This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Experimental realization of a high precision tunable hexagonal optical lattice

Jin-Yu Liu\authormark1,2,3,\dagger    Guang-Quan Luo\authormark2,3,\dagger    Xiao-Qiong Wang\authormark2,3    Andreas Hemmerich\authormark4    and Zhi-Fang Xu\authormark2,3,5,6 \authormark1Department of Physics, Harbin Institute of Technology, Harbin 150001, China
\authormark2Department of Physics, Southern University of Science and Technology, Shenzhen 518055, China
\authormark3Shenzhen Institute for Quantum Science and Engineering, Southern University of Science and Technology, Shenzhen 518055, China
\authormark4Institute of Laser Physics, University of Hamburg, Luruper Chaussee 149, 22761 Hamburg, Germany
\authormark5International Quantum Academy, Shenzhen 518048, China
\authormark6Guangdong Provincial Key Laboratory of Quantum Science and Engineering, Southern University of Science and Technology, Shenzhen 518055, China
\authormark\daggerThese authors contributed equally to this work.
\authormark*[email protected], [email protected]
Abstract

Hexagonal optical lattices offer a tunable platform to study exotic orbital physics in solid state materials. Here, we present a versatile high-precision scheme to implement a hexagonal optical lattice potential, which is engineered by overlapping two independent triangular optical sublattices generated by laser beams with slightly different wavelengths around 1064 nm. This enables us to precisely control the detailed structure of the hexagonal lattice by adjusting the relative position and the relative lattice depth of the two triangular optical sublattices. Taking advantage of the sensitive dependence of the second Bloch band on small lattice deformations, we propose a strategy to optimize the optical lattice geometry with an extremely high precision. This method can also be extended to other lattice configurations involving more than two sublattices. Our work provides the experimental requirements in the search for novel orbital physics of ultracold atoms, for example, in the flat pp-band of the hexagonal optical lattice.

journal: osac

1 Introduction

As an additional degree of freedom, different from charge and spin, orbital physics plays a central role for superconductivity, magnetism and transport in transition-metal-oxides due to two fundamental features, i.e., orbital degeneracy and orbital anisotropy[1]. In periodic crystals, orbitals describe the motion of electrons localized at a lattice site. Electron clouds with different symmetries can arise such as ss, pp, and dd orbitals, distinguished by their angular momentum. To simulate orbital degrees of freedom with ultracold atoms, one may apply artificial periodic potentials, called optical lattices, by interfering several laser beams, and prepare the atoms in higher Bloch bands. Wannier functions associated with the Bloch bands describe the character of the orbitals. As a versatile platform for quantum simulation [2, 3, 4, 5, 6, 7], ultracold atoms in optical lattices have become an ideal system for exploring orbital physics in electronic condensed matter. In contrast to the widely studied ss orbital physics in optical lattices [4, 8, 9], the interplay of orbital degeneracy and spatial anisotropy of higher orbitals can induce more complex many-body scenarios with remarkable new features [10].

Over the past decade, pp orbitals in optical lattices have been attracting a great deal of attention theoretically. Related experimental developments have provided the means to verify many of the theoretical predictions. In addition to ultracold fermionic gases taking the role of electrons in solid state materials [11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22], also bosons, for example loaded into the pp band, exhibit exotic many-body quantum phases due to interaction effects [23, 24, 25, 26, 27, 28, 29, 30, 31, 32, 33, 34, 35]. A pioneering experimental work of Ref. [24] successfully realized a pp-orbital superfluid in a bipartite optical square lattice. In this work, bosons condensed at two quasi-momenta located at the centers of the four boundaries forming the rectangular first Brillouin zone (BZ). The corresponding complex-valued locally chiral px±ipyp_{x}\pm ip_{y} order parameter breaks time-reversal symmetry. Recently, experimental efforts on pp-orbital physics have been directed to hexagonal optical lattices [36, 37, 38, 39]. In our recent work [38], we observed the formation of a globally chiral atomic superfluid, induced by interaction-driven time-reversal symmetry breaking in the second Bloch band of a hexagonal boron nitride (BN) optical lattice. The emergent orbital superfluid supports topological Bogoliubov excitations and edge states.

In experiments with bosons in higher Bloch bands with orbital degrees of freedom, time-reversal symmetry and discrete rotation symmetries must be implemented with high precision in order to realize and exploit orbital degeneracies. In this manuscript, we report a versatile high precision scheme to construct a hexagonal BN optical lattice and propose a strategy to precisely optimize the optical lattice geometry. Unlike other schemes [40, 36, 41, 37], we create the hexagonal optical lattice by overlapping two triangular optical lattices, which contribute sublattice sites 𝒜\mathcal{A} and \mathcal{B} to the resulting BN lattice. Each triangular lattice is generated by intersecting three monochromatic running wave laser beams with out-of-plane linear polarization. The potential depth and spatial position of each sublattice can be adjusted independently in the experiment. We swap the potential depths of the two sublattice wells to load atoms into the desired second band. In this process we added an extra step of reducing the superimposed optical dipole trap to further cool the atoms, which turned out crucial for the realization of long-lived condensates with very low temperature.

