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

Bosonic fractional quantum Hall conductance in shaken honeycomb optical lattices without flat bands

Shiwan Miao    Zhongchi Zhang    Yajuan Zhao    Zihan Zhao    Huaichuan Wang Department of Physics and State Key Laboratory of Low Dimensional Quantum Physics, Tsinghua University, Beijing, 100084, China    Jiazhong Hu [email protected] Department of Physics and State Key Laboratory of Low Dimensional Quantum Physics, Tsinghua University, Beijing, 100084, China Frontier Science Center for Quantum Information, Beijing, 100084, China
Abstract

We propose a scheme to realize bosonic fractional quantum Hall conductance in shaken honeycomb optical lattices. This scheme does not require a very flat band, and the necessary long-range interaction relies on the s-wave scattering which is common in many ultracold-atom experiments. By filling the lattice at 1/4 with identical bosons under Feshbach resonance, two degenerate many-body ground states share one Chern number of 1 and exactly correspond to the fractional quantum Hall conductance of 1/2. Meanwhile, we prove that the fractional quantum Hall state can be prepared by adiabatically turning on the lattice shaking, and the fractional conductance is robust in the shaken lattice. This provides an easy way to initialize and prepare the fractional quantum Hall states in ultracold-atom platforms, and it paves a new way to investigate and simulate strong-correlated quantum matters with degenerate quantum gas.

I Introduction

The fractional quantum Hall (FQH) effect is one of the most fascinating phenomena in past decades Stormer et al. (1999), where multiple many-body ground states share one integer Chern number and effectively each of the band obtains a fractional number to characterize the conductivity. In previous studies Stormer et al. (1999); Cooper (2020); Wen (2017), it was proved that the FQH effect can be achieved either in fermions or bosons. However, due to the fermionic nature of electrons in conventional materials, the FQH effect has only been observed experimentally in fermions. Thus, this leaves an open question about how to realize the bosonic FQH effect experimentally, for example, to realize and probe the Hall conductance corresponding to one half.

With the development of quantum simulations Bloch et al. (2008); Cooper et al. (2019), the platforms of ultracold atoms now provide good opportunities to study and simulate strong-correlated many-body systems Sørensen et al. (2005); Hafezi et al. (2007); Cooper and Dalibard (2013a); Wang et al. (2011); Neupert et al. (2011); Hudomal et al. (2019); Sun et al. (2011); Tang et al. (2011); Grushin et al. (2014); Cooper and Dalibard (2013b), particularly for bosonic FQH effect. There are many pioneer experiments in realizing non-trivial Chern numbers or large synthetic gauge fields Miyake et al. (2013); Aidelsburger et al. (2013); Jotzu et al. (2014); Aidelsburger et al. (2015, 2011); Wintersperger et al. (2020); Beeler et al. (2013); Stuhl et al. (2015); Lin et al. (2009) in order to reach regimes of strong correlations. Inspired by previous studies of FQH effect in “Haldane-like” models Wang et al. (2011); Neupert et al. (2011); Haldane (1988); Jotzu et al. (2014), we find that the bosonic FQH conductance can be achieved and experimentally initialized in shaken honeycomb optical lattices without synthetic magnetic fields. This scheme relies on the Feshbach resonance at s-wave scatterings Chin et al. (2010) which does not require special long-range interactions.

Meanwhile, compared to previous studies Wang et al. (2011); Neupert et al. (2011); Sun et al. (2011); Tang et al. (2011) requiring a flat band where the band gap is much larger than the band width, our scheme is realized with the nearest-neighbor-hopping model and does not require a very flat band. Our band gap is almost the same as the band width. This avoids special designs for high-order hoppings for far-apart lattice sites to flatten the bands, and reduces the complexity of Hamiltonian engineering. Furthermore, we find that the states with FQH conductance can be prepared by adiabatically turning on the shaking of static optical lattices which is topologically trivial. This simplifies the procedures to initialize the FQH states in optical lattices. We believe this scheme will inspire new opportunities to experimentally study the bosonic FQH effect in an easier way.

II Shaken honeycomb optical lattices and single-body topology

The honeycomb optical lattice is formed by the interference of three red-detuned lasers at the same frequency, which have a relative angle at 120 degrees and are in one incident plane (Fig. 1a). The polarizations of lattice beams are parallel to the incident plane. The dipole potential VopV_{op} of the optical lattice can be written in the form of

Vop\displaystyle V_{op} =\displaystyle= VD|eikyx^+eik(32x+12y)(12x^+32y^)\displaystyle-{V_{D}}\left|e^{-iky}\hat{x}+e^{ik({\sqrt{3}\over 2}x+{1\over 2}y)}(-{1\over 2}\hat{x}+{\sqrt{3}\over 2}\hat{y})\right. (1)
+eik(32x+12y)(12x^32y^)|2\displaystyle\left.+e^{ik({-\sqrt{3}\over 2}x+{1\over 2}y)}(-{1\over 2}\hat{x}-{\sqrt{3}\over 2}\hat{y})\right|^{2}
=\displaystyle= VD[3cos3kxcos3k(x/2+3y/2)\displaystyle-V_{D}\left[3-\cos\sqrt{3}kx-\cos\sqrt{3}k(x/2+\sqrt{3}y/2)\right.
cos3k(x/2+3y/2)],\displaystyle\left.-\cos\sqrt{3}k(-x/2+\sqrt{3}y/2)\right],

where VDV_{D} is the trap depth in our definition, k=2π/λk=2\pi/\lambda is the wave vector, and λ\lambda is the wavelength of lattice lasers. This gives a lattice spacing a=λ/23a=\lambda/2\sqrt{3}, corresponding to a hexagon whose side length is λ/23\lambda/2\sqrt{3}. When cold atoms are trapped in honeycomb lattices, they can be described by a tight-binding model, whose Hamiltonian H0H_{0} is H0/=i,jt0c^ic^jH_{0}/\hbar=-\sum_{\langle i,j\rangle}t_{0}\hat{c}^{\dagger}_{i}\hat{c}_{j} where t0t_{0} is the nearest-neighbor hopping amplitude; c^i\hat{c}_{i}^{\dagger} (c^i\hat{c}_{i}) is the creation (annihilation) operator on lattice cite ii; \langle\cdot\rangle corresponds to the summation of all nearest-neighbor sites.

In order to create non-trivial topological bands in tight-binding honeycomb lattice, we apply a periodic modulation on the phases of the lattice beams to break the time reversal symmetry, where phases of each laser beam are modulated according to ϕ1=ϕAcos(Ωt+π/2)\phi_{1}=\phi_{A}\cos(\Omega t+\pi/2), ϕ2=ϕAcos(Ωtπ/6)\phi_{2}=\phi_{A}\cos(\Omega t-\pi/6), and ϕ3=ϕAcos(Ωt+5π/6)\phi_{3}=\phi_{A}\cos(\Omega t+5\pi/6). This creates a shaken lattice of which each site ii follows a circular motion ri(t)\vec{r}_{i}(t) where ri(t)=ri,0A[cos(Ωt)x^+sin(Ωt)y^]\vec{r}_{i}(t)=\vec{r}_{i,0}-A[\cos(\Omega t)\hat{x}+\sin(\Omega t)\hat{y}] and AA is orbital radius of the circular trajectory. The lattice shaking alternates the original Hamiltonian H0H_{0} into a time-dependent form H^(t)\hat{H}^{\prime}(t), and

H^(t)/=i,jeizijsin(Ωt+ϕij)t0c^ic^j+h.c.,\hat{H}^{\prime}(t)/\hbar=-\sum_{\langle i,j\rangle}e^{iz_{ij}\sin(\Omega t+{\phi}_{ij})}t_{0}\hat{c}^{\dagger}_{i}\hat{c}_{j}+h.c., (2)

where i,j\langle i,j\rangle corresponds to a pair of nearest-neighbor lattice sites, zij=maΩAρij/z_{ij}=m_{a}\Omega A\rho_{ij}/\hbar, mam_{a} is the mass of an atom, and ρijeiϕij=(ri,0rj,0)(x^+y^eiπ/2)\rho_{ij}e^{i\phi_{ij}}=(\vec{r}_{i,0}-\vec{r}_{j,0})\cdot(\hat{x}+\hat{y}e^{-i\pi/2}). Here zijz_{ij} is the ratio of maΩ2Aρijm_{a}\Omega^{2}A\rho_{ij} to Ω\hbar\Omega, where the numerator is the product of the centrifugal shaking force maΩ2Am_{a}\Omega^{2}A and the distance ρij\rho_{ij}, and the denominator is the Floquet energy Ω\hbar\Omega.

A periodic Hamiltonian is decomposed into a Fourier transformation that H^(t)=lH^leilΩt\hat{H}^{\prime}(t)=\sum_{l\in\mathbb{Z}}\hat{H}_{l}e^{il\Omega t}. When Ω\Omega is large, we obtain an effective time-indepedent Floquet Hamiltonian H^fl\hat{H}_{fl} based on high frequency expansion method Goldman and Dalibard (2014); Eckardt and Anisimovas (2015); Rahav et al. (2003); Bukov et al. (2015). We rewrite the effective Floquet Hamiltonian H^fl\hat{H}_{fl} based on the nearest-neighbor (NN), next-nearest-neighbor (NNN), next-next-nearest-neighbor (NNNN), and next-next-next-nearest-neighbor (NNNNN) hoppings, i.e.

H^fl/\displaystyle\hat{H}_{fl}/\hbar =\displaystyle=- i,jt~0c^ic^ji,j2t~1c^ic^ji,j3t~2c^ic^j\displaystyle\sum_{\langle i,j\rangle}\tilde{t}_{0}\hat{c}^{\dagger}_{i}\hat{c}_{j}-\sum_{\langle i,j\rangle_{2}}\tilde{t}_{1}\hat{c}^{\dagger}_{i}\hat{c}_{j}-\sum_{\langle i,j\rangle_{3}}\tilde{t}_{2}\hat{c}^{\dagger}_{i}\hat{c}_{j} (3)
i,j4t~3c^ic^j+h.c.,\displaystyle-\sum_{\langle i,j\rangle_{4}}\tilde{t}_{3}\hat{c}^{\dagger}_{i}\hat{c}_{j}+h.c.,

where \langle\cdot\rangle, 2\langle\cdot\rangle_{2}, 3\langle\cdot\rangle_{3}, and 4\langle\cdot\rangle_{4} correspond to the summations of the NN, NNN, NNNN, and NNNNN sites. We plot t~m\tilde{t}_{m} in Fig. 1c and the detailed formula of t~m\tilde{t}_{m} can be found in the appendix. The effective NNN hopping amplitude has an imaginary part, which breaks the time-reversal symmetry, opens the band gap, and gives a non-zero Chern number to each band. The band flatness ratio is usually used to characterize the potential of many-body topology Parameswaran et al. (2013), where the ratio is defined by the band gap divided by the band width of the ground band. In Fig. 1, we present the flatness ratio, hopping strength and energy bands under different parameters. Usually a flat band is required to show the dominated bosonic FQH effect Wang et al. (2011); Neupert et al. (2011), while flattening the band is a challenge in ultracold-atom experiments since the intrinsic long-range hoppings are strongly suppressed for remote lattice sites. In our scheme, the flatness ratio is less than 2 where the gap is near the same as the band width. We find it is still suitable to realize the FQH conductance of 1/2.

