic scheme of our method is pictured in Fig. 1.Consider a periodic potential V (x) of period d, whichis abruptly shifted by the distance d/N at stroboscopictimes tn = (n/N)T , n 2 Z, leading to a time-periodicpotential V (x, t) of period T . Provided that T is muchsmaller than typical timescales of atomic motion, theatoms experience an e↵ective time-averaged potentialVe↵

(x) =RV (x, t)dt/T . A simple calculation shows that

Ve↵

(x) is given by the sum of all harmonics of the po-tential V (x), whose orders are multiples of N [20]. Thee↵ective potential V

e↵

(x) is thus spatially periodic, ofspatial period d

e↵

= d/N .Conventional optical lattices present a spatial modu-

lation proportional to the intensity pattern of interferinglight waves, which exhibit spatial frequencies of at mosttwice the light momentum k. Thus, applying the stro-boscopic scheme in Fig. 1 to these potentials could notlead to e↵ective lattices of period d

e↵

< �/2. This re-striction does not apply to spinful particles subjectedto spin-dependent optical lattices. As an illustration,consider a spin-1/2 particle evolving in the potentialV (x) = V

L

cos(2kx)�z + VB

�x, where �u (u = x, y, z)are the Pauli matrices. In a dressed state picture, theatom may follow adiabatically the state of lowest energyV�(x) = �

pV 2

L

cos2(2kx) + V 2

B

. As this potential ex-hibits harmonics of the spatial frequency 2k of all orders,the lattice spacings achievable by applying the strobo-scopic scheme to V�(x) can be made arbitrarily small.

We describe in the following a modified, more practical,version of this scheme, which consists of a spin-dependentoptical lattice with smooth temporal variations, given by

V (x, t) = VL

cos(2kx� ⌦t)�z + VB

cos(N⌦t)�x. (1)

This potential satisfies V (x + d/N, t + T/N) = V (x, t),with d = ⇡/k, thus, it can be viewed as a continuousversion of the stroboscopic scheme. Understanding thephysical e↵ects of the potential (1) falls within the de-scription of time-periodic Hamiltonian systems [13–16].Following Ref. [15], we describe the dynamics of an atombetween the times t

i

and tf

as

U(ti

! tf

) = e�iK(tf )e�i~ (tf�ti)Heff eiK(ti), (2)

where we introduce a time-independent, e↵ective Hamil-tonian H

e↵

and a time-periodic kick operator K(t). Thethree operators in (2) describe, from right to left, therole of the initial phase of the Hamiltonian at time t

i

, theevolution from t

i

to tf

according to a stationary Hamil-tonian, and the micro-motion related to the final phaseof the Hamiltonian at time t

f

.The expressions for the e↵ective Hamiltonian

He↵

and kick operator K(t) can be calculatedthrough a perturbative expansion in powers of1/⌦, see Refs. [14, 15]. To lowest-order, this yields

0

5

10(a)

�1 0 1

0

50

100

150

q [k]

![E

r/~]

0

5

10

![E

e↵

r/~]

(b)

�4 �2 0 2 4

0

50

100

150

q [k]

FIG. 2. Band structure of a dynamic optical lattice of spac-ing de↵ = d/4, corresponding to the parameters N = 4, andVL = VB = ~⌦ = 200Ee↵

r . In (a), we make use of the spa-tial and temporal translational symmetries T

x

, Tt

and labelthe eigenstates by their quasi-momentum �k q < k andquasi-energy �~⌦/2 ~! < ~⌦/2. The Bloch-Floquet bandscan be unfolded using the additional symmetry T ⇤, leadingto the band structure in (b), indexed by the modified quasi-momentum �4k q < 4k. The unfolding of the band struc-ture can be followed from the di↵erent coloring of successivebands.

He↵

=p2

2m+ V

e↵

(x), (3)

Ve↵

(x) =Ue↵

2cos(2Nkx)�x, U

e↵

=2V

B

N !