Since the energy minima of the second band sensitively depend on the relative positions of the two triangular lattices, small variations of the experimental parameters would cause a significant change in the locations of Bragg resonances in momentum spectra recorded by ballistic time-of-flight (TOF) and subsequent absorption imaging. We thus developed a strategy to optimize the optical lattice geometry. First, we calibrate the positions of Bragg peaks from the momentum spectra, which reflect the corresponding quasi-momenta for the energy minima in the second band. Second, we deduce the current experimental parameters assisted by numerical calculations and tune these parameters to correct the lattice deformation. This strategy has been readily applied in the experiment and we believe that it could be easily generalized to other optical lattice systems.

The remainder of this manuscript is organized as follows: In Sec. 2, we introduce the hexagonal optical lattice potential and analyze its deformation with the variation of experimental parameters. In Sec. 3, we describe the protocols for building and controlling the experimental setup. In Sec. 4, we show the experimental results of the TOF images of the condensates, and discuss our strategy to optimize the optical lattice potential. A summary and an outlook are given in Sec. 5.

2 Theoretical scheme for creating hexagonal optical lattice potential

Theoretically, we apply three laser beams intersecting at 120120^{\circ} angle with respect to each other in the xyxy-plane to construct the hexagonal lattice potential. The polarizations of all laser beams are parallel to the zz-axis. Each laser beam comprises two spectral components with slightly different wavelengths λ1\lambda_{1} and λ2\lambda_{2} around 10641064\,nm. The interference among each set of spectral components with the same frequency generates a triangular lattice. Therefore, the total hexagonal optical lattice potential is a superposition of two triangular lattices with corresponding lattice sites denoted respectively by 𝒜\mathcal{A} and \mathcal{B} and can be written as Vh(𝐫)=Vλ1(𝐫)+Vλ2(𝐫)V_{h}(\mathbf{r})=V_{\lambda_{1}}(\mathbf{r})+V_{\lambda_{2}}(\mathbf{r}) with

Vλ1(𝐫)=V1{3+2i,jcos[(𝐤i𝐤j)𝐫+(θiθj)]},V_{\lambda_{1}}(\mathbf{r})=-V_{1}\left\{3+2\sum_{\left<i,j\right>}\cos\left[\left(\mathbf{k}_{i}-\mathbf{k}_{j}\right)\cdot\mathbf{r}+\left(\theta_{i}-\theta_{j}\right)\right]\right\}, (1a)
Vλ2(𝐫)=V2{3+2i,jcos[(𝐤i𝐤j)𝐫+(θiθj)]}.V_{\lambda_{2}}(\mathbf{r})=-V_{2}\left\{3+2\sum_{\left<i,j\right>}\cos\left[\left(\mathbf{k}_{i}-\mathbf{k}_{j}\right)\cdot\mathbf{r}+\left(\theta^{\prime}_{i}-\theta^{\prime}_{j}\right)\right]\right\}. (1b)

Here, V1,20V_{1,2}\geq 0 as λ1,2\lambda_{1,2} are far red detuned from the atomic resonance. The summation i,j\left<i,j\right> is limited to 1,2\left<1,2\right>, 2,3\left<2,3\right>, 3,1\left<3,1\right>. The wave vectors are given by 𝐤1=kL(3/2,1/2)\mathbf{k}_{1}=k_{\mathrm{L}}(-\sqrt{3}/2,1/2), 𝐤2=kL(3/2,1/2)\mathbf{k}_{2}=k_{\mathrm{L}}(\sqrt{3}/2,1/2), and 𝐤3=kL(0,1)\mathbf{k}_{3}=k_{\mathrm{L}}\left(0,-1\right) with kL=2π/λ1k_{\mathrm{L}}=2\pi/\lambda_{1}. The phase angles are determined by θj=2πν1Lj/c+θ0\theta_{j}=2\pi\nu_{1}L_{j}/c+\theta_{0} and θj=2πν2Lj/c+θ0\theta^{\prime}_{j}=2\pi\nu_{2}L_{j}/c+\theta_{0}^{\prime}, where ν1,2=c/λ1,2\nu_{1,2}=c/\lambda_{1,2} are frequencies and LjL_{j} denotes the optical path length of the jj-th beam from the splitting point to the center of the lattice. cc denotes the speed of light. θ0\theta_{0} and θ0\theta_{0}^{\prime} include the phase information for laser beams and are irrelevant to the lattice structure.

This optical lattice system has various degrees of freedom to be conveniently adjusted in the experiment. V1,2V_{1,2} control the potential well depths of the two triangular sublattices independently. The relative position of the two triangular lattices is described by Δ𝐫\Delta\mathbf{r} which is defined via the equation Vλ1(𝐫)/V1=Vλ2(𝐫Δ𝐫)/V2V_{\lambda_{1}}(\mathbf{r})/V_{1}=V_{\lambda_{2}}(\mathbf{r}-\Delta\mathbf{r})/V_{2} and can be explicitly written as

Δ𝐫=3a04π(3Δθ12,Δθ122Δθ23).\Delta\mathbf{r}=\frac{\sqrt{3}a_{0}}{4\pi}\left(\sqrt{3}\Delta\theta_{12},-\Delta\theta_{12}-2\Delta\theta_{23}\right). (2)

Here, a0=4π/(33kL)a_{0}=4\pi/(3\sqrt{3}k_{\mathrm{L}}) and the relative phase differences Δθij\Delta\theta_{ij} are defined as