Refer to caption
Figure 1: (a) The schematic of shaken honeycomb optical lattices. The definition of the mm- (nn-) axis of lattice coordinate and the hopping terms are marked on the lattice. When the phases of three beams are modulated by ϕ1=ϕAcos(Ωt+π/2)\phi_{1}=\phi_{A}\cos(\Omega t+\pi/2), ϕ2=ϕAcos(Ωtπ/6)\phi_{2}=\phi_{A}\cos(\Omega t-\pi/6), and ϕ3=ϕAcos(Ωt+5π/6)\phi_{3}=\phi_{A}\cos(\Omega t+5\pi/6), the lattice sites move along a counterclockwise circular trajectory whose angle frequency is Ω\Omega and orbital radius is proportional to ϕA\phi_{A}. (b) The band flatness ratio versus modulation parameters z0z_{0} and t0/Ωt_{0}/\Omega. (c) Effective tunnelings t~m\tilde{t}_{m} versus z0z_{0} at t0/Ω=0.1t_{0}/\Omega=0.1 (red line in panel b). (d) The single-particle band at z0=2.3z_{0}=2.3, t0/Ω=0.108t_{0}/\Omega=0.108 (red cross in panel b), where each band has a non-zero Chern number.

III Many-body topology with strongly interacting bosons

Now we switch our description from single-body physics to many-body physics. To achieve the fractional conductance, we need both strong on-site interactions and finite nearest-neighbor interactions. This can be realized by conventional s-wave Feshbach resonance Chin et al. (2010). For a given trap depth VDV_{D} in Eq. 1, we calculate the Wannier functions of each site by the methods in Ref. Marzari et al. (2012) (See Appendix B for more details). The on-site interaction UU is obtained by the integral U=4π2asmad3𝒓w(𝒓)w(𝒓)w(𝒓)w(𝒓)U={4\pi\hbar^{2}a_{s}\over m_{a}}\int d^{3}\bm{r}w^{\dagger}(\bm{r})w^{\dagger}(\bm{r})w(\bm{r})w(\bm{r}) where asa_{s} is the scattering length and mam_{a} is the mass of one atom. w(𝒓𝒓i)w(\bm{r}-\bm{r}_{i}) is the Wannier function centered at the position 𝒓i\bm{r}_{i}. Usually in Bose-Hubbard models, we only keep the on-site interactions and ignore the NN interaction V1V_{1}. However, once the scattering length asa_{s} approaches to infinity under Feshbach resonance, the NN interaction V1V_{1} becomes significant with a form of Zhai (2021)

V1\displaystyle V_{1} =\displaystyle= 8π2asmad3rw(𝒓)w(𝒓𝒂)w(𝒓𝒂)w(𝒓)\displaystyle{8\pi\hbar^{2}a_{s}\over m_{a}}\int d^{3}rw^{\dagger}(\bm{r})w^{\dagger}(\bm{r}-\bm{a})w(\bm{r}-\bm{a})w(\bm{r}) (4)
=\displaystyle= 2Ud3rw(𝒓)w(𝒓𝒂)w(𝒓𝒂)w(𝒓)d3rw(𝒓)w(𝒓)w(𝒓)w(𝒓).\displaystyle 2U{\int d^{3}rw^{\dagger}(\bm{r})w^{\dagger}(\bm{r}-\bm{a})w(\bm{r}-\bm{a})w(\bm{r})\over\int d^{3}rw^{\dagger}(\bm{r})w^{\dagger}(\bm{r})w(\bm{r})w(\bm{r})}.

Here 𝒂\bm{a} corresponds to the relative vector between two nearest-neighbor sites.

In rubidium 85, there is a Feshbach resonance at 155.3 G Courteille et al. (1998); Claussen et al. (2003) with a width around 11 G and a background scattering length -441a0a_{0} where a0a_{0} is the Bohr radius. Considering that magnetic-field fluctuations can be controlled within 1 mG in most of labs, the scattering length can be tuned up to 10000a0a_{0} with a relative uncertainty less than 0.25%. This offers the opportunity that the on-site interaction reaches the hard-core regime while the nearest-neighbor interaction is still important in the tight-binding model. In Table 1, we list the hopping coefficients and interaction strength versus VDV_{D} in unshaken lattices. At the condition of as=10000a0a_{s}=10000a_{0} and VD=28ErV_{D}=28E_{r} (Er=2k22maE_{r}={\hbar^{2}k^{2}\over 2m_{a}}), the original t0t_{0} is 0.031Er0.031E_{r} and the on-site interaction UU is 88Er88E_{r}, which satisfies Ut0U\gg t_{0}. UU strongly repulses any doublons in one site. Meanwhile, the NN interaction V1V_{1} is 0.56t00.56\hbar t_{0} and is significant compared with the hopping amplitudes.

Table 1: V1V_{1} versus the trap depth VDV_{D}. We calculate the intrinsic NN tunneling t0t_{0}, NNN tunneling t1t_{1}, and NNNN tunneling t2t_{2} of the original unshaken lattice, where t1t_{1} and t2t_{2} are usually negligible compared with t0t_{0}. Here we assume there is a perpendicular trap tightly confining cold atoms into 2D degenerate gas, and the tight trap is described by a harmonic trap with a vibrational frequency at 50 kHz.
VD/ErV_{D}/E_{r} t0/Er\hbar t_{0}/E_{r} t1/t0\hbar t_{1}/t_{0} 104t2/t010^{4}t_{2}/t_{0} U/ErU/E_{r} V1/2ErV_{1}/2E_{r} V1/t0V_{1}/t_{0}
20 0.059 -0.020 1616 71.34 0.034 1.16
22 0.050 -0.016 1111 75.47 0.024 0.96
24 0.043 -0.014 7.47.4 79.41 0.017 0.78
26 0.037 -0.011 5.25.2 83.19 0.012 0.64
28 0.031 -0.0096 3.63.6 86.82 0.0088 0.58
30 0.027 -0.0081 2.62.6 90.31 0.0063 0.48
32 0.023 -0.0068 1.81.8 93.68 0.0047 0.40

In the strong-correlation regime, the single-particle band description cannot capture the actual physics. Therefore, using the twisted boundary condition, we write out the many-body Floquet Hamiltonian H^flmany\hat{H}_{fl-many} of hard-core bosons with the nearest-neighbor interaction V1V_{1} under the trap depth VD=28ErV_{D}=28E_{r}. Here the twisted boundary condition is ψ(𝒓𝒊+Lm𝒎^)=eiθxψ(𝒓𝒊)\psi(\bm{r_{i}}+L_{m}\hat{\bm{m}})=e^{i\theta_{x}}\psi(\bm{r_{i}}) and ψ(𝒓𝒊+Ln𝒏^)=eiθyψ(𝒓𝒊)\psi(\bm{r_{i}}+L_{n}\hat{\bm{n}})=e^{i\theta_{y}}\psi(\bm{r_{i}}) where LmL_{m}(LnL_{n}) is the lattice size along mm- (nn-) axis and the axes are marked in Fig. 1a. At θx=θy=0\theta_{x}=\theta_{y}=0, the twisted boundary condition is the same as the periodic boundary condition. H^flmany\hat{H}_{fl-many} is a giant-sized sparse matrix (see Appendix E for more information). Then we apply the exact diagonalization to calculate the lowest four energy levels and their corresponding many-body bands in the case of 6 bosons in 24 lattice sites (H^flmany\hat{H}_{fl-many} has a size around 105×10510^{5}\times 10^{5}), where the bands are characterized by θx\theta_{x} and θy\theta_{y} instead of the quasi-momenta kxk_{x} and kyk_{y}.

By scanning the Floquet parameter z0z_{0} and NN interaction V1V_{1}, there are some regions that the lowest two many-body bands cross each other while they are away from the third band. In Fig. 2a, we present the energy difference between the second and third energy states (E3E_{3}-E2E_{2}) under the twisted boundary condition. Here we take the inherent long-range hopping beyond the tight-binding models (t1t_{1} and t2t_{2}) into accounts in calculating the 0th-order effective Hamiltonian, and neglect them in higher-order expansions. We mark the candidates for the FQH conductance at 1/2 in Fig. 2a according to the gap opening. In Fig. 2c, we plot the many-body bands versus θx\theta_{x} and θy\theta_{y} in this phase regime. It shows that two lowest bands cross each other and are away from the third band.

To further verify the conductance, we calculate the total Chern number 𝒞\mathcal{C} of the lowest two bands. Here 𝒞\mathcal{C} equals 12πK=1,2𝑑θx𝑑θyFxy,K(θx,θy){1\over 2\pi}\sum_{K=1,2}\int d\theta_{x}d\theta_{y}F_{xy,K}(\theta_{x},\theta_{y}) where Fxy,K(θx,θy)=Im(ψKθy|ψKθxψKθx|ψKθy)F_{xy,K}(\theta_{x},\theta_{y})=\textrm{Im}(\langle\frac{\partial\psi_{K}}{\partial\theta_{y}}|\frac{\partial\psi_{K}}{\partial\theta_{x}}\rangle-\langle\frac{\partial\psi_{K}}{\partial\theta_{x}}|\frac{\partial\psi_{K}}{\partial\theta_{y}}\rangle) is the Berry curvature of the KK-th state in the ground state manifold. We divide the whole θx\theta_{x}-θy\theta_{y} space into 20×\times20 pieces, and calculate the energies and the discrete Berry curvatures Fukui et al. (2005) in Fig. 2c and d. The two lowest bands share one integer Chern number 𝒞=1\mathcal{C}=-1 together, corresponding to FQH conductance of 1/2. We calculate cases in lattices of different sizes to show robustness of our scheme against the finite size effect (Fig. 2b), while the size of H^flmany\hat{H}_{fl-many} increases dramatically (H^flmany\hat{H}_{fl-many} has a size of 108×10810^{8}\times 10^{8} for 36 sites).

Refer to caption
Figure 2: (a) Different phases distinguished by many body gap E3E2E_{3}-E_{2}. We also label out NOON phase here. While the system is in NOON states, the honeycomb lattice is decomposed into two individual sets of triangle lattices. The atoms fill one of the triangle lattices and leaves the other empty. The actual ground states are the superpositions of these two possibilities, like a NOON state. (b) The variation of many-body energy levels with lattice sites NN under VD=28ErV_{D}=28E_{r}. Here it shows the same behavior as Ref Wang et al. (2011) and maintains the FQH band gap. (c) Many-body bands at VD=28ErV_{D}=28E_{r}. (d) Discrete Berry curvature of two lowest states while the integral is -1.