✓VL

~⌦

◆N

, (4)

K(t) =�V

L

~⌦ sin(2kx� ⌦t)�z +VB

N~⌦ sin(N⌦t)�x. (5)

The expression (4) is derived under the assumption thatN is an even integer. In the Supplementary material weshow that the perturbative expansion can be resumed,with respect to either the variable V

L

/(~⌦) or VB

/(~⌦)[20]. The e↵ective potential (4) describes a periodicpotential of depth U

e↵

and spatial period de↵

= d/N .In order to test the validity of the e↵ective Hamil-

tonian, we performed a numerical study of the time-periodic Hamiltonian using the Floquet formalism. Sincethe Hamiltonian H is invariant under the symmetriesTx : x ! x+d and Tt : t ! t+T , we look for eigenstateswritten as Bloch-Floquet wave functions q,!(x, t) =ei(qx�!t)uq,!(x, t), where uq,!(x, t) is d-periodic in x andT -periodic in t [21, 22]. Eigenstates are labelled bytheir quasi-momentum �k < q k and quasi-energy0 ~! < ~⌦. An example of the band structure calcu-lated numerically for N = 4 is plotted in Fig. 2a. Theband structure exhibits gap openings once every fourbands, at the momenta Nkp, where p 2 Z⇤, as expectedfor a lattice of spacing d/N .The band structure can be unfolded, making use of the

additional symmetry T ⇤ : x ! x + d/N, t ! t + T/N .

3

�4 �3 �2 �1 0 1 2 3 4

x [d]

n(x)[a.u.]

(a)

05

10

x [d]0

1

2

t [⌧B

]

n(x,t)

(b)

0 0.5 1

0

0.5

1

t [⌧B

]

xcm(t)[W

e↵/F]

(c)

FIG. 3. (a) Atomic density of a wave-packet loaded into adynamic lattice of spacing de↵ = d/4. We start from a gaus-sian wave-packet, spin-polarized along x, of wave function (x, t = 0) = exp[�x2/(2�2)], with � ' 1.4 d (red dashedline). The lattice depth VL is slowly ramped up for a du-ration tramp = 20~/Ee↵

r from VL = 0 to VL = V 0L , and lat-

tice parameters N = 4, V 0L = VB = ~⌦ = 200Ee↵

r . Theatom density after loading is spatially modulated, with a pe-riod d/4 (blue line). (b) Evolution of the density distributionduring Bloch oscillations, calculated for the dynamic latticeparameters of (a), and for a force F = We↵/(8 de↵), whereWe↵ ' 0.06Ee↵

r is the expected bandwidth of the lowest bandfor Ue↵ = 10.9Ee↵

r . (c) Evolution of the center-of-mass posi-tion during Bloch oscillations, calculated for a standard op-tical lattice of depth Ue↵ = 10.9Ee↵

r (red dashed line), andfor the dynamic optical lattice (blue line). The time and spa-tial coordinates are plotted in units of the ideal Bloch period⌧B

= 2N~k/F and amplitude We↵/F .

As explained in the Supplementary Material, eigenstatesassociated with the symmetries Tx, Tt and T ⇤ can bewritten as q,!(x, t) = ei(qx�!t)vq,!(x, t), where vq,!(x, t)is d/N -periodic in x and 2⇡-periodic in (kx�⌦t) [20]. Weshow the band structure calculated within this formalismin Fig. 2b, which is very close to that expected for alattice of spacing d/N and depth U

e↵

' 10.9Ee↵

r

[23] .

The practical relevance of the short-spacing lattice de-scribed above is based on the ability to load atoms intothe ground state of the e↵ective potential (4). The anal-ysis of this loading protocol requires special care, as thee↵ective-Hamiltonian approach inherent to Eq. (2) as-sumes a constant lattice depth [15]. In fact, we find thatthe concept of the e↵ective Hamiltonian can be modifiedso as to describe the time-evolution under a ramp of themoving-lattice depth V

L

, see Ref. [20]. We simulate thelattice loading from a numerical calculation of the fulldynamics of an atomic wave packet under the action of

�0.5 0 0.5

0

5

10

15

q [2Nk]