Δθij=(θiθj)(θiθj)=2πΔνc(LiLj)\Delta\theta_{ij}=(\theta_{i}-\theta_{j})-(\theta_{i}^{\prime}-\theta_{j}^{\prime})=\frac{2\pi\Delta\nu}{c}(L_{i}-L_{j}) (3)

with Δν=ν1ν2\Delta\nu=\nu_{1}-\nu_{2}. Note that there are only two independent degrees of freedom as Δθ12+Δθ23+Δθ31=0\Delta\theta_{12}+\Delta\theta_{23}+\Delta\theta_{31}=0. Thus, Δ𝐫\Delta\mathbf{r} can be rewritten as

Δ𝐫=3a0Δν2c(3L13L2,L1L2+2L3).\Delta\mathbf{r}=\frac{\sqrt{3}a_{0}\Delta\nu}{2c}\left(\sqrt{3}L_{1}-\sqrt{3}L_{2},-L_{1}-L_{2}+2L_{3}\right). (4)

When fixing L3=(L1+L2)/2L_{3}=(L_{1}+L_{2})/2, we can change L1L2L_{1}-L_{2} or Δν\Delta\nu to tune the relative position of the two triangular sublattices along the xx-direction. If we further choose (L1L2,L2L3,L3L1)=(6.04,3.02,3.02)cm(L_{1}-L_{2},L_{2}-L_{3},L_{3}-L_{1})=(-6.04,3.02,3.02)\ \text{cm} and Δν=3.308\Delta\nu=3.308 GHz, Δ𝐫=(a0,0)\Delta\mathbf{r}=(-a_{0},0) and a regular hexagonal optical lattice is formed. The unit cell in this case is a hexagon as shown in Fig. 1 (a). Adjacent vertices are associated with potential wells denoted by 𝒜\mathcal{A} and \mathcal{B}, respectively. The potential maximum is located at the center of the unit cell. A typical resulting hexagonal optical lattice potential is shown in Fig. 1 (b).

Refer to caption
Figure 1: (𝐚)\mathbf{(a)} A hexagonal lattice is obtained by superposing two triangular sublattices. The red and blue filled circles indicate the potential minima of two triangular sublattices. (𝐛)\mathbf{(b)} The hexagonal optical lattice potential with V1=7.88ERV_{1}=7.88E_{R} and V2=7.24ERV_{2}=7.24E_{R}, where ER=2kL2/(2m)E_{R}=\hbar^{2}k_{L}^{2}/(2m). kL=2π/λ1k_{L}=2\pi/\lambda_{1} is the wave numberr of the laser and mm is the atom mass of 87Rb. The red hexagon indicates the unit cell.

3 Experimental construction of hexagonal optical lattices

Our experiment starts with a 87Rb condensate with approximately 4×1044\times 10^{4} atoms in the hyperfine ground state |F=1,mF=1|F=1,m_{F}=-1\rangle without discernable thermal atoms, produced in a pancake optical dipole trap. The dipole trap is formed by two elliptical beams perpendicular to each other at the center of an ultra-high dodecagonal vacuum metal chamber, which is shown in Fig. 2 (a). The trapping frequencies are {ωx,ωy,ωz}=2π×{26.4,26.7,70.6}\{\omega_{x},\omega_{y},\omega_{z}\}=2\pi\times\{26.4,26.7,70.6\}Hz. A two-dimensional hexagonal optical lattice is formed by superimposing two sets of triangular lattices generated by three two-color beams.

Refer to caption
Figure 2: (𝐚)\mathbf{(a)} Schematic diagram of the experimental setup. The pancake optical dipole trap is created by the elliptical Gaussian beams colored in purple. The hexagonal lattice is formed by the three Gaussian beams colored in pink. (𝐛)\mathbf{(b)} The lattice depths (|V12|,|V23|,|V31|)(|V_{12}|,|V_{23}|,|V_{31}|) are plotted versus the frequency difference Δν\Delta\nu. The black squares, red circles, and blue triangles indicate the experimental data of |V12||V_{12}|, |V23||V_{23}|, and |V31||V_{31}| respectively. The black solid, red solid, and blue dashed lines are the fitting curves for the experimental data symbols with the same color. The central vertical gray dashed line and the gray solid circles mark the values of Δν\Delta\nu for two specific configurations, with a one-dimensional (1D) and hexagonal lattice geometry, respectively.

Experimentally, to create a regular hexagonal lattice we need to fine tune all parameters including L1L_{1}, L2L_{2}, L3L_{3}, and Δν\Delta\nu. First, we measure and roughly set the values of (L1L2,L2L3,L3L1)(L_{1}-L_{2},L_{2}-L_{3},L_{3}-L_{1}) as (6,3,3)(-6,3,3) cm. Second, we turn on two out of three lattice beams to form a 1D optical lattice. Loading a Bose-Einstein condensation into the lattice, we calibrate the lattice depth via amplitude modulated spectroscopy and obtain information for the lattice center difference when we turn on all three lattice beams. As an example, we consider the ii-th and jj-th lattice beams with four wave vectors (𝐤i,𝐤j,𝐤i,𝐤j)(\mathbf{k}_{i},\mathbf{k}_{j},\mathbf{k}_{i}^{\prime},\mathbf{k}_{j}^{\prime}). In the beginning, we tune the power of each spectral component of each beam to make sure that V1=V2V_{1}=V_{2} (denoted as V0V_{0}). Their combination gives rise to a 1D lattice potential, which is given by

Vij(1D)(𝐫)=4V0Vijcos{(𝐤i𝐤j)𝐫+12[(θiθj)+(θiθj)]}\displaystyle V_{ij}^{\rm(1D)}(\mathbf{r})=-4V_{0}-V_{ij}\cos\left\{\left(\mathbf{k}_{i}-\mathbf{k}_{j}\right)\cdot\mathbf{r}+\frac{1}{2}\left[(\theta_{i}-\theta_{j})+(\theta_{i}^{\prime}-\theta_{j}^{\prime})\right]\right\} (5)

where Vij=4V0cos(πΔvc(LiLj))V_{ij}=4V_{0}\cdot\cos\left(\frac{\pi\Delta v}{c}\left(L_{i}-L_{j}\right)\right).