IV Robustness against the higher band excitation and Floquet heating

The tight-binding Hamiltonian in Eq. 2 only considers the contributions from s-bands, while a large scattering length may cause the s-band to mix with higher-excited bands. We follow the same treatment in Ref Viebahn et al. (2021) which successfully describes the interband transitions in optical lattices. We build a model with two sites, two particles, and both s-bands and p-bands to estimate the contributions from higher bands. The detailed numerical calculations and analyses are listed in the appendix C, and only the conclusion is presented in the main text. The dominant interband transitions induced by a large scattering length is that two neighboring particles in s-bands will scatter to p-bands together due to collisions. The coupling matrix element of this process is less than 0.1ErE_{r}, but the energy detuning is more than 10ErE_{r}. This suggests that there will be a suppressing factor more than 10410^{-4} for higher-band mixture, which is negligible in our systems.

Besides the phenomena of higher-band mixture due to a large scattering length, the Floquet modulation may also cause the transitions to the excited bands. A particle at site-ii may absorb Δnp\Delta n_{p} Floquet ”photons”, and be excited to a higher band. The effective coupling strength QΔnpQ_{\Delta n_{p}} to excite a particle to higher bands via a resonant Δnp\Delta n_{p}-photon process is (see Appendix D for detailed calculations)

QΔnp=BΔnp(z0z0,th)Δnp1.Q_{\Delta n_{p}}=B_{\Delta n_{p}}(\frac{z_{0}}{z_{0,th}})^{\Delta n_{p}-1}. (5)

Here BΔnpB_{\Delta n_{p}} is a coefficient whose magnitude has weak dependence on Δnp\Delta n_{p} and we list the detailed form in Appendix D. The threshold z0,thz_{0,th} is a dimensionless quantity which is approximately the photon number Δnp\Delta n_{p} divided by the Euler’s number ee. In the case of VD=28ErV_{D}=28E_{r}, it requires at least 28 photons to excite a particle to the higher bands so z0,thz_{0,th} is around 10, while z0z_{0} in our proposal is 2.3. Then (z0z0,th)Δnp1(\frac{z_{0}}{z_{0,th}})^{\Delta n_{p}-1} in the equation provides a factor of 101910^{-19}. Therefore, the interband heating caused by Floquet modulation is negligible in our scenario.

V Adiabatic preparation of FQH states

Another potential question about our model is whether it’s appropriate to derive effective Hamiltonian by the high-frequency-expansion method. To eliminate this concern and prove that we can prepare such FQH states adiabatically in cold-atom platforms, in the following calculations we apply the original time-dependent Hamiltonian, which contains intrinsic long-range hoppings and does not include Floquet treatments. The original time-dependent Hamiltonian is in the form of

H(t)=\displaystyle H(t)=- [\displaystyle\Big{[} <i,j>eizijsin(Ωt+ϕij)t0c^ic^j\displaystyle\sum_{<i,j>}e^{iz_{ij}\sin(\Omega t+{\phi}_{ij})}t_{0}\hat{c}^{\dagger}_{i}\hat{c}_{j} (6)
+\displaystyle+ <i,j>2eizijsin(Ωt+ϕij)t1c^ic^j\displaystyle\sum_{<i,j>_{2}}e^{iz_{ij}\sin(\Omega t+{\phi}_{ij})}t_{1}\hat{c}^{\dagger}_{i}\hat{c}_{j}
+\displaystyle+ <i,j>3eizijsin(Ωt+ϕij)t2c^ic^j\displaystyle\sum_{<i,j>_{3}}e^{iz_{ij}\sin(\Omega t+{\phi}_{ij})}t_{2}\hat{c}^{\dagger}_{i}\hat{c}_{j}
+\displaystyle+ <i,j>4eizijsin(Ωt+ϕij)t3c^ic^j+\displaystyle\sum_{<i,j>_{4}}e^{iz_{ij}\sin(\Omega t+{\phi}_{ij})}t_{3}\hat{c}^{\dagger}_{i}\hat{c}_{j}+...
+\displaystyle+ h.c.]+i,jVi,jn^in^j.\displaystyle h.c.\Big{]}+\sum_{\langle i,j\rangle}V_{i,j}\hat{n}_{i}\hat{n}_{j}.

Here t0t_{0}, t1t_{1}, t2t_{2} and t3t_{3} are intrinsic NN, NNN, NNNN, and NNNNN hopping coefficients for static honeycomb optical lattices (Table I). The on-site interaction UU does not appear in the expression since we limit the Hilbert space into the states where each site can be filled with one particle at most. zijz_{ij} and ϕij\phi_{ij} have the same meaning as those in Eq. 2 and have different values for different hoppings.

Therefore, we simulate how to prepare the target FQH state, which is the ground state of the effective Hamiltonian under the twisted boundary condition, with FQH conductance of 1/2 . Initially, the lattice is static and unshaken, and we start with the ground state |ψ(z0=0)|\psi(z_{0}=0)\rangle under twisted boundary conditions. Then we fix the shaking frequency Ω\Omega at 0.108 and linearly ramp the shaking amplitude AA to increase z0z_{0} from 0 to 2.3 in CC Floquet-modulation cycles, and we calculate the fidelity, which is the module square of inner products, between the time-evolved state and the target FQH state.

In Fig. 3a, we plot the fidelity versus z0z_{0} and different ramping rates (or total modulation cycles CC). The fidelity at the end of ramping reaches over 85% for both cycles C=300C=300 and 400400, and approaches to a constant while C>400C>400. We find if we ignore t~3\tilde{t}_{3} (NNNNN hoppings) in the effective Hamiltonian, the fidelity drops below 80%. Although t~3\tilde{t}_{3} is only 1/30 of other major terms, it hurts the calculations of fidelities. It suggests that the ground state of the effective Hamiltonian is not identical to the experimentally-prepared state by adiabatic ramping, and this causes the fidelity to fall below 100%.

Refer to caption
Figure 3: (a) Fidelity versus z0z_{0} and CC. The inset is a zoom-in plot to show the difference near z0=2.3z_{0}=2.3. Slowly turning on the Floquet modulation helps the fidelity, and 400 modulation cycles will be slow enough to reach the adiabatic limit. (b) The spectrum of energy EnE_{n} versus θx\theta_{x} for energy-lowest eight states of the effective Hamiltonian. The energy is normalized by t0t_{0} and θx\theta_{x} is changed from 0 to 4π4\pi. Two lowest states share the Chern number 1 together, and it is a typical signature of fractional quantum Hall conductance at 1/21/2. (c) Fidelity within the manifold of the lowest two states while changing θx\theta_{x}. Green belt for the adiabatic-evolved state; red dotted line for the ground state of the effective Hamiltonian; blue solid line for the ground state of a topologically trivial Hamiltonian.

Following the discussion of FQH conductance in Ref. Niu et al. (1985), the ground state and the first excited state cross each other in the many-body energy spectrum while they are isolated from the higher excited states. Besides this level isolation, one of the 1/2-FQH states evolves into the other one when the boundary condition θx\theta_{x} is changed by 2π2\pi. Then, 4π4\pi instead of 2π\pi becomes the period for the state to evolve back to itself. In Fig. 3b, we show that the states of the effective Hamiltonian have a 4π4\pi-period and the level isolation.

Therefore, for the adiabatic-evolved states without tight-binding approximations, we apply the same arguments to prove the FQH conductance. We use State I to denote the adiabatic-evolved state during the preparation. We tune θx\theta_{x} by 2π2\pi very slowly, and State I evolves and cannot come back to itself since the gap is closed during the change of boundary conditions. We extract the orthogonal component at 2π2\pi change and refer it as State II. This state maintains a good fidelity (always above 75%) with the first excited state of the effective Hamiltonian.

We change θx\theta_{x} for multiple 2π2\pi (Fig. 3c) and calculate the overall fidelity within this manifold of State I and II for the actual scenario. In Fig. 3c, we show how the fidelity changes while the boundary condition is being varied (green belts). When θx\theta_{x} leaves 0 or 2π2\pi, the overlap between State I and State II decreases since the state is evolving. When θx\theta_{x} reaches integer times of 2π2\pi, the fidelity comes back again. After a few 2π2\pi, the fidelity is stably oscillating and does not decay while θx\theta_{x} is varied. The shape of belts is due to rapid Floquet modulations. Within each modulation, the fidelity is oscillating and the long-time-scale fidelity behaves as an area instead of a line. Here the fidelity does not exactly come back to 1 and we believe this is mainly due to the numerical errors of the giant matrix and data storage.

To convince this point, we numerically calculate another two cases for direct comparisons. One case is for the effective Hamiltonian under Floquet theory (red dotted line), and another is for a topological-trivial Hamiltonian (blue solid line). For the effective Hamiltonian, the change of fidelity behaves very similar to the case of adiabatic-evolved state. It is robust and does not decay while we continue changing θx\theta_{x}. Since we analytically solve this model, the ideal case should be that the fidelity reaches 1 again when θx\theta_{x} is integer times of 2π2\pi. The fidelity below 1 is due to the issues of numerical errors of a large-size matrix. The same reason applies to the case of the adiabatic-evolved state. Therefore, the robustness of the fidelity in the manifold of State I and State II proves the properties of FQH conductance. For the case of the topological-trivial Hamiltonian, we find the fidelity is decreasing while we change θx\theta_{x}. It indicates that the wave function is leaking out from this manifold. There is no gap separating the energy-lowest two states with the other bands, which cannot support FQH conductance.

VI Conclusions

In conclusion, we find a scheme realizing the bosonic FQH conductance of 1/2 in optical lattices without flat band. By circularly shaking the optical lattice and applying Feshbach resonance, the system displays a fractional quantum Hall conductance of 1/2. We show that the state with this conductance can be experimentally prepared by adiabatic turning on the lattice shaking. It provides a convenient and robust way to investigate the bosonic FQH effect Kjäll and Moore (2012); Wen (1990); Estienne et al. (2015); Repellin and Goldman (2019); Price and Cooper (2012); Račiūnas et al. (2018) in cold-atom experiments.

We acknowledge financial supports from the National Natural Science Foundation of China under grant no. 11974202, and 92165203.