!�

qv/~[E

e↵

r/~]

v = 0

�0.5 0 0.5

�5

0

5

10

q [2Nk]

v = vlatt

FIG. 4. Band structure corresponding to the dynamic latticefor the parameters N = 2, VL = VB = ~⌦ = 10Ee↵

r . Thepanels correspond to di↵erent frames of reference, of veloc-ity v = 0 (left) and v = vlatt = ⌦/(2k) (right). The bluepoints correspond to the band structure of an optical latticeof spacing de↵ = d/N and depth Ue↵ ' 2Ee↵

r , at rest in thelaboratory frame. The red dots correspond to the band struc-ture of an optical lattice of spacing d and depth U ' 74Er

(' 9Ue↵), at rest in the frame of velocity v = vlatt [24].

the potential (1). Starting from a gaussian wave packet,spin-polarized along x, we solve the Schrodinger equa-tion, discretized in space and time, with a lattice depthVL

slowly ramped up for a duration tramp

. As shownin Fig. 3a, a ramp duration t

ramp

= 20~/Ee↵

r

leads to astate with strong spatial modulations of spacing d/N , asexpected for a wavepacket prepared in the lowest bandof the e↵ective lattice (4). The calculated populationin the e↵ective lowest band is 93%, close to the valueexpected with standard optical lattices for such a rampduration. In the Supplementary Information we analyzethe momentum distribution, which corresponds to theone expected for the ground state of the e↵ective lattice,slightly modified by the micro-motion [20].

The system description as an e↵ective d/N lattice isalso supported by a numerical simulation of Bloch os-cillations. We calculate the action of a linear potential�Fx applied to the state obtained after the lattice load-ing. As shown in Fig. 3b,c, the wave packet undergoesBloch oscillations, revealed as real-space oscillations ofits center of mass. Both the amplitude and period of thisoscillation agree well with those expected for an e↵ectivelattice of period d/N and depth U

e↵

inferred from bandstructure calculations.

The potential V (x, t) written in (1) corresponds tothe sum of a time-modulated magnetic field and a spin-dependent optical lattice moving at the velocity v

latt

=⌦/(2k). In the above discussion we considered the ef-fect of this potential as an e↵ective static optical lattice.

4

�x

�x

�y

�y

�z

�z

qx

qy

a

e1

e2

e3

b

�x

1

�y

i�z

1

⇡/2

FIG. 5. (a) Momentum-space representation of the e↵ectivecouplings in Eq. 6, illustrated as arrows of length 2Nk, ori-ented along the unit vectors±e

i

(i = 1, 2, 3), and proportionalto Pauli matrices. Quantum states are represented in the ba-sis {|+

z

i (filled dots), |�z

i (circles)}. (b) Phase accumulatedaround a triangular subcell of the k-space lattice. Due tothe internal-state degree of freedom, the unit cell of the lat-tice is formed by four triangular subcells. The same phaseof � = ⇡/2 is found to be accumulated around all subcells,indicating that the lowest energy band is associated with anon-trivial Chern number ⌫Ch = 1 [25].

An alternative view is obtained in the frame of referencemoving at the velocity v = v

latt

, where the potentialV (x0 = x � vt, t) consists of the sum of a modulatedmagnetic field and a static lattice V

L

cos(2kx0)�z, withVL

⇠ ~⌦ � Ue↵

. Both points of view can be reconciledby a proper interpretation of the band structure, as illus-trated for the case N = 2 in Fig. 4. Among the eigenener-gies (q,!) calculated numerically in the laboratory framev = 0, we identify the Bloch bands corresponding to astatic e↵ective lattice of spacing d

e↵

. The eigenenergies(q,!0) corresponding to a frame of reference moving at avelocity v can be deduced from those in the laboratoryframe using the relation !0 = !�qv/~. In the frame mov-ing at v = v

latt

, we observe Bloch bands correspondingto a very deep static optical lattice of period d.