We then change the frequency difference among two spectral components in a range of (0,10)(0,10) GHz and measure (|V12|,|V23|,|V31|)(|V_{12}|,|V_{23}|,|V_{31}|) accordingly. Fig. 2(b) shows our final results after we carefully set L1L_{1}, L2L_{2}, and L3L_{3}. Note that we use the fact that V23=V31V_{23}=V_{31} holds for arbitrary values of Δν\Delta\nu when L2L3=L3L1L_{2}-L_{3}=L_{3}-L_{1} to fine tune L3L_{3}. In Fig. 2(b), we fit the experimental data by applying the formula

|Vij(Δν)|=aij|cos2πΔνbij|,\displaystyle|V_{ij}(\Delta\nu)|=a_{ij}\left|\,\cos\frac{2\pi\Delta\nu}{b_{ij}}\,\right|, (6)

where aija_{ij} and bijb_{ij} are the fitting parameters. The fitting results are (a12=7.347(8)ER,b12=9.909(6)GHz)(a_{12}=7.347(8)\ E_{R},b_{12}=9.909(6)\,\rm GHz), (a23=7.38(2)ER,b23=19.87(2)GHz)(a_{23}=7.38(2)\ E_{R},b_{23}=19.87(2)\,\rm GHz), (a31=7.38(1)ER,b31=19.84(2)GHz)(a_{31}=7.38(1)\ E_{R},b_{31}=19.84(2)\,\rm GHz) in the confidence level of 95%\% for curves. Based on these parameters, we obtain more precision values for optical path differences as (L1L2,L2L3,L3L1)=(6.04,3.02,3.02)(L_{1}-L_{2},L_{2}-L_{3},L_{3}-L_{1})=(-6.04,3.02,3.02) cm. When the frequency difference is set at Δν=3.308GHz\Delta\nu=3.308\,\rm GHz or 6.607GHz6.607\,\rm GHz as marked by two circles in Fig. 2(b), a regular hexagonal is created.

4 Optimization of the hexagonal lattice geometry

4.1 Bose-Einstein condensation in a hexagonal optical lattice

To further determine and fine tune the geometry of the hexagonal lattice after turning on all three laser beams, we load atoms into the second Bloch band of a hexagonal BN lattice with V1V2V_{1}\neq V_{2}. Initially, atoms are loaded into the ss orbitals in the deep wells of a hexagonal BN lattice. Applying the population swapping technique [42, 43] by a rapid quench of the lattice depths in 100 μs\mu s, such that the sizes of V1V_{1}, V2V_{2} and hence the depth of the two subclasses of wells are swapped, the atoms are transferred into the ss orbitals of the shallow wells. After a hold time, the atoms recondense at the band mimina of the second Bloch band. The momentum distribution of the condensate is recorded by suddenly switching off all potentials followed by 2020\,ms free expansion before an absorption image is taken. More details can be found in Ref. [38]. We thus distinguish the degree of the deformation of the hexagonal lattice due to improper choices of the optical paths and the frequency difference Δν\Delta\nu.

Fixing the optical paths as (L1L2,L2L3,L3L1)=(6.04,3.02,3.02)(L_{1}-L_{2},L_{2}-L_{3},L_{3}-L_{1})=(-6.04,3.02,3.02) cm, we have adjusted different values of Δν\Delta\nu between 3.2823.282\,GHz and 3.3323.332\,GHz. Fig. 3(a) shows the corresponding theoretically obtained second band dispersions (first row), obtained from plane-wave expansion method, and the experimentally recorded momentum distributions for the condensate in the second band (second row). For the case of Δν=3.308GHz\Delta\nu=3.308\,\rm GHz, the second band minima are located at the two KK points denoted as KΔK_{\Delta} and KK_{\nabla} (cf. Fig. 3(b)). Changing Δν\Delta\nu, we observe that the two minima shift along the kyk_{y} direction. In Fig. 3(c), we compare the location of the second band minima with the experimental observed momentum distribution and find that Bragg peaks in experimental TOF images are in good agreement with theoretical predictions for various Δν\Delta\nu. In the low frequency difference limit, two band minima merge together at an MM point (cf. Fig. 3(b)). This phenomenon is quite similar to the band dispersion change in a shaking 1D lattice [44, 45]. Thus, a similar ferromagnetic quantum phase transition would also occur in our system when we change Δν\Delta\nu.

Refer to caption
Figure 3: (𝐚)\mathbf{(a)} The upper panel shows the single-particle energy spectra of the second band in the reciprocal space with (V1,V2)=(7.88,7.24)ER(V_{1},V_{2})=(7.88,7.24)\ E_{R}. The red dots indicate the energy minima. The lower panel shows the corresponding experimental TOF images. The black hexagons indicate the first BZ. The frequency differences Δν\Delta\nu are chosen between 3.2823.282\,GHz to 3.3323.332\,GHz. (𝐛)\mathbf{(b)} The first BZ of the hexagonal lattice in momentum space. The blue dots indicate the high symmetry points. (𝐜)\mathbf{(c)} The black lines indicate the positions of the energy minima obtained by theoretical calculations. The blue dots indicate the average positions of the atoms obtained from experimental TOF images. The upper and lower blanches indicate the top right corner and bottom right corner, respectively.