Appendix A Effective Floquet Hamiltonian

In this section, we derive the effective Floquet Hamiltonian for shaken optical lattices. The lattice moves along a trajectory 𝒓l(t)\bm{r}_{l}(t) where the subscript ll corresponds to the displacement of the whole lattice. In the lattice coordinate, there is an inertial force 𝑭(t)=m𝒓¨l(t)\bm{F}(t)=-m\bm{\ddot{r}}_{l}(t) applying to each atom whose mass is mm, and this leads to a site-dependent potential 𝑭(t)𝒓i-\bm{F}(t)\cdot\bm{r}_{i}. Therefore, the Hamiltonian in the lattice coordinate has a form of

H^(t)\displaystyle\hat{H}(t) =\displaystyle= i,jtijc^ic^ji(𝑭(t)𝒓i)c^ic^i\displaystyle-\sum_{i,j}\hbar t_{ij}\hat{c}^{\dagger}_{i}\hat{c}_{j}-\sum_{i}(\bm{F}(t)\cdot\bm{r}_{i})\hat{c}^{\dagger}_{i}\hat{c}_{i} (7)
+\displaystyle+ i,jVi,jn^in^j+12iUin^i(n^i1)\displaystyle\sum_{{\langle}i,j\rangle}V_{i,j}\hat{n}_{i}\hat{n}_{j}+{1\over 2}\sum_{i}U_{i}\hat{n}_{i}(\hat{n}_{i}-1)

where ii or jj may represent any lattice site while i,j\langle i,j\rangle corresponds to a nearest-neighbor pair of lattice sites. By introducing a unitary transformation

U^(t)=exp[ii(ma𝒓˙l(t)𝒓i)c^ic^i],\hat{U}(t)=\exp[{i\over\hbar}\sum_{i}(-m_{a}\bm{\dot{r}}_{l}(t)\cdot\bm{r}_{i})\hat{c}_{i}^{\dagger}\hat{c}_{i}], (8)

the Hamiltonian is converted into

H^(t)=U^(t)H^(t)U^(t)iU^(t)tU^(t).\hat{H}^{\prime}(t)=\hat{U}^{\dagger}(t)\hat{H}(t)\hat{U}(t)-i\hbar\ \hat{U}^{\dagger}(t)\partial_{t}\hat{U}(t). (9)

The first term on the right hand side leaves the inertial-force potential and interaction terms unchanged, and the second term cancels out the inertial-force potential. By applying the Baker-Compell-Hausdorff formula, the hopping terms after the transformation become

U^(t)(i,jtijc^ic^j)U^(t)=i,jtijc^ic^jeima𝒓˙l(t)𝒓ij,\hat{U}^{\dagger}(t)(-\sum_{i,j}\hbar t_{ij}\hat{c}^{\dagger}_{i}\hat{c}_{j})\hat{U}(t)=-\hbar\sum_{i,j}t_{ij}\ \hat{c}^{\dagger}_{i}\hat{c}_{j}e^{{i\over\hbar}m_{a}\bm{\dot{r}}_{l}(t)\cdot\bm{r}_{ij}}, (10)

where 𝒓ij=𝒓i𝒓j\bm{r}_{ij}=\bm{r}_{i}-\bm{r}_{j} is the relative position between the site ii and jj. And the Hamiltonian after the transformation is

H^(t)\displaystyle\hat{H}^{\prime}(t) =\displaystyle= i,jtijc^ic^jeima𝒓˙l(t)𝒓ij\displaystyle-\hbar\sum_{i,j}t_{ij}\ \hat{c}^{\dagger}_{i}\hat{c}_{j}e^{{i\over\hbar}m_{a}\bm{\dot{r}}_{l}(t)\cdot\bm{r}_{ij}} (11)
+\displaystyle+ i,jVi,jn^in^j+12iUin^i(n^i1).\displaystyle\sum_{\langle i,j\rangle}V_{i,j}\hat{n}_{i}\hat{n}_{j}+{1\over 2}\sum_{i}U_{i}\hat{n}_{i}(\hat{n}_{i}-1).

When the lattice moves along a circular trajectory with a radius AA, i.e. 𝒓l(t)=𝒓l,0A[cos(Ωt)𝒙^+sin(Ωt)𝒚^]\bm{r}_{l}(t)=\bm{r}_{l,0}-A[\cos(\Omega t)\bm{\hat{x}}+\sin(\Omega t)\bm{\hat{y}}]. Then we introduce the symbols ρij\rho_{ij} and ϕij\phi_{ij} to simplify the equation, whose definitions are ρijeiϕij=(ri,0rj,0)(x^+y^eiπ/2)\rho_{ij}e^{i\phi_{ij}}=(\vec{r}_{i,0}-\vec{r}_{j,0})\cdot(\hat{x}+\hat{y}e^{-i\pi/2}). Then H^(t)\hat{H}^{\prime}(t) is

H^(t)\displaystyle\hat{H}^{\prime}(t) =\displaystyle= i,jtijc^ic^jeimaAΩρijsin(Ωt+ϕij)\displaystyle-\hbar\sum_{i,j}t_{ij}\ \hat{c}^{\dagger}_{i}\hat{c}_{j}e^{{i\over\hbar}m_{a}A\Omega\rho_{ij}\sin(\Omega t+\phi_{ij})} (12)
+\displaystyle+ i,jVi,jn^in^j+12iUin^i(n^i1).\displaystyle\sum_{\langle i,j\rangle}V_{i,j}\hat{n}_{i}\hat{n}_{j}+{1\over 2}\sum_{i}U_{i}\hat{n}_{i}(\hat{n}_{i}-1).

Then we use a dimensionless parameter zij=maAΩρij/z_{ij}=m_{a}A\Omega\rho_{ij}/\hbar to characterize the system. In the main text z0z_{0} is the nearest-neighbor zijz_{ij} and the higher order of zijz_{ij} can be derived based on this formula and the value of z0z_{0}. A periodic Hamiltonian can be Fourier-decomposed by the Jacobi-Anger method into H^(t)=nH^neinΩt\hat{H}^{\prime}(t)=\sum_{n\in\mathbb{Z}}\hat{H}_{n}e^{in\Omega t}, and the nn-th order Fourier term for H^(t)\hat{H}^{\prime}(t) is

H^n\displaystyle\hat{H}_{n} =\displaystyle= i,jeinΩtJn(zij)einϕijtijc^ic^j\displaystyle-\hbar\sum_{i,j}e^{in\Omega t}J_{n}(z_{ij})e^{in\phi_{ij}}t_{ij}\ \hat{c}^{\dagger}_{i}\hat{c}_{j} (13)
+\displaystyle+ δn0(i,jVi,jn^in^j+12iUin^i(n^i1)),\displaystyle\delta_{n0}(\sum_{\langle i,j\rangle}V_{i,j}\hat{n}_{i}\hat{n}_{j}+{1\over 2}\sum_{i}U_{i}\hat{n}_{i}(\hat{n}_{i}-1)){\color[rgb]{1,0,0},}

where Jn(x)J_{n}(x) is the nn-th Bessel function.

In the high frequency region, the time-dependent Hamiltonian can be approximated by a time-independent Floquet Hamiltonian H^fl\hat{H}_{fl} based on high frequency expansion method Goldman and Dalibard (2014); Eckardt and Anisimovas (2015); Rahav et al. (2003); Bukov et al. (2015) with a form of H^fl=H^0Ω+H^1Ω+H^2Ω+\hat{H}_{fl}=\hat{H}_{0\Omega}+\hat{H}_{1\Omega}+\hat{H}_{2\Omega}+\ldots. Here H^0Ω\hat{H}_{0\Omega}, H^1Ω\hat{H}_{1\Omega}, and H^2Ω\hat{H}_{2\Omega} are obtained via the commutation relations of H^n\hat{H}_{n}, i.e.

H^0Ω\displaystyle\hat{H}_{0\Omega} =H^0,\displaystyle=\hat{H}_{0}, (14)
H^1Ω\displaystyle\hat{H}_{1\Omega} =1Ωn=11n[H^n,H^n],\displaystyle=\frac{1}{\hbar\Omega}\sum_{n=1}^{\infty}\frac{1}{n}[\hat{H}_{n},\hat{H}_{-n}], (15)
H^2Ω\displaystyle\hat{H}_{2\Omega} =122Ω2n=11n2([H^n,H^0],H^n]+h.c.)\displaystyle=\frac{1}{2\hbar^{2}\Omega^{2}}\sum_{n=1}^{\infty}\frac{1}{n^{2}}([\hat{H}_{n},\hat{H}_{0}],\hat{H}_{-n}]+h.c.) (16)
+132Ω2n,n=11nn([H^n,[H^n,H^nn]]\displaystyle+\frac{1}{3\hbar^{2}\Omega^{2}}\sum_{n,n^{\prime}=1}^{\infty}\frac{1}{nn^{\prime}}([\hat{H}_{n},[\hat{H}_{n^{\prime}},\hat{H}_{-n-n^{\prime}}]]
[H^n,[H^n,H^n+n]]+h.c.).\displaystyle-[\hat{H}_{n},[\hat{H}_{-n^{\prime}},\hat{H}_{-n+n^{\prime}}]]+h.c.).

Considering an unshaken lattice with only nearest-neighbor (NN) hopping t0t_{0}, the commutator of two NN hopping produces a new hopping term with longer range, i.e. [c^ic^j,c^jc^k]=c^ic^k[\hat{c}^{\dagger}_{i}\hat{c}_{j},\hat{c}^{\dagger}_{j}\hat{c}_{k}]=\hat{c}^{\dagger}_{i}\hat{c}_{k}. Therefore, the Floquet Hamiltonian has nearest-neighbor (NN), next-nearest-neighbor (NNN), next-next-nearest-neighbor (NNNN), and next-next-next-nearest-neighbor (NNNNN) hoppings, with a form of.