We now consider a 2D extension of our scheme. Thetime-dependent part of the Hamiltonian is taken as

V (r, t) = VL

cos(2ke1

· r� ⌦1

t)�x + VB

cos(N⌦1

t)�y

+ VL

cos(2ke2

· r� ⌦2

t)�y + VB

cos(N⌦2

t)�z

+ VL

cos(2ke3

· r� ⌦3

t)�z + VB

cos(N⌦3

t)�x,

where the unit vectors e1,2,3 have directions as repre-

sented in Fig. 5. For a suitable choice of the frequencies⌦

1,2,3 [26], each line of the equation above can be treatedindividually, which results in an e↵ective potential of theform [27]

Ve↵

(r) ' Ue↵

2[ cos(2Nke

1

· r)�x + cos(2Nke2

· r)�y

+ cos(2Nke3

· r)�z], (6)

where N is taken to be an even integer. These couplings

are illustrated in quasi-momentum space in Fig. 5a. Fol-lowing Ref. [25], the topological Chern number associ-ated with the lowest energy band can be readily obtainedfrom these couplings. Indeed, the Chern number mea-sures the flux of the Berry curvature ⌦(q) over the entire(momentum-space) unit cell:

⌫Ch

=1

2⇡

Z

unit cell

⌦(q)d2q, (7)

which can be directly evaluated by calculating the phasesaccumulated by a state as it performs a loop aroundthe triangular subcells [25]. For the e↵ective lattice de-scribed by eq. (6), each unit cell is constituted of fourtriangular subcells, and we find an accumulated phase of⇡/2 within each of them (see Fig. 5b). In this configu-ration, the Chern number of the lowest band is given by⌫Ch

= (1/2⇡) ⇥ 4 ⇥ (⇡/2) = 1. Generally the reasoningabove is valid only in the weak-binding regime; however,for the coupling (6), ⌫

Ch

is unchanged for all values ofUe↵

. Note that the size of the unit cell in real spacescales as 1/N2; we thus expect the flux density to be in-creased by a factor of N2 compared to standard opticallattices.In conclusion, we introduced a novel scheme to en-

gineer spatially periodic atom traps of sub-wavelengthspacing, based on the application of spin-dependent op-tical lattices. The scheme could be implemented withalkali atoms, and more favorably with Lanthanide atomsfor which one expects lower Rayleigh scattering e↵ects[28, 29]. We mention that, since short-spacing latticesare associated with typically higher energy scales com-pared to usual optical lattices, their realization requireslarge laser intensities. A natural extension of this workwould be to include interactions between atoms in thee↵ective lattice description, and to understand whethermicro-motion plays a significant role in scattering proper-ties [30–32]. This aspect will play a central role for inves-tigating quantum many-body physics with short-spacinglattices.The authors are pleased to acknowledge Fabrice Ger-

bier and Jerome Beugnon for valuable discussions.This work is supported by IFRAF, ANR (ANR-12-BLANAGAFON), ERC (Synergy UQUAM), the RoyalSociety of London and the EPSRC. N.G. is financed bythe FRS-FNRS Belgium and by the BSPO under the PAIproject P7/18 DYGEST.

[email protected][1] M. Lewenstein, A. Sanpera, V. Ahufinger, B. Damski,

A. Sen, and U. Sen, Adv. Phys. 56, 243 (2007).[2] I. Bloch, J. Dalibard, and W. Zwerger, Rev. Mod. Phys.

80, 885 (2008).[3] J. Sebby-Strabley, M. Anderlini, P. Jessen, and J. Porto,

Phys. Rev. A 73, 033605 (2006).

5

[4] S. Folling, S. Trotzky, P. Cheinet, M. Feld, R. Saers,A. Widera, T. Muller, and I. Bloch, Nature 448, 1029(2007).

[5] P. Soltan-Panahi, J. Struck, P. Hauke, A. Bick,W. Plenkers, G. Meineke, C. Becker, P. Windpassinger,M. Lewenstein, and K. Sengstock, Nature Phys. 7, 434(2011).

[6] L. Tarruell, D. Greif, T. Uehlinger, G. Jotzu, andT. Esslinger, Nature 483, 302 (2012).

[7] W. Yi, A. Daley, G. Pupillo, and P. Zoller, New J. Phys.10, 073015 (2008).

[8] B. Dubetsky and P. Berman, Phys. Rev. A 66, 045402(2002).