4.2 A new optimization strategy

From the momentum distribution of the condensate in the second band, we find that the positions of the Bragg peaks are very sensitive to the frequency difference Δν\Delta\nu and the optical paths L1L_{1}, L2L_{2}, and L3L_{3}, which does not occur for atoms residing in the lowest band. Hence, we can take advantage of this sensitivity to further fine correct the distortion of the hexagonal optical lattices after implementing the method mentioned in Sec. 3. Details of the optimization strategy are presented below.

Refer to caption
Figure 4: (𝐚)\mathbf{(a)} Sketch of the momentum distribution in a deformed hexagonal lattice. The red and green triangles indicate the three KΔK_{\Delta}^{\prime} and KK_{\nabla}^{\prime} points, respectively. The relative distance between the centers of the two triangles is labeled by 𝐒\mathbf{S}. (𝐛)\mathbf{(b)} The experimental TOF images correspond to δL3=173μm\delta L_{3}=-173\,\mathrm{\mu m}, δL3=98μm\delta L_{3}=-98\,\mathrm{\mu m}, and δL3=20μm\delta L_{3}=-20\,\mathrm{\mu m} respectively. The optical lattice potential is set as (V1,V2)=(7.87,7.23)ER(V_{1},V_{2})=(7.87,7.23)\ E_{R}. The black hexagon indicates the first BZ.

To avoid confusion, we use KΔK_{\Delta}^{\prime} and KK_{\nabla}^{\prime} to label the locations of the six major Bragg peaks in the momentum distributions in Fig. 4, which correspond to the condensate momenta in reciprocal space. KΔK_{\Delta} (KK_{\nabla}) and KΔK_{\Delta}^{\prime} (KK_{\nabla}^{\prime}) are identical in the case of a perfect regular hexagonal lattice. The three KΔK_{\Delta}^{\prime} and KK_{\nabla}^{\prime} points form two triangles Δ\Delta^{\prime} and \nabla^{\prime}, respectively. We note that Δ\Delta^{\prime} and \nabla^{\prime} are two regular triangles with the same size, which means that directions and powers of three lattice beams are set accurately in the experiment. In addition, every KΔK_{\Delta}^{\prime} point has a inverse KK_{\nabla}^{\prime} point and vice versa, which is ensured by the time-reversal symmetry of the system. In summary, the only significant error is the misalignment of the two centers of Δ\Delta^{\prime} and \nabla^{\prime}, which are not identical unless all KK^{\prime} points form a regular hexagon leading to a perfect regular hexagonal lattice (see Fig. 4 (a)).

The displacement from the center of \nabla^{\prime} to the center of Δ\Delta^{\prime} is labeled as 𝐒\mathbf{S}. The positions of all KK^{\prime} points can be derived by 𝐒\mathbf{S}. For example, the top right vertex of the hexagon is located at

(kx(1),ky(1))=(3kLS(1)2,kLS(2)2).(k_{x}^{(1)},k_{y}^{(1)})=(\frac{\sqrt{3}k_{L}-S(1)}{2},\frac{k_{L}-S(2)}{2}). (7)

When 𝐒=0\mathbf{S}=0, we obtain a regular hexagonal lattice with (kx(1),ky(1))=(3kL/2,kL/2)(k_{x}^{(1)},k_{y}^{(1)})=(\sqrt{3}k_{L}/2,k_{L}/2). When 𝐒0\mathbf{S}\neq 0, we only need to optimize the phases θi\theta_{i} and θi\theta_{i}^{\prime} by fine tuning L3L_{3} and Δν\Delta\nu. Due to the constraint i,jΔθij=0\sum_{\langle i,j\rangle}\Delta\theta_{ij}=0, we can choose two out of three relative phase differences, e.g. Δθ12\Delta\theta_{12} and Δθ23\Delta\theta_{23}, to determine a certain displacement 𝐒\mathbf{S}. And a displacement 𝐒\mathbf{S}, in turn, can also be used to derive the correct value of Δθ12\Delta\theta_{12} and Δθ23\Delta\theta_{23}.

We apply three steps to determine the relative phase differences Δθ12\Delta\theta_{12} and Δθ23\Delta\theta_{23}. First, we extract the positions of the six major Bragg peaks from a TOF image. The centers of two triangles Δ\Delta^{\prime} and \nabla^{\prime} are derived by averaging three KΔK_{\Delta}^{\prime} and three KK_{\nabla}^{\prime} points respectively. The displacement 𝐒\mathbf{S} between the two triangle centers is thus obtained. Secondly, we use Eq. 7 to calculate the position of the top right vertex (kx(1),ky(1))(k_{x}^{(1)},k_{y}^{(1)}), which is more accurate than extracting the position of the corresponding Bragg peak directly. Finally, Δθ12\Delta\theta_{12} and Δθ23\Delta\theta_{23} of the TOF image are obtained by finding the solution of f(Δθ12,Δθ23)=0f(\Delta\theta_{12},\Delta\theta_{23})=0, where the objective function is defined as

f(Δθ12,Δθ23)={[E2(𝐤)kx]2+[E2(𝐤)ky]2}|𝐤=(kx(1),ky(1)).f(\Delta\theta_{12},\Delta\theta_{23})=\left.\left\{\left[\frac{\partial E_{2}(\mathbf{k})}{\partial k_{x}}\right]^{2}+\left[\frac{\partial E_{2}(\mathbf{k})}{\partial k_{y}}\right]^{2}\right\}\right|_{\mathbf{k}=(k_{x}^{(1)},k_{y}^{(1)})}. (8)