H^fl/\displaystyle\hat{H}_{fl}/\hbar =\displaystyle=- i,jt~0c^ic^ji,j2t~1c^ic^ji,j3t~2c^ic^j\displaystyle\sum_{\langle i,j\rangle}\tilde{t}_{0}\hat{c}^{\dagger}_{i}\hat{c}_{j}-\sum_{\langle i,j\rangle_{2}}\tilde{t}_{1}\hat{c}^{\dagger}_{i}\hat{c}_{j}-\sum_{\langle i,j\rangle_{3}}\tilde{t}_{2}\hat{c}^{\dagger}_{i}\hat{c}_{j} (17)
i,j4t~3c^ic^j+h.c.,\displaystyle-\sum_{\langle i,j\rangle_{4}}\tilde{t}_{3}\hat{c}^{\dagger}_{i}\hat{c}_{j}+h.c.,

where \langle\cdot\rangle, 2\langle\cdot\rangle_{2}, 3\langle\cdot\rangle_{3}, and 4\langle\cdot\rangle_{4} correspond to the summations of the NN, NNN, NNNN, and NNNNN sites. The effective hopping amplitudes are

t~0\displaystyle\tilde{t}_{0} =\displaystyle= t0J0(z0)+2t03Ω2s=11s2Js(z0)J0(z0)Js(z0)×[2cos(2sπ/3)2cos(sπ/3)\displaystyle t_{0}J_{0}(z_{0})+\frac{2t^{3}_{0}}{\Omega^{2}}\sum_{s=1}^{\infty}\frac{1}{s^{2}}J_{s}(z_{0})J_{0}(z_{0})J_{-s}(z_{0})\times[2\cos(2s\pi/3)-2\cos(s\pi/3) (18)
+2cos(2sπ/3)3cos(sπ)+1]+4t033Ω2s,s=11ss{Js(z0)Js(z0)Jss(z0)×\displaystyle+2\cos(2s\pi/3)-3\cos(s\pi)+1]+\frac{4t^{3}_{0}}{3\Omega^{2}}\sum_{s,s^{\prime}=1}^{\infty}\frac{1}{ss^{\prime}}\{J_{s}(z_{0})J_{s^{\prime}}(z_{0})J_{-s-s^{\prime}}(z_{0})\times
[2cos(2sπ/3sπ/3)2cos(sπ+sπ/3)+2cos(2sπ/3sπ)\displaystyle[2\cos(2s\pi/3-s^{\prime}\pi/3)-2\cos(s\pi+s^{\prime}\pi/3)+2\cos(-2s\pi/3-s^{\prime}\pi)
2cos(sπ/3+sπ)+cos(sπ)cos(sπ+sπ)](ss)},\displaystyle-2\cos(s\pi/3+s^{\prime}\pi)+\cos(s^{\prime}\pi)-\cos(s\pi+s^{\prime}\pi)]-(s^{\prime}\rightarrow-s^{\prime})\},
t~1\displaystyle\tilde{t}_{1} =\displaystyle= 2it02Ωs=11sJs(z0)Js(z0)sin(sπ/3),\displaystyle-\frac{2it^{2}_{0}}{\Omega}\sum_{s=1}^{\infty}\frac{1}{s}J_{s}(z_{0})J_{-s}(z_{0})\sin(s\pi/3), (19)
t~2\displaystyle\tilde{t}_{2} =\displaystyle= 4t03Ω2s=11s2Js(z0)J0(z0)Js(z0)×[cos(2sπ/3)cos(sπ/3)]\displaystyle\frac{4t_{0}^{3}}{\Omega^{2}}\sum_{s=1}^{\infty}\frac{1}{s^{2}}J_{s}(z_{0})J_{0}(z_{0})J_{-s}(z_{0})\times[\cos(2s\pi/3)-\cos(s\pi/3)] (20)
+8t033Ω2s,s=11ss{Js(z0)Js(z0)Jss(z0)×[cos(2sπ/3+sπ/3)\displaystyle+\frac{8t_{0}^{3}}{3\Omega^{2}}\sum_{s,s^{\prime}=1}^{\infty}\frac{1}{ss^{\prime}}\{J_{s}(z_{0})J_{s^{\prime}}(z_{0})J_{-s-s^{\prime}}(z_{0})\times[\cos(2s\pi/3+s^{\prime}\pi/3)
cos(sπ/3sπ/3)](ss)},\displaystyle-\cos(s\pi/3-s^{\prime}\pi/3)]-(s^{\prime}\rightarrow-s^{\prime})\},
t~3\displaystyle\tilde{t}_{3} =\displaystyle= 2t03Ω2s=11s2Js(z0)J0(z0)Js(z0)×[1cos(sπ/3)]\displaystyle\frac{2t_{0}^{3}}{\Omega^{2}}\sum_{s=1}^{\infty}\frac{1}{s^{2}}J_{s}(z_{0})J_{0}(z_{0})J_{-s}(z_{0})\times[1-\cos(s\pi/3)] (21)
+4t033Ω2s,s=11ss{Js(z0)Js(z0)Jss(z0)×[cos(sπ/3)\displaystyle+\frac{4t_{0}^{3}}{3\Omega^{2}}\sum_{s,s^{\prime}=1}^{\infty}\frac{1}{ss^{\prime}}\{J_{s}(z_{0})J_{s^{\prime}}(z_{0})J_{-s-s^{\prime}}(z_{0})\times[\cos(-s^{\prime}\pi/3)
cos(sπ/3sπ/3)](ss)}.\displaystyle-\cos(-s\pi/3-s^{\prime}\pi/3)]-(s^{\prime}\rightarrow-s^{\prime})\}.

Here we have not taken the interaction terms into accounts. The influence of the NN interaction V1V_{1} on the Floquet Hamiltonian is proportional to 1Ω2{1\over\Omega^{2}} and produces number-dependent NNN hopping. In the high frequency region, The NNN hopping induced by V1V_{1} is less than one tenth of that in Eq. 13 so we neglect it in the calculation of effective hopping. This approximation affects the calculation of the ground state of effective Hamiltonian, so we simulate the wavefunction under exact time-dependent Hamiltonian in the main text and show that the negligence is reasonable. Because we are interested in the scenario of hard-core bosons (UU\rightarrow\infty), it requires a high energy cost for two particels to occupy the same site and the Floquet photons cannot provide such a large energy. In the region where the modulation frequency is much smaller than the energy scale of interaction, the high frequency expansion method is not applicable. In Appendix D, we show that it is safe to confine the Hilbert space in the subspace composed by single-occupation states despite the Floquet modulation. Therefore, the on-site interaction does not appear in the Hamiltonian.

Appendix B Wannier functions in honeycomb optical lattices

We calculate the ground-state Wannier function in honeycomb optical lattices in order to characterize the on-site and nearest-neighbor interactions. First, we calculate the Bloch functions |ψ1,𝒌|\psi_{1,\bm{k}}\rangle and |ψ2,𝒌|\psi_{2,\bm{k}}\rangle of the first and the second energy bands at a 50×5050\times 50 𝒌\bm{k}-point mesh. The Wannier function is the Fourier transform of Bloch function with a form of

|𝑹n=V(2π)3FBZ𝑑𝒌|ψn,𝒌ei𝒌𝑹|\bm{R}n\rangle={V\over(2\pi)^{3}}\int_{\textrm{FBZ}}d\bm{k}|\psi_{n,\bm{k}}\rangle e^{-i\bm{k}\cdot\bm{R}} (22)

where 𝑹\bm{R} is lattice vector, |𝑹n|\bm{R}n\rangle is the nn-th Wannier function in the primitive cell at 𝑹\bm{R}, the integral domain is the first Brillouin zone (FBZ).

The choice of Bloch function has a gauge freedom that permits the replacement of |ψn,𝒌|\psi_{n,\bm{k}}\rangle by eiθ(𝒌)|ψn,𝒌e^{i\theta(\bm{k})}|\psi_{n,\bm{k}}\rangle. Therefore, a good gauge should make the derivative of Bloch function |ψn,𝒌\nabla|\psi_{n,\bm{k}}\rangle well defined in the FBZ. If there are JJ bands having crossover with each other, the gauge freedom is generalized to a unitary transform Umn(𝒌)U_{mn}^{(\bm{k})} among JJ Bloch functions

|ψ~n𝒌=m=1J|ψm𝒌Umn(𝒌).|\tilde{\psi}_{n\bm{k}}\rangle=\sum_{m=1}^{J}|\psi_{m\bm{k}}\rangle U_{mn}^{(\bm{k})}. (23)

Here, the gauge is determined by the projection method Marzari and Vanderbilt (1997); Marzari et al. (2012). We use the ground state of the harmonic trap localized at AA- or BB- site of honeycomb cells as the trial functions |h1|h_{1}\rangle and |h2|h_{2}\rangle. We define a matrix A(𝒌)mn=ψm,𝒌|hnA(\bm{k})_{mn}=\langle\psi_{m,\bm{k}}|h_{n}\rangle, and the unitary transform Umn(𝒌)U_{mn}^{(\bm{k})} can be represented by the singular value decomposition of A(𝒌)=Us(𝒌)Ds(𝒌)Vs𝒌)A(\bm{k})=U_{s}(\bm{k})D_{s}(\bm{k})V^{\dagger}_{s}\bm{k}), where Us(𝒌)U_{s}(\bm{k}) and Vs(𝒌)V_{s}(\bm{k}) are unitary and Ds(𝒌)D_{s}(\bm{k}) is diagonal. Then, the gauge matrix Umn(𝒌)U_{mn}^{(\bm{k})} is Umn(𝒌)=Us(𝒌)Vs(𝒌)U_{mn}^{(\bm{k})}=U_{s}(\bm{k})V_{s}^{\dagger}(\bm{k}).

Then, we use the results above to calculate the required inputs of Wannier90 Pizzi et al. (2020) and optimize the output Wannier functions. We find that the spread of optimized Wannier functions reaches the minimal which verifies that the Wannier function obtained from the projection method is the maximally localized Wannier function (MLWF). It’s required that the MLWF should be real all over the space, and the maximal magnitude of imaginary parts in our result is 101310^{-13} less than the maximal values of real parts. It supports that the non-vanishing imaginary parts are caused by numerical precision and do not hurt our calculations.

In Fig. 4(a), the MLWF centered at AA site wA(r)w_{A}(r) is plotted in a logarithmic scale. In Fig. 4(b) and (c), the MLWFs along xx-axis are plotted, and they are compared with the trial functions hA(r)h_{A}(r) and hB(r)h_{B}(r). Both MLWFs and trial functions are normalized so that the squared-integral in the xyx-y plane is 1.

Refer to caption
Figure 4: (a) MLWF centered at A site. The trap depth VD=30ErV_{D}=30E_{r}. Here aa is the lattice constant corresponding to the distance between the nearest AA site and BB site. (b) Comparison of MLWFs and trial functions (Gaussian functions from harmonic traps). (c) A zoom-in plot of panel b for comparison of MLWFs and trial functions around 0. All the panels are plotted in the logarithmic scale for the wave function values.

The calculation of interactions in BH model requires three-dimensional Wannier function, so we assume a harmonic trap along zz-axis with a vibrational frequency ωz=2π×50\omega_{z}=2\pi\times 50 kHz. We use the ground state wave function wAw_{A}(z) of the harmonic trap along zz-axis for the third dimension. The integral contribution along zz-axis is 𝑑zwA(z)wA(z)wA(z)wA(z)=maωz\int dzw_{A}^{\dagger}(z)w_{A}^{\dagger}(z)w_{A}(z)w_{A}(z)=\sqrt{m_{a}\omega_{z}\over\hbar}, where AA corresponds to AA site. We use rubidium 85 as an example and the integral contribution is 4.16μm14.16\ \mu\textrm{m}^{-1}. In Table 2, we present the integral in the xyx-y plane and the corresponding interaction at the scattering length as=10000a0a_{s}=10000a_{0} (a0a_{0} is the Bohr radius). The integral for the on-site overlap is

I1=𝑑x𝑑ywA(x,y)wA(x,y)wA(x,y)wA(x,y).I_{1}=\int dxdyw_{A}^{\dagger}(x,y)w_{A}^{\dagger}(x,y)w_{A}(x,y)w_{A}(x,y). (24)

The integral of the NN overlap is

I2=𝑑zwA(x,y)wB(x,y)wB(x,y)wA(x,y).I_{2}=\int dzw_{A}^{\dagger}(x,y)w_{B}^{\dagger}(x,y)w_{B}(x,y)w_{A}(x,y). (25)

The hopping terms are numerically acquired by fitting the Bloch bands to the tight-binding model with higher hopping terms Ibañez Azpiroz et al. (2013). The results are listed in the main text.