[9] G. Ritt, C. Geckeler, T. Salger, G. Cennini, andM. Weitz, Phys. Rev. A 74, 063622 (2006).

[10] M. Gullans, T. Tiecke, D. Chang, J. Feist, J. Thompson,J. I. Cirac, P. Zoller, and M. D. Lukin, Phys. Rev. Lett.109, 235309 (2012).

[11] J. Thompson, T. Tiecke, N. de Leon, J. Feist, A. Akimov,M. Gullans, A. Zibrov, V. Vuletic, and M. Lukin, Science340, 1202 (2013).

[12] O. Romero-Isart, C. Navau, A. Sanchez, P. Zoller, andJ. I. Cirac, Phys. Rev. Lett. 111, 145304 (2013).

[13] P. Avan, C. Cohen-Tannoudji, J. Dupont-Roc, andC. Fabre, J. Phys. 37, 993 (1976).

[14] S. Rahav, I. Gilary, and S. Fishman, Phys. Rev. A 68,013820 (2003).

[15] N. Goldman and J. Dalibard, Phys. Rev. X 4, 031027(2014).

[16] M. Bukov, L. D’Alessio, and A. Polkovnikov,arXiv:1407.4803 (2014).

[17] N. Goldman, G. Juzeliunas, P. Ohberg, and I. B. Spiel-man, Rep. Prog. Phys 77, 126401 (2014).

[18] E. J. Bergholtz and Z. Liu, Int. J. Mod. Phys. B 27(2013).

[19] S. A. Parameswaran, R. Roy, and S. L. Sondhi, C. R.Phys. 14, 816 (2013).

[20] See supplementary material for a discussion on the stro-boscopic method, the Bloch-Floquet calculation, the re-summation of the perturbative expansion of He↵ , micro-motion e↵ects in momentum space, and the descriptionof the lattice loading.

[21] M. Holthaus, Z. Phys. B Cond. Mat. 89, 251 (1992).[22] M. Grifoni and P. Hanggi, Phys. Rep. 304, 229 (1998).[23] This depth value slightly di↵ers from the perturbative

result Ue↵ ' 16.7Ee↵r , from eq. (4), since V

L,B

6⌧ ~⌦. Wechecked numerically that the di↵erence can be accountedfor by higher-order terms.

[24] The modulated magnetic field VB

cos(N⌦t)�x

renormal-izes the depth U of the optical lattice according toU = 2J0[VB

/(N⌦)]VL

, as observed in the calculatedband structures.

[25] N. R. Cooper and R. Moessner, Phys. Rev. Lett. 109,215302 (2012).

[26] The frequencies ⌦1,2,3 should not be chosen too close toeach other, and their ratios should not approach simplefractions.

[27] N. R. Cooper, Phys. Rev. Lett. 106, 175301 (2011).[28] S. Nascimbene, J. Phys. B: At. Mol. Opt. Phys. 46,

134005 (2013).[29] X. Cui, B. Lian, T.-L. Ho, B. L. Lev, and H. Zhai, Phys.

Rev. A 88, 011601 (2013).[30] S. Choudhury and E. J. Mueller, Phys. Rev. A 90, 013621

(2014).

[31] T. Bilitewski and N. R. Cooper, Phys. Rev. A 91, 033601(2015).

[32] A. Eckardt and E. Anisimovas, arXiv:1502.06477 (2015).

Supplementary Material for

Dynamic optical lattices of sub-wavelength spacing for ultracold atoms

Sylvain Nascimbene,1, ⇤ Nathan Goldman,1, 2 Nigel Cooper,3 and Jean Dalibard1

1Laboratoire Kastler Brossel, College de France, ENS-PSL Research University,CNRS, UPMC-Sorbonne Universites, 11 place Marcelin Berthelot, 75005 Paris, France

2CENOLI, Faculte des Sciences, Universite Libre de Bruxelles (U.L.B.), B-1050 Brussels, Belgium3T.C.M. Group, Cavendish Laboratory, J.J. Thomson Avenue, Cambridge CB3 0HE, United Kingdom

(Dated: June 1, 2015)

S.I. EFFECTIVE POTENTIAL CREATED BYTHE STROBOSCOPIC SCHEME

In this Section, we present the calculation of the ef-fective lattice potential produced via the stroboscopicscheme illustrated in Fig. 1 (main text). We start froma periodic potential V (x) of period d, which can be de-composed in Fourier series as

V (x) =X

p2ZVp e

i2⇡px/d.