Here, E2(𝐤)E_{2}(\mathbf{k}) is the energy dispersion for the second band and depends on the values of Δθ12\Delta\theta_{12} and Δθ23\Delta\theta_{23}. Obviously, f(Δθ12,Δθ23)=0f(\Delta\theta_{12},\Delta\theta_{23})=0 when (kx(1),ky(1))(k_{x}^{(1)},k_{y}^{(1)}) is a minimum of the energy spectrum. We thus optimize the objective function f(Δθ12,Δθ23)f(\Delta\theta_{12},\Delta\theta_{23}) in the neighborhood of (4π/3,2π/3)(-4\pi/3,2\pi/3) to find the correct values of Δθ12\Delta\theta_{12} and Δθ23\Delta\theta_{23}.

Applying this strategy, we can extract Δθ12\Delta\theta_{12} and Δθ23\Delta\theta_{23} from an arbitrary TOF image. If their values are different from our target values Δθ120\Delta\theta^{0}_{12} and Δθ230\Delta\theta^{0}_{23}, as estimated we can increase L3L_{3} and Δν\Delta\nu, respectively, by δL3\delta L_{3} and δΔν\delta\Delta\nu, to reach the target values, where

δL3=\displaystyle\delta L_{3}= (Δθ23Δθ12Δθ230Δθ120)cΔθ122πΔν,\displaystyle\left(\frac{\Delta\theta_{23}}{\Delta\theta_{12}}-\frac{\Delta\theta_{23}^{0}}{\Delta\theta_{12}^{0}}\right)\frac{c\Delta\theta_{12}}{2\pi\Delta\nu}, (9)
δΔν=\displaystyle\delta\Delta\nu= Δθ120Δθ12Δθ12Δν.\displaystyle\frac{\Delta\theta_{12}^{0}-\Delta\theta_{12}}{\Delta\theta_{12}}\Delta\nu.

When (Δθ12,Δθ23)=(4π/3,2π/3)(\Delta\theta_{12},\Delta\theta_{23})=(-4\pi/3,2\pi/3), a regular hexagonal lattice is created. In our experiments, we repeat the measurements and superimpose TOF images to increase the ratio of the signal to noise to evaluate the correct values of Δθ12\Delta\theta_{12} and Δθ23\Delta\theta_{23}.

To demonstrate the power of the optimization strategy, we perform experiments and tune parameters according to the strategy, which is shown in Fig. 4(b). First, we set Δν=3.308GHz\Delta\nu=3.308\,\rm GHz, and repeat experiments of loading atoms into the second band and hold for 105105\,ms. The corresponding momentum distribution from averaging over 65 experimental runs is shown in the left panel of Fig. 4(b). After evaluation, we find that to obtain a regular hexagonal lattice, we still need to change L3L_{3} and Δν\Delta\nu by small values with δL3=173μm\delta L_{3}=-173\mu m and δΔν=1.5MHz\delta\Delta\nu=1.5\,\rm MHz. We thus decrease L3L_{3} by 80μm80\,\rm\mu m and keep frequency difference Δν\Delta\nu the same since the main phase error is from L3L_{3}. After that, we perform another series of experiment and obtain the momentum distribution shown in the middle panel of Fig. 4(b). By further evaluation, we obtain δL3=98μm\delta L_{3}=-98\,\rm\mu m and δΔν=0.7MHz\delta\Delta\nu=0.7\,\rm MHz. Later, we decrease L3L_{3} by 80μm80\,\rm\mu m again and perform a third series of experiments to obtain the momentum distribution shown in the right panel of Fig. 4(b). Finally, we obtain δL3=20μm\delta L_{3}=-20\,\rm\mu m and δΔν=0.6MHz\delta\Delta\nu=0.6\,\rm MHz. It corresponds to Δθ12=4π/3+0.0002π\Delta\theta_{12}=-4\pi/3+0.0002\pi and Δθ23=2π/30.0006π\Delta\theta_{23}=2\pi/3-0.0006\pi, which is close to the precision limit we can achieve. We thus conclude that we have already created a regular hexagonal optical lattice.

5 Summary and outlook

In conclusion, we apply an effective high-precision scheme to construct a highly tunable precisely hexagonal optical lattice in the experiment. Furthermore, we propose and implement strategies to optimize the lattice geometry by carefully determining the positions of Bragg peaks from momentum distributions of the condensate in the second Bloch band. All experimental results are in good agreement with theoretical calculations. Our work paves the way for investigating many-body physics in the hexagonal optical lattice, especially for novel higher orbital physics, which sensitively depends on the lattice rotational symmetry. The schemes of constructing and optimizing the optical lattice can potentially be extended to other non-square lattices.

\bmsection

Acknowledgments This work is supported by the Key-Area Research and Development Program of Guangdong Province (Grant No. 2019B030330001), the National Key R&D Program of China (Grant No. 2018YFA0307200), NSFC (Grant No. U1801661), and a fund from Guangdong province (Grant No. 2019ZT08X324). X.-Q. W. acknowledges support from the China Postdoctoral Science Foundation (No. 2021M691444). A.H. acknowledges support by Cluster of Excellence CUI: Advanced Imaging of Matter of the Deutsche Forschungsgemeinschaft (DFG) - EXC 2056 - project ID 390715994.