Table 2: The overlap contributions in the xx-yy plane for on-site and nearest-neighbor versus the trap depth VDV_{D}. The integral of overlap I1I_{1} and I2I_{2} is normalized to the units of λ2\lambda^{-2}, where λ\lambda is the wavelength of lasers for optical lattices. The energy is normalized to the unit of the recoil energy Er=h2/(2mλ2)=2k22maE_{r}=h^{2}/(2m\lambda^{2})=\frac{\hbar^{2}k^{2}}{2m_{a}}.
VDV_{D} (Er)(E_{r}) I1I_{1} (λ2)(\lambda^{-2}) I2I_{2} (λ2)(\lambda^{-2}) UU (Er)(E_{r}) V1/2V_{1}{/2} (Er)(E_{r})
20 25.9009 0.0123 71.34 0.0339
22 27.3995 0.0087 75.47 0.0240
24 28.8305 0.0062 79.41 0.0171
26 30.2013 0.0044 83.19 0.0121
28 31.5182 0.0032 86.82 0.0088
30 32.7867 0.0023 90.31 0.0063
32 34.0115 0.0017 93.68 0.0047

Appendix C Two-particle, Two-site and Two-band Model

For bosons in lattices, the field operator is

ψ^(𝒓)=m,𝑹ib^m,iwm(𝒓𝑹i),\hat{\psi}(\bm{r})=\sum_{m,\bm{R}_{i}}\hat{b}_{m,i}w_{m}(\bm{r}-\bm{R}_{i}), (26)

where b^m,i\hat{b}_{m,i} is the annihilation operator of bosons with site index ii and band index mm, wm(𝒓Ri)w_{m}(\bm{r}-R_{i}) is the corresponding Wannier function. Considering a delta-function potential, the general form of the lattice model is Zhai (2021)

H^=ijmtijmb^m,ib^m,j+12ijijmnmnUijklmnmnb^m,ib^m,ib^n,jb^n,j,\hat{H}=-\sum_{ijm}t_{ij}^{m}\hat{b}^{\dagger}_{m,i}\hat{b}_{m,j}+{1\over 2}\sum_{iji^{\prime}j^{\prime}}^{mnm^{\prime}n^{\prime}}U_{ijkl}^{mnm^{\prime}n^{\prime}}\hat{b}^{\dagger}_{m,i}\hat{b}^{\dagger}_{m^{\prime},i^{\prime}}\hat{b}_{n,j}\hat{b}_{n^{\prime},j^{\prime}}, (27)

where

tijm=d3𝒓wm(𝒓𝑹i)(222m+Vop(𝒓))wm(𝒓𝑹j),t_{ij}^{m}=-\int d^{3}\bm{r}w_{m}(\bm{r}-\bm{R}_{i})\left(-\frac{\hbar^{2}\nabla^{2}}{2m}+V_{op}(\bm{r})\right)w_{m}(\bm{r}-\bm{R}_{j}), (28)

and

Uijklmnmn\displaystyle U_{ijkl}^{mnm^{\prime}n^{\prime}} =4π2asmad3𝒓\displaystyle={4\pi\hbar^{2}a_{s}\over m_{a}}\int d^{3}\bm{r}
wm(𝒓𝑹i)wn(𝒓𝑹j)wm(𝒓𝑹k)wn(𝒓𝑹l).\displaystyle w_{m}(\bm{r}-\bm{R}_{i})w_{n}(\bm{r}-\bm{R}_{j})w_{m^{\prime}}(\bm{r}-\bm{R}_{k})w_{n^{\prime}}(\bm{r}-\bm{R}_{l}).

Following the previous reference Viebahn et al. (2021), we use the two-particle, two-site, and two-band model to estimate the band mixing due to large scattering length. We calculate the Wannier functions of the p-bands and plot them In Fig. 5 for both AA- and BB-sites. Because pyp_{y}-orbitals are anti-symmetric in yy-direction and more localized along xx-direction, the major contributions of band mixing with s-bands come from the pxp_{x}-orbital. Therefore, we keep the ss-band and pxp_{x}-band in AA- and BB-sites to estimate the band mixture. The lattice Hamiltonian is further simplified into

H\displaystyle H =\displaystyle= tABg(b^g,Ab^g,B+b^g,Bb^g,A)tABe(b^e,Ab^e,B+b^e,Bb^e,A)\displaystyle-t^{g}_{AB}(\hat{b}^{\dagger}_{g,A}\hat{b}_{g,B}+\hat{b}^{\dagger}_{g,B}\hat{b}_{g,A})-t^{e}_{AB}(\hat{b}^{\dagger}_{e,A}\hat{b}_{e,B}+\hat{b}^{\dagger}_{e,B}\hat{b}_{e,A}) (30)
tAAg(b^g,Ab^g,A+b^g,Bb^g,B)tAAe(b^e,Ab^e,A+b^e,Bb^e,B)\displaystyle-t^{g}_{AA}(\hat{b}^{\dagger}_{g,A}\hat{b}_{g,A}+\hat{b}^{\dagger}_{g,B}\hat{b}_{g,B})-t^{e}_{AA}(\hat{b}^{\dagger}_{e,A}\hat{b}_{e,A}+\hat{b}^{\dagger}_{e,B}\hat{b}_{e,B})
+12ijij{A,B}mnmn{e,g}Uijklmnmnb^m,ib^m,ib^n,jb^n,j\displaystyle+{1\over 2}\sum_{iji^{\prime}j^{\prime}\in\{A,B\}}^{mnm^{\prime}n^{\prime}\in\{e,g\}}U_{ijkl}^{mnm^{\prime}n^{\prime}}\hat{b}^{\dagger}_{m,i}\hat{b}^{\dagger}_{m^{\prime},i^{\prime}}\hat{b}_{n,j}\hat{b}_{n^{\prime},j^{\prime}}

Here the index gg (ee) corresponds to the ss- (pxp_{x}-) band particles, and the index AA (BB) corresponds to the AA- (BB-) site. We set tAAg=0t^{g}_{AA}=0 as the zero energy point, and then tAAet^{e}_{AA} is characterizing the band separation between ss- and pp-bands. In Table 3, we list the typical numerical magnitudes in this model. Since we are interested in the effect on the ground state b^g,Ab^g,B|vac\hat{b}^{\dagger}_{g,A}\hat{b}^{\dagger}_{g,B}|vac\rangle, where |vac|vac\rangle is the vacuum state, we focus on the transitions from the ground state to the states with higher energy. In Table 4, all the possible first-order transitions are listed with the energy costs and the coupling strengths. The ratios between the coupling strengths and the energy cost are less than 1/100, so the excited populations due to Feshbach resonance are less than 10410^{-4}. Therefore, the band mixture due to large scattering length is highly suppressed by the off resonant coupling.

Refer to caption
Figure 5: pxp_{x}-orbital and pyp_{y}-orbital at AA- or BB-sites.
Table 3: The typical numerical values in the two-particle, two-site and two-band model at as=10000a0a_{s}=10000a_{0} and VD=28ErV_{D}=28E_{r}. The calculation is based on ss-orbital Wannier functions and pxp_{x}-orbital Wannier functions at AA and BB-sites. The Wannier function along zz-direction is the ground state wave function in a harmonic trap with a vibrational frequency at 50 kHz. Here all the parameters are in the unit of the recoil energy Er=2k22maE_{r}={\hbar^{2}k^{2}\over 2m_{a}}.
numeric value (ErE_{r})
tAAe8.271\hbar t_{AA}^{e}\approx-8.271 UABABegeg=0.6409U^{egeg}_{ABAB}=0.6409 UABABeggg=0.0212U^{eggg}_{ABAB}=0.0212
UAAAAgggg=86.82U_{AAAA}^{gggg}=86.82 UAAABgggg=0.2828U_{AAAB}^{gggg}=-0.2828
UAAAAegeg=34.6818U_{AAAA}^{egeg}=34.6818 UAAABeggg=0.2648U_{AAAB}^{eggg}=-0.2648
UABABeeee=2.1601U_{ABAB}^{eeee}=2.1601 UABABeegg=0.0676U_{ABAB}^{eegg}=-0.0676
UAAAAeeee=55.0342U_{AAAA}^{eeee}=55.0342 UAAABeegg=0.1704U^{eegg}_{AAAB}=0.1704
Table 4: The first-order transition channels.
transition energy cost coupling
b^g,Ab^g,B|vacb^g,Ab^e,B|vac\hat{b}_{g,A}^{\dagger}\hat{b}_{g,B}^{\dagger}|vac\rangle\rightarrow\hat{b}_{g,A}^{\dagger}\hat{b}_{e,B}^{\dagger}|vac\rangle tAAe+2UABABegeg-t_{AA}^{e}+2U^{egeg}_{ABAB} 2UABABeggg2U_{ABAB}^{eggg}
(or b^g,Bb^e,A|vac\hat{b}_{g,B}^{\dagger}\hat{b}_{e,A}^{\dagger}|vac\rangle)
b^g,Ab^g,B|vacb^g,Ab^g,A|vac\hat{b}_{g,A}^{\dagger}\hat{b}_{g,B}^{\dagger}|vac\rangle\rightarrow\hat{b}_{g,A}^{\dagger}\hat{b}_{g,A}^{\dagger}|vac\rangle UAAAAggggU^{gggg}_{AAAA} UAAABggggU^{gggg}_{AAAB}
(or b^g,Bb^g,B|vac\hat{b}_{g,B}^{\dagger}\hat{b}_{g,B}^{\dagger}|vac\rangle)
b^g,Ab^g,B|vacb^g,Ab^e,A|vac\hat{b}_{g,A}^{\dagger}\hat{b}_{g,B}^{\dagger}|vac\rangle\rightarrow\hat{b}_{g,A}^{\dagger}\hat{b}_{e,A}^{\dagger}|vac\rangle tAAe+2UAAAAegeg-t_{AA}^{e}+2U^{egeg}_{AAAA} 2UAAABeggg2U^{eggg}_{AAAB}
(or b^g,Bb^e,B|vac\hat{b}_{g,B}^{\dagger}\hat{b}_{e,B}^{\dagger}|vac\rangle)
b^g,Ab^g,B|vacb^e,Ab^e,B|vac\hat{b}_{g,A}^{\dagger}\hat{b}_{g,B}^{\dagger}|vac\rangle\rightarrow\hat{b}_{e,A}^{\dagger}\hat{b}_{e,B}^{\dagger}|vac\rangle 2tAAe+2UABABeeee-2t_{AA}^{e}+2U^{eeee}_{ABAB} 2UABABeegg2U^{eegg}_{ABAB}
b^g,Ab^g,B|vacb^e,Ab^e,A|vac\hat{b}_{g,A}^{\dagger}\hat{b}_{g,B}^{\dagger}|vac\rangle\rightarrow\hat{b}_{e,A}^{\dagger}\hat{b}_{e,A}^{\dagger}|vac\rangle 2tAAe+UAAAAeeee-2t_{AA}^{e}+U^{eeee}_{AAAA} UAAABeeggU_{AAAB}^{eegg}
(or b^e,Bb^e,B|vac\hat{b}_{e,B}^{\dagger}\hat{b}_{e,B}^{\dagger}|vac\rangle)

Appendix D Floquet Heating

In this section, we estimate the interband excitations caused by Floquet modulations. Since the energy scale for interband excitations is much larger than the modulation frequency, the high frequency expansion method is not applicable. The quasienergy operator and a set of new bases, combining stationary states and Floquet photons, are applied to solve this problem Eckardt and Anisimovas (2015). Although there are lots of interaction terms in Eq. LABEL:eqn:interaction, only the on-site interaction has larger energy than a Floquet photon. Therefore, we focus on the on-site interaction and ignore the nearest-neighbor interaction.