The method consists in shifting the potential V (x) ofthe distance d/N , after each time interval T/N . For Ninteger, this leads to a time-periodic potential V (x, t)of time period T . For a su�ciently short period T ,the atomic motion is governed by the e↵ective potentialVe↵

(x), equal to the time average of V (x, t):

Ve↵

(x) =1

T

Z T

0

V (x, t)dt

=1

N

NX

j=1

V (x+ jd/N)

=X

p2ZVp e

i2⇡px/d 1

N

NX

j=1

ei2⇡pj/N

=X

p multiple of N

Vp ei2⇡px/d.

It is then apparent that the e↵ective potential Ve↵

(x) isperiodic, of period d

e↵

= d/N .

S.II. EXPRESSION FORTHE BLOCH-FLOQUET HAMILTONIAN

The modulated potential (1) is invariant under thespace and time translational symmetries Tx, Tt and T ⇤,which all commute with each other. The eigenstates of

[email protected]

the Hamiltonian can thus be written as eigenstates ofthose symmetries, which can be expressed as

q,!(x, t) = ei(qx�!t)X

j,l2Zcj,l e

il(kx�⌦t)eijNkx,

where �Nk < q Nk and 0 ! < ⌦. The spinorcoe�cients cj,l are determined by the equations

~(! + l⌦)cj,l =~2[q + (l +Nj)k]2

2mcj,l

+VL

2�x(cj,l+1

+ cj,l�1

)

+VB

2�z(cj+1,l�N + cj�1,l+N ).

The numerical data represented in Fig. 2 (main text) iscalculated using the above equations, in a truncated basis�10 j, l 10.

S.III. RESUMMATION OFTHE PERTURBATIVE EXPANSION OF He↵

The e↵ective potential Ve↵

can be calculated as a se-ries expansion in powers of the (potentially small) dimen-sionless parameters V

L

/(~⌦) and VB

/(~⌦), in the high-frequency limit ⌦ ! 1. In the main text, we provide itsexpression in Eq. (4), which corresponds to the lowest-order term. We note that this derivation, which is basedon the general formula of Ref. [S1], was obtained by ne-glecting the non-commutativity of the position and mo-mentum operators; indeed, we verified that the momen-tum operator is irrelevant in the derivation of the e↵ectivepotential, which essentially relies on the spin-dependenttime-modulated components of the Hamiltonian. Thus,in the following of this Section, which aims to derive thee↵ective potential in the strong-modulation regime, weexplicitly neglect any e↵ects associated with the kineticenergy term of the full Hamiltonian.In this Section, we first derive the expression for the

e↵ective potential Ve↵

, in the case where VL

/(~⌦) is al-lowed to take arbitrary large values (still assuming thatVB

⌧ VL

, ~⌦). Following Refs. [S2, S3], we perform aunitary transformation

| 0i = R(t) | i , R(t) = exp

✓�i

VL

~⌦ sin(kx� ⌦t)�z

◆,

2

0 1 2 3 4 50

0.5

1

1.5

VL [~⌦]

Ue↵[V

B]

(a)

0 2 4 60

0.05

0.1

VB [~⌦]

Ue↵[V

N L/(~⌦)N

�1]

(b)

FIG. S1. Depth Ue↵ of the e↵ective lattice, calculated forarbitrary values of VL/(~⌦) (a) or VB/(~⌦) (b) in the caseN = 4. The dashed lines correspond to the lowest-order per-turbation result (4), and the solid lines to the resummationresults (S.1) and (S.2).

which removes the diverging term ⇠ VL ⇠ ~⌦ from thetime-dependent potential V (x, t) in Eq. 1 (main text).This leads to a novel time-dependent potential

V 0(x, t) = R(t)V (x, t)(t)R†(t) + i~@tR(t)R†(t)

= R(t) [VB

cos(N⌦t)�x]R†(t).