\bmsection

Disclosures The authors declare no conflicts of interest.

\bmsection

Data Availability Statement Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

References

  • [1] Y. Tokura and N. Nagaosa, “Orbital physics in transition-metal oxides,” \JournalTitlescience 288, 462–468 (2000).
  • [2] D. Jaksch and P. Zoller, “The cold atom Hubbard toolbox,” \JournalTitleAnnals of Physics 315, 52–79 (2005).
  • [3] M. Lewenstein, A. Sanpera, V. Ahufinger, B. Damski, A. Sen, and U. Sen, “Ultracold atomic gases in optical lattices: mimicking condensed matter physics and beyond,” \JournalTitleAdvances in Physics 56, 243–379 (2007).
  • [4] I. Bloch, J. Dalibard, and W. Zwerger, “Many-body physics with ultracold gases,” \JournalTitleRev. Mod. Phys. 80, 885–964 (2008).
  • [5] I. Bloch, J. Dalibard, and S. Nascimbène, “Quantum simulations with ultracold quantum gases,” \JournalTitleNature Physics 8, 267–276 (2012).
  • [6] N. Goldman, J. C. Budich, and P. Zoller, “Topological quantum matter with ultracold gases in optical lattices,” \JournalTitleNature Physics 12, 639–645 (2016).
  • [7] C. Gross and I. Bloch, “Quantum simulations with ultracold atoms in optical lattices,” \JournalTitleScience 357, 995–1001 (2017).
  • [8] M. Lewenstein, A. Sanpera, and V. Ahufinger, Ultracold Atoms in Optical Lattices: Simulating quantum many-body systems (Oxford University Press, 2012).
  • [9] O. Dutta, M. Gajda, P. Hauke, M. Lewenstein, D.-S. Lühmann, B. A. Malomed, T. Sowiński, and J. Zakrzewski, “Non-standard hubbard models in optical lattices: a review,” \JournalTitleReports on Progress in Physics 78, 066001 (2015).
  • [10] X. Li and W. V. Liu, “Physics of higher orbital bands in optical lattices: a review,” \JournalTitleReports on Progress in Physics 79, 116401 (2016).
  • [11] C. Wu, D. Bergman, L. Balents, and S. Das Sarma, “Flat bands and wigner crystallization in the honeycomb optical lattice,” \JournalTitlePhys. Rev. Lett. 99, 070401 (2007).
  • [12] C. Wu and S. Das Sarma, “px,y{p}_{x,y}-orbital counterpart of graphene: Cold atoms in the honeycomb optical lattice,” \JournalTitlePhys. Rev. B 77, 235107 (2008).
  • [13] K. Wu and H. Zhai, “Theory of quantum antiferromagnetism of fermions in an optical lattice with a half-filled pp band,” \JournalTitlePhys. Rev. B 77, 174431 (2008).
  • [14] C. Wu, “Orbital ordering and frustration of pp-band mott insulators,” \JournalTitlePhys. Rev. Lett. 100, 200406 (2008).
  • [15] E. Zhao and W. V. Liu, “Orbital order in mott insulators of spinless pp-band fermions,” \JournalTitlePhys. Rev. Lett. 100, 160403 (2008).
  • [16] C. Wu, “Orbital analogue of the quantum anomalous hall effect in pp-band systems,” \JournalTitlePhys. Rev. Lett. 101, 186807 (2008).
  • [17] M. Zhang, H.-h. Hung, C. Zhang, and C. Wu, “Quantum anomalous hall states in the pp-orbital honeycomb optical lattices,” \JournalTitlePhys. Rev. A 83, 023615 (2011).
  • [18] K. Sun, W. V. Liu, A. Hemmerich, and S. Das Sarma, “Topological semimetal in a fermionic optical lattice,” \JournalTitleNature Physics 8, 67–70 (2012).
  • [19] B. Liu, X. Li, B. Wu, and W. V. Liu, “Chiral superfluidity with p-wave symmetry from an interacting s-wave atomic Fermi gas,” \JournalTitleNature Communications 5, 5064 (2014).
  • [20] Z. Zhou, E. Zhao, and W. V. Liu, “Spin-orbital exchange of strongly interacting fermions in the pp band of a two-dimensional optical lattice,” \JournalTitlePhys. Rev. Lett. 114, 100406 (2015).
  • [21] Z.-F. Xu, A. Hemmerich, and W. V. Liu, “Odd-parity topological superfluidity for fermions in a bond-centered square optical lattice,” \JournalTitlePhys. Rev. A 96, 053607 (2017).
  • [22] M. Hachmann, Y. Kiefer, J. Riebesehl, R. Eichberger, and A. Hemmerich, “Quantum degenerate fermi gas in an orbital optical lattice,” \JournalTitlePhys. Rev. Lett. 127, 033201 (2021).
  • [23] C. Wu, W. V. Liu, J. Moore, and S. D. Sarma, “Quantum stripe ordering in optical lattices,” \JournalTitlePhys. Rev. Lett. 97, 190406 (2006).
  • [24] G. Wirth, M. Ölschläger, and A. Hemmerich, “Evidence for orbital superfluidity in the p-band of a bipartite optical square lattice,” \JournalTitleNature Physics 7, 147–153 (2011).