Similar to Eq. 7, the complete tight-binding Hamiltonian in the moving reference frame is

H^c(t)\displaystyle\hat{H}_{c}(t) =\displaystyle= i,j,m,m(𝑭(t)𝒓i,jmm)c^i,mc^j,m\displaystyle-\sum_{i,j,m,m^{\prime}}(\bm{F}(t)\cdot\bm{r}_{i,j}^{m^{\prime}m})\hat{c}^{\dagger}_{i,m^{\prime}}\hat{c}_{j,m}
\displaystyle- i,j,mtijmc^i,mc^j,m+12i,mUiiiimmmmn^i,m(n^i,m1).\displaystyle\sum_{i,j,m}\hbar t_{ij}^{m}\hat{c}^{\dagger}_{i,m}\hat{c}_{j,m}+{1\over 2}\sum_{i,m}U_{iiii}^{mmmm}\hat{n}_{i,m}(\hat{n}_{i,m}-1).

Here mm is band index, tiimt_{ii}^{m} is the mean energy of the mm-th band, and 𝒓i,jmm=d3𝒓wm(𝒓𝑹i)𝒓wm(𝒓𝑹j)\bm{r}_{i,j}^{m^{\prime}m}=\int d^{3}\bm{r}w_{m^{\prime}}(\bm{r}-\bm{R}_{i})\ \bm{r}\ w_{m}(\bm{r}-\bm{R}_{j}). If only the contributions from the s-bands are taken into account, 𝒓i,jss\bm{r}_{i,j}^{ss} will equal to δij𝒓i\delta_{ij}\bm{r}_{i} where 𝒓i\bm{r}_{i} is the position of the lattice site ii in the moving reference frame. Then we obtain the same results as from Eq. 7.

For interband transitions, there is relation of 𝒓i,imm𝒓i,j(i)mm\bm{r}_{i,i}^{mm^{\prime}}\gg\bm{r}_{i,j(\neq i)}^{mm^{\prime}}.Therefore, by inspecting the first term in the right hand side of Eq. LABEL:eq:D1, the excitation by Floquet photons is mainly from the in situ transitions where the particle is still in the same spatial position but jumps to a higher band. Then we will focus on this contribution and demonstrate that it is not a concern. For simplicity, the repeated subscripts or superscripts will be contracted, e.g. ri,im,mrimr^{m,m}_{i,i}\rightarrow r^{m}_{i}. The Hamiltonian is simplified to

H^c(t)\displaystyle\hat{H}_{c}(t) =\displaystyle= i,j,mtijmc^i,mc^j,mi,m,m(𝑭(t)𝒓imm)c^i,mc^i,m+12i,mUimn^i,m(n^i,m1)\displaystyle-\sum_{i,j,m}\hbar t_{ij}^{m}\hat{c}^{\dagger}_{i,m}\hat{c}_{j,m}-\sum_{i,m,m^{\prime}}(\bm{F}(t)\cdot\bm{r}_{i}^{m^{\prime}m})\hat{c}^{\dagger}_{i,m^{\prime}}\hat{c}_{i,m}+{1\over 2}\sum_{i,m}U_{i}^{m}\hat{n}_{i,m}(\hat{n}_{i,m}-1) (32)
=\displaystyle= i,j,mtijmc^i,mc^j,mΩz0i,m,m[cos(Ωt)ximmasin(Ωt)yimma]c^i,mc^i,m+12i,mUimn^i,m(n^i,m1).\displaystyle-\sum_{i,j,m}\hbar t_{ij}^{m}\hat{c}^{\dagger}_{i,m}\hat{c}_{j,m}-{\hbar\Omega z_{0}}\sum_{i,m,m^{\prime}}[\cos(\Omega t)\frac{x_{i}^{m^{\prime}m}}{a}-\sin(\Omega t)\frac{y_{i}^{m^{\prime}m}}{a}]\hat{c}^{\dagger}_{i,m^{\prime}}\hat{c}_{i,m}+{1\over 2}\sum_{i,m}U_{i}^{m}\hat{n}_{i,m}(\hat{n}_{i,m}-1).

By introducing a unitary transformation

U^c(t)=exp[ii,m(ma𝒓˙l(t)𝒓im)c^i,mc^i,m],\hat{U}_{c}(t)=\exp[{i\over\hbar}\sum_{i,m}(-m_{a}\bm{\dot{r}}_{l}(t)\cdot\bm{r}_{i}^{m})\hat{c}_{i,m}^{\dagger}\hat{c}_{i,m}], (33)

the Hamiltonian is converted to

H^c(t)\displaystyle\hat{H}_{c}^{\prime}(t) =\displaystyle= i,j,mtijmc^i,mc^j,meimaAΩρijsin(Ωt+ϕij)+12i,mUimn^i,m(n^i,m1)\displaystyle-\hbar\sum_{i,j,m}t_{ij}^{m}\ \hat{c}^{\dagger}_{i,m}\hat{c}_{j,m}e^{{i\over\hbar}m_{a}A\Omega\rho_{ij}\sin(\Omega t+\phi_{ij})}+{1\over 2}\sum_{i,m}U_{i}^{m}\hat{n}_{i,m}(\hat{n}_{i,m}-1) (34)
\displaystyle- Ωz0i,mm(cos(Ωt)ximmasin(Ωt)yimma)c^i,mc^i,mei[ma𝒓˙l(t)(𝒓im𝒓im)].\displaystyle{\hbar\Omega z_{0}}\sum_{i,m\neq m^{\prime}}(\cos(\Omega t)\frac{x_{i}^{m^{\prime}m}}{a}-\sin(\Omega t)\frac{y_{i}^{m^{\prime}m}}{a})\hat{c}^{\dagger}_{i,m^{\prime}}\hat{c}_{i,m}e^{{i\over\hbar}[m_{a}\bm{\dot{r}}_{l}(t)\cdot(\bm{r}_{i}^{m^{\prime}}-\bm{r}_{i}^{m})]}.

Based on the symmetry of honeycomb lattices, the centers of Wannier functions wm(𝒓𝑹i)w_{m}(\bm{r}-\bm{R}_{i}) along yy-direction for all bands are the same as the yy-center of the lattice site ii. However, the centers of Wannier functions wm(𝒓𝑹i)w_{m}(\bm{r}-\bm{R}_{i}) along xx-direction are not the same as the xx-center of the lattice site. We define the dimensionless parameters ηmmx=ximma\eta_{m^{\prime}m}^{x}={x_{i}^{m^{\prime}m}\over a} and ηmmy=yimma\eta_{m^{\prime}m}^{y}={y_{i}^{m^{\prime}m}\over a} to further simplify H^c(t)\hat{H}_{c}^{\prime}(t) to

H^c\displaystyle\hat{H}_{c} (t)=i,j,mtijmc^i,mc^j,meiz0sin(Ωt+ϕij)Ωz0i,mm{}^{\prime}(t)=-\hbar\sum_{i,j,m}t_{ij}^{m}\ \hat{c}^{\dagger}_{i,m}\hat{c}_{j,m}e^{{i}z_{0}\sin(\Omega t+\phi_{ij})}-{\hbar\Omega z_{0}}\sum_{i,m\neq m^{\prime}} (35)
[cos(Ωt)ηmmxsin(Ωt)ηmmy]c^i,mc^i,meiz0Δlimmsin(Ωt)\displaystyle[\cos(\Omega t)\eta_{m^{\prime}m}^{x}-\sin(\Omega t)\eta_{m^{\prime}m}^{y}]\hat{c}^{\dagger}_{i,m^{\prime}}\hat{c}_{i,m}e^{iz_{0}{\Delta l_{i}^{m^{\prime}m}}\sin(\Omega t)}
+12i,mUimn^i,m(n^i,m1).\displaystyle\ \ +{1\over 2}\sum_{i,m}U_{i}^{m}\hat{n}_{i,m}(\hat{n}_{i,m}-1).

Here Δlimm=ximxima\Delta l_{i}^{m^{\prime}m}={x_{i}^{m^{\prime}}-x_{i}^{m}\over a} is characterizing the difference between the centers of mm and mm^{\prime} Wannier functions.