Making use of the identity e�i��z�xei��z = cos(2�)�x +

i sin(2�)�y, we obtain the expression

V 0(x, t) = VB

cos(N⌦t)

cos

✓VL

~⌦ sin(kx� ⌦t)

◆�x

+ sin

✓VL

~⌦ sin(kx� ⌦t)

◆�y

�.

In the large-frequency limit ⌦ ! 1, the atom dynam-ics can be described by an e↵ective stationary potential,given by [S2, S3]

Ve↵

(x) =1

T

Z T

0

V 0(x, t)dt

= JN

✓2V

L

~⌦

◆VB

cos(2Nkx)�x, (S.1)

assuming N even, and where JN is a Bessel function ofthe first kind. This e↵ective potential corresponds to a

spin-dependent optical lattice of spacing d/N and depthUe↵

= 2JN�2VL~⌦

�VB

.A similar resummation with respect to V

B

/(~⌦) canalso be derived. Here, we make use of the Floquet repre-sentation of time-periodic Hamiltonians. We first writethe exact eigenstates of the coupling VB

2

cos(N⌦t)�x,which read

||n, sxi =X

p2ZJp

✓2sxVB

~⌦

◆|n+ pN, sxi,

where n denotes the Floquet quantum number, and sxis the spin projection along x. The energy of the state||n, sxi is equal to n~⌦. The e↵ect of the couplingVL

cos(kx�⌦t)�x can be understood using perturbationtheory in the degenerate subspace ||n,±i, which must beperformed at order N . We obtain the expression

Ve↵

(x) =Ue↵

2cos(2Nkx)�x,

Ue↵

= 4~⌦✓

VL

2~⌦

◆N������

X

PNi=1 pi=�1

QNi=1

Jpi [(�1)i 2VBN~⌦ ]QN�1

i=1

Pij=1

(1 +Npj)

������.

(S.2)

We plot in Fig. S1 the lattice depth Ue↵

given by the re-summation formulas in Eqs. (S.1)-(S.2) discussed above.We checked that the formulas (S.1) and (S.2) accountwell for the numerical results obtained via direct di-agonalization of the Bloch-Floquet equations (see Sec-tion S.II).

S.IV. MICRO-MOTION EFFECTSIN THE MOMENTUM DISTRIBUTION

In this section, we analyze how the micro-motion as-sociated with the time-modulation in Eq. (1) a↵ects themomentum distribution of atoms prepared in the e↵ec-tive potential V

e↵

of spatial period d/N [Eq. (4)]. Specif-ically, we consider an atom prepared in the ground stateof the e↵ective potential. This state can be expanded onthe family of states of momentum multiple of 2Nk (seeFig. S2a).The actual state created using time-modulated lattices

is expected to be modified by the micro-motion, as

| (t)i = e�iK(t) | 0

i , (S.3)

where the expression for the kick operator K(t) is givenin the main text [Eq. (5)]. The latter leads to additionaldi↵raction peaks at all momenta multiple of 2k, whoseamplitude vary periodically in time, with a period T/N(see Fig. S2b). This shows that Bragg di↵raction doesnot give a direct information on the ground state of thee↵ective lattice.

3

0

1

2

3n(k)[a.u

.](a)

0

1

2t = 0

(b)

0

1

2t = 0.25T/4

0

1

2t = 0.5T/4

�8 �4 0 4 80

1

2t = 0.75T/4

q [2k]

FIG. S2. (a) Momentum distribution associated with theground state of the e↵ective lattice with spacing d/4 anddepth Ue↵ = 10.9Ee↵

r . (b) Momentum distribution of thestate in (a), taking into account the micro-motion expectedfor the dynamic lattice parameters, according to Eq. S.3. Themicro-motion leads to a more complex structure comparedto (a), periodically evolving in time. The lattice parameterscorrespond to the ones of Fig. 2 in the main text.