  • [25] P. Soltan-Panahi, D.-S. Lühmann, J. Struck, P. Windpassinger, and K. Sengstock, “Quantum phase transition to unconventional multi-orbital superfluidity in optical lattices,” \JournalTitleNature Physics 8, 71–75 (2012).
  • [26] X. Li, Z. Zhang, and W. V. Liu, “Time-reversal symmetry breaking of pp-orbital bosons in a one-dimensional optical lattice,” \JournalTitlePhys. Rev. Lett. 108, 175302 (2012).
  • [27] F. Pinheiro, G. M. Bruun, J.-P. Martikainen, and J. Larson, “xyzxyz quantum heisenberg models with pp-orbital bosons,” \JournalTitlePhys. Rev. Lett. 111, 205302 (2013).
  • [28] T. Sowiński, M. Łącki, O. Dutta, J. Pietraszewicz, P. Sierant, M. Gajda, J. Zakrzewski, and M. Lewenstein, “Tunneling-induced restoration of the degeneracy and the time-reversal symmetry breaking in optical lattices,” \JournalTitlePhys. Rev. Lett. 111, 215302 (2013).
  • [29] M. Ölschläger, T. Kock, G. Wirth, A. Ewerbeck, C. M. Smith, and A. Hemmerich, “Interaction-induced chiral px ±\pm ipy superfluid order of bosons in an optical lattice,” \JournalTitleNew Journal of Physics 15, 083041 (2013).
  • [30] T. Kock, M. Ölschläger, A. Ewerbeck, W.-M. Huang, L. Mathey, and A. Hemmerich, “Observing chiral superfluid order by matter-wave interference,” \JournalTitlePhys. Rev. Lett. 114, 115301 (2015).
  • [31] T. Kock, C. Hippler, A. Ewerbeck, and A. Hemmerich, “Orbital optical lattices with bosons,” \JournalTitleJournal of Physics B: Atomic, Molecular and Optical Physics 49, 042001 (2016).
  • [32] Y. Li, J. Yuan, A. Hemmerich, and X. Li, “Rotation-symmetry-enforced coupling of spin and angular momentum for pp-orbital bosons,” \JournalTitlePhys. Rev. Lett. 121, 093401 (2018).
  • [33] S. Lieu, A. F. Ho, D. K. K. Lee, and P. Coleman, “Intertwined superfluidity and density wave order in a pp-orbital bose condensate,” \JournalTitlePhys. Rev. B 99, 014504 (2019).
  • [34] H. Chen and X. C. Xie, “Orbital order in a bosonic pp-band triangular lattice,” \JournalTitlePhys. Rev. B 103, 205144 (2021).
  • [35] Y. Li, J. Yuan, X. Zhou, and X. Li, “Spin-induced orbital frustration in a hexagonal optical lattice,” \JournalTitlePhys. Rev. Research 3, 033274 (2021).
  • [36] M. Weinberg, C. Staarmann, C. Ölschläger, J. Simonet, and K. Sengstock, “Breaking inversion symmetry in a state-dependent honeycomb lattice: artificial graphene with tunable band gap,” \JournalTitle2D Materials 3, 024005 (2016).
  • [37] S. Jin, W. Zhang, X. Guo, X. Chen, X. Zhou, and X. Li, “Evidence of potts-nematic superfluidity in a hexagonal s p 2 optical lattice,” \JournalTitlePhysical Review Letters 126, 035301 (2021).
  • [38] X.-Q. Wang, G.-Q. Luo, J.-Y. Liu, W. V. Liu, A. Hemmerich, and Z.-F. Xu, “Evidence for an atomic chiral superfluid with topological excitations,” \JournalTitleNature 596, 227–231 (2021).
  • [39] T. Klafka, “Bose-einstein condensation in higher bloch bands of the optical honeycomb lattice,” Ph.D. thesis, Staats-und Universitätsbibliothek Hamburg Carl von Ossietzky (2021).
  • [40] P. Soltan-Panahi, J. Struck, P. Hauke, A. Bick, W. Plenkers, G. Meineke, C. Becker, P. Windpassinger, M. Lewenstein, and K. Sengstock, “Multi-component quantum gases in spin-dependent hexagonal lattices,” \JournalTitleNature Physics 7, 434–440 (2011).
  • [41] L. Duca, T. Li, M. Reitter, I. Bloch, M. Schleier-Smith, and U. Schneider, “An aharonov-bohm interferometer for determining bloch band topology,” \JournalTitleScience 347, 288–292 (2015).
  • [42] M. Anderlini, P. J. Lee, B. L. Brown, J. Sebby-Strabley, W. D. Phillips, and J. V. Porto, “Controlled exchange interaction between pairs of neutral atoms in an optical lattice,” \JournalTitleNature 448, 452–456 (1999).
  • [43] S. Fölling, S. Trotzky, P. Cheinet, M. Feld, R. Saers, A. Widera, T. Müller, and I. Bloch, “Direct observation of second-order atom tunnelling,” \JournalTitleNature 448, 1029–1032 (2007).
  • [44] C. V. Parker, L.-C. Ha, and C. Chin, “Direct observation of effective ferromagnetic domains of cold atoms in a shaken optical lattice,” \JournalTitleNature Physics 9, 769–774 (2013).
  • [45] L. W. Clark, L. Feng, and C. Chin, “Universal space-time scaling symmetry in the dynamics of bosons across a quantum phase transition,” \JournalTitleScience 354, 606–610 (2016).