S.V. EFFECTIVE HAMILTONIANDURING LATTICE LOADING

In this Section, we analyze the adiabatic preparation ofthe ground state associated with the e↵ective potentialVe↵

(x) of depth U0

e↵

. We consider a slow ramp of themoving-lattice depth V

L

(t) = 0 ! V 0

L

during the timeinterval 0 t t

ramp

, such that the e↵ective potential’sdepth U0

e↵

corresponds to the final value VL

(tramp

) = V 0

L

.As the definition (2) of the e↵ective Hamiltonian andkick operators assumes a constant lattice depth [S1], weexpect these notions to be modified during the ramp. Itis the aim of this Section to show how the adiabatic ramp

can still be captured by an e↵ective-Hamiltonian picture.To analyze this situation, we decompose the ramp into

N steps, and we assume that the time interval �t =tramp

/N is short enough, such that VL

can be consideredto remain constant within each step. More precisely, weassume that the lattice depth is equal to V

L

(j�t) duringthe step j�t t<(j+1)�t. We then apply the e↵ective-Hamiltonian formalism of Ref. [S1] within each time-step,and write the full time-evolution operator as

Uramp

=0Y

j=N�1

Uj ,

Uj = e�iK0[VL(j�t)]e�iHeff [VL(j�t)]�teiK0[VL(j�t)].

In the latter expression, and for the sake of simplicity,we assumed that �t was a multiple of the modulationperiod, so that the kick operators at the beginning andat the end of each step only depend on the value of V

L

(in fact, they correspond to the kick operator at the timet = 0, hence the notation K

0

).Assuming �t short enough, we write

e�iK0[VL((j+1)�t)]e�iK0[VL(j�t)] ' e�i�t(dVL/dt)dK0/dVL ,

leading to

Uramp

= e�iK(tramp)T⇢exp

✓�i

ZHramp

e↵

(t)dt/~◆�

,

where T denotes time-ordering, and where one intro-duced the slowly varying Hamiltonian

Hramp

e↵

(t) = He↵

|VL(t)+ ~dVL

dt

dK0

(tramp

)

dVL

����VL(t)

=Ue↵

(t)

2cos(2Nkx)�x � 1

dVL

dtsin(2kx)�z.

(S.4)

We now estimate a criterion for identifying the adia-batic regime of the lattice loading. Describing the latticeloading solely with the first term of eq. (S.4) would lead tothe standard adiabaticity criterion t

ramp

� ~U0

e↵

/(Ee↵

r

)2

[S4]. We expect the second term in (S.4) to drive non-adiabatic transitions for V

L

& ⌦Er

. Adiabatic latticeloading thus requires the additional constraint t

ramp

�(V 0

L

/Er

)/⌦. As we choose VL

. ~⌦, this constraintshould not be the most restrictive.

[S1] N. Goldman and J. Dalibard, Phys. Rev. X 4, 031027(2014).

[S2] P. Hauke, O. Tieleman, A. Celi, C. Olschlager, J. Si-monet, J. Struck, M. Weinberg, P. Windpassinger,

K. Sengstock, M. Lewenstein, et al., Phys. Rev. Lett.109, 145301 (2012).

[S3] N. Goldman, J. Dalibard, M. Aidelsburger, and N. R.Cooper, Phys. Rev. A 91, 033632 (2015).

4

0

0.5

1

VL(t)cos(⌦t)

(a)

0 0.5 1 1.5

0

0.5

1

U0 U1U2

�t

UN�2UN�1

t [tramp]

VL(t)cos(⌦t)

(b)

FIG. S3. (a) Evolution of the amplitude of the moving optical lattice at position x = 0, during and after the lattice ramp ofduration tramp. (b) Scheme of the ramp discretization: the time interval 0 t tramp is decomposed into N steps of duration�t. Within each step the depth VL is constant, leading to a time-periodic potential.

[S4] J. H. Denschlag, J. Simsarian, H. Ha↵ner, C. McKenzie,A. Browaeys, D. Cho, K. Helmerson, S. Rolston, and

W. Phillips, J. Phys. B: At. Mol. Opt. Phys. 35, 3095(2002).