According to the Floquet theory, the solution of Schrödinger equation idt|ψ(t)=H^c(t)|ψ(t)i\hbar d_{t}|\psi(t)\rangle=\hat{H}_{c}^{\prime}(t)|\psi(t)\rangle has a form of |ψν(t)=|uν(t)eitϵν|\psi_{\nu}(t)\rangle=|u_{\nu}(t)\rangle e^{-{i\over\hbar}t\epsilon_{\nu}}, where |uν(t)|u_{\nu}(t)\rangle is a periodic function with a period of T=2π/ΩT=2\pi/\Omega. |ψ(t)ν|\psi(t)_{\nu}\rangle is called the Floquet state, |uν(t)|u_{\nu}(t)\rangle is the Floquet mode, and ϵν\epsilon_{\nu} is the quasienergy. |ψν(t)|\psi_{\nu}(t)\rangle is also an eigenstate of the time-evolution operator in one period TT, i.e.,

U^(t0+T,t0)|ψν(t0)=eiTϵν|ψν(t0),\hat{U}(t_{0}+T,t_{0})\ |\psi_{\nu}(t_{0})\rangle=e^{-{i\over\hbar}T\epsilon_{\nu}}|\psi_{\nu}(t_{0})\rangle, (36)

where U^(t0+T,t0)\hat{U}(t_{0}+T,t_{0}) denotes the time-evolution operator from t0t_{0} to t0+Tt_{0}+T and the eigenvalue eiTϵνe^{-{i\over\hbar}T\epsilon_{\nu}} does not depend on the start time t0t_{0}. By solving the eigenvalue problem of the time-evolution operator, the phase factor eiTϵνe^{-{i\over\hbar}T\epsilon_{\nu}} and the Floquet state |ψν(t)|\psi_{\nu}(t)\rangle are uniquely defined, while the corresponding quasienergies and Floquet modes are not unique. A Floquet state can be written as |ψν(t)=|uνnp(t)eitϵνnp|\psi_{\nu}(t)\rangle=|u_{\nu n_{p}}(t)\rangle e^{-{i\over\hbar}t\epsilon_{\nu n_{p}}}, where ϵνnp=ϵν+npΩ\epsilon_{\nu n_{p}}=\epsilon_{\nu}+n_{p}\hbar\Omega and |uνnp(t)=|uν(t)einpΩt|u_{\nu n_{p}}(t)\rangle=|u_{\nu}(t)\rangle e^{in_{p}\Omega t}. For a particular ν\nu, there are a series of orthogonal Floquet modes |uνnp(t)|u_{\nu n_{p}}(t)\rangle with quasienergies ϵνnp\epsilon_{\nu n_{p}}. The quasienergies and the Floquet modes are the eigenstates and eigenenergies of quasienergy operator Q^(t)=H^c(t)idt\hat{Q}(t)=\hat{H}_{c}^{\prime}(t)-i\hbar d_{t}, i.e.,

Q^|uνnp=ϵνnp|uνnp.\hat{Q}|u_{\nu n_{p}}\rangle\rangle=\epsilon_{\nu n_{p}}|u_{\nu n_{p}}\rangle\rangle. (37)

Here the time-dependent state |u(t)|u(t)\rangle with a period TT is written as a double-ket |u|u\rangle\rangle. The scalar product for such a state is given by u|v=1T0T𝑑tu(t)|v(t)\langle\langle u|v\rangle\rangle={1\over T}\int_{0}^{T}dt\ \langle u(t)|v(t)\rangle. Similar to spatially periodic Hamiltonians, one can fix all quasienergies in the same interval of width ω\hbar\omega, called a Brillouin zone. Therefore, all Floquet states |ψν(t)|\psi_{\nu}(t)\rangle can be constructed from the Floquet modes whose quasienergies lie in a single Brillouin zone.

For the driven optical lattices, a useful set of bases is

|m,i,np=c^i,m|vaceinpΩt.|m,i,n_{p}\rangle\rangle=\hat{c}_{i,m}^{\dagger}|vac\rangle e^{in_{p}\Omega t}. (38)

Here npn_{p} is the number of Floquet photons. Then the matrix elements of the quasienergy operator Q^\hat{Q} is

m,i,np|Q^|m,i,np=m,i|H^c,npnp+δnpnpnpΩ|m,i,\langle\langle m^{\prime},i^{\prime},n^{\prime}_{p}|\hat{Q}|m,i,n_{p}\rangle\rangle=\langle m^{\prime},i^{\prime}|\hat{H}^{\prime}_{c,n^{\prime}_{p}-n_{p}}+\delta_{n^{\prime}_{p}n_{p}}n_{p}\hbar\Omega|m,i\rangle, (39)

where H^c,n\hat{H}^{\prime}_{c,n} is obtained by the Fourier transformation of H^c(t)=nH^c,seisΩt\hat{H}^{\prime}_{c}(t)=\sum_{n}\hat{H}^{\prime}_{c,s}e^{is\Omega t} with a form of

H^\displaystyle\hat{H} =c,si,j,mtijmJs(z0)eisϕijc^i,mc^j,mΩz0i,mm{}^{\prime}_{c,s}=-\sum_{i,j,m}t_{ij}^{m}J_{s}(z_{0})e^{is\phi_{ij}}\hat{c}_{i,m}^{\dagger}\hat{c}_{j,m}-\hbar\Omega z_{0}\sum_{i,m\neq m^{\prime}} (40)
[ηmm+Js1(z0Δlimm)+ηmmJs+1(z0Δlimm)]c^i,mc^i,m.\displaystyle\left[\eta_{m^{\prime}m}^{+}J_{s-1}(z_{0}\Delta l_{i}^{m^{\prime}m})+\eta_{m^{\prime}m}^{-}J_{s+1}(z_{0}\Delta l_{i}^{m^{\prime}m})\right]\hat{c}^{\dagger}_{i,m^{\prime}}\hat{c}_{i,m}.
+12δs0i,mUimn^i,m(n^i,m1)\displaystyle\ \ +{1\over 2}\delta_{s0}\sum_{i,m}U_{i}^{m}\hat{n}_{i,m}(\hat{n}_{i,m}-1)

Here ηmm+\eta_{m^{\prime}m}^{+} (ηmm\eta_{m^{\prime}m}^{-}) corresponds to ηmmx+iηmmy2{\eta^{x}_{m^{\prime}m}+i\eta^{y}_{m^{\prime}m}\over 2} (ηmmxiηmmy2{\eta^{x}_{m^{\prime}m}-i\eta^{y}_{m^{\prime}m}\over 2}). In Table 5, we list the related numeric values of η\eta and Δl\Delta l for ss- and pp- bands.

Then we will estimate the resonant coupling strength of interband transitions via absorbing Δnp\Delta n_{p} Floquet photons. In our case, the Floquet photon energy Ω=t/0.108=0.287Er\Omega={t}/0.108=0.287E_{r} and the band gap tiip8.271Er-t_{ii}^{p}\approx 8.271E_{r}, so it requires around 28 photons to excite a particle from ss-band to pp-band. For a Δnp\Delta n_{p}-photon transition process where Δnp\Delta n_{p} is large enough, the coupling strength is

\displaystyle\langle\langle Px\displaystyle P_{x} ,i,np|Q^|S,i,np+ΔnpJΔnp1(z0Δlimm)\displaystyle,i,n_{p}|\hat{Q}|S,i,n_{p}+\Delta n_{p}\rangle\rangle\sim J_{\Delta n_{p}-1}(z_{0}\Delta l_{i}^{m^{\prime}m}) (41)
=\displaystyle= k=0(1)kk!(k+Δnp)!(z0Δlimm2)2k+Δnp1\displaystyle\sum_{k=0}^{\infty}{(-1)^{k}\over k!(k+\Delta n_{p})!}\left(\frac{z_{0}\Delta l_{i}^{m^{\prime}m}}{2}\right)^{2k+\Delta n_{p}-1}
\displaystyle\sim 1Δnp!(z0Δlimm2)Δnp1\displaystyle\frac{1}{\Delta n_{p}!}\left(\frac{z_{0}\Delta l_{i}^{m^{\prime}m}}{2}\right)^{\Delta n_{p}-1}
\displaystyle\sim (ez0Δlimm2Δnp)Δnp1=(z0zth)Δnp1.\displaystyle\left(e\frac{z_{0}\Delta l_{i}^{m^{\prime}m}}{2\Delta n_{p}}\right)^{\Delta n_{p}-1}=\left(\frac{z_{0}}{z_{th}}\right)^{\Delta n_{p}-1}.

Here we apply Stirling formula (n)!=2πn(ne)n(n)!=\sqrt{2\pi n}({n\over e})^{n} to the factorial. Because Δlimm\Delta l_{i}^{m^{\prime}m} is less than 1 between all bands, this suggests zth>2Δnpez_{th}>2\frac{\Delta n_{p}}{e}. When z0z_{0} is smaller than zthz_{th}, the coupling strength is exponentially suppressed.

Besides absorbing Δnp\Delta n_{p} photons directly, the Δnp\Delta n_{p}-order transition to the target state via Δnp1\Delta n_{p}-1 intermediate states may also heat the system. For the Δnp\Delta n_{p}-order transition, the particle absorbs a single photon Δnp\Delta n_{p} times. For each time the coupling strength is

Px,i,np|\displaystyle\langle\langle P_{x},i,n_{p}| Q^\displaystyle\hat{Q} |S,i,np+1\displaystyle|S,i,n_{p}+1\rangle\rangle (42)
\displaystyle\approx Ωz0(ηmmx+iηmmy2)J0(z0Δlimm)\displaystyle\hbar\Omega z_{0}\left({\eta^{x}_{m^{\prime}m}+i\eta^{y}_{m^{\prime}m}\over 2}\right)J_{0}(z_{0}\Delta l_{i}^{m^{\prime}m})
<\displaystyle< Ωz0(ηmmx+iηmmy2).\displaystyle\hbar\Omega z_{0}\left({\eta^{x}_{m^{\prime}m}+i\eta^{y}_{m^{\prime}m}\over 2}\right).

The coupling for the Δnp\Delta n_{p}-order process is Px,i,np|Q^|S,i,np+1Δnp\langle\langle P_{x},i,n_{p}|\hat{Q}|S,i,n_{p}+1\rangle\rangle^{\Delta n_{p}} divided by the product of all energetic detuning of intermediate states. According to the discussion on high order transition processes in Ref. Sträter and Eckardt (2016); Weinberg et al. (2015), this product has the same order of magnitude as 1(Δnp1)!(Ω)Δnp1\frac{1}{(\Delta n_{p}-1)!}(\hbar\Omega)^{\Delta n_{p}-1}. Therefore, the coupling term for a Δnp\Delta n_{p}-order process also behaves as (z0Δlimmzth)Δnp1\left({z_{0}\Delta l_{i}^{m^{\prime}m}\over z_{th}}\right)^{\Delta n_{p}-1}. For a harmonic trap, the dipole matrix element is nonzero only for two states whose difference of the vibrational energy level is 1. Based on Table 5, the dimensionless dipole matrix elements are less than 1, so the threshold zthz_{th} for Δnp\Delta n_{p}-order process is larger than Δnp1e\Delta n_{p}-1\over e.

Table 5: Numeric value relevant to Floquet heating.
ηpxsx\eta_{p_{x}s}^{x} 0.1292 Δlipxs\Delta l_{i}^{p_{x}s} 0.0891
ηpysx\eta_{p_{y}s}^{x} 0 Δlipys\Delta l_{i}^{p_{y}s} 0
ηpxsy\eta_{p_{x}s}^{y} 0
ηpysy\eta_{p_{y}s}^{y} 0.1292

In our case, the target value of the modulation parameter z0z_{0} is 2.3, which is less than the threshold, so the interband excitation caused by Floquet modulations is exponentially suppressed and negligible.

In the derivation of Eq. 17, the on-site interaction is neglected. Although the modulation couple the single-occupation state gAgB|vacg^{\dagger}_{A}g^{\dagger}_{B}|vac\rangle to the doublon state gAgA|vacg_{A}^{\dagger}g_{A}^{\dagger}|vac\rangle, the coupling strength is suppressed by JΔnp(z0)J_{\Delta n_{p}}(z_{0}) where Δnp\Delta n_{p} is over 100. Therefore, it’s safe to apply single-occupation state space under the Floquet modulation.

Appendix E Many-body Hamiltonian

In this section, we use the case of 24 lattice sites (3×4×23\times 4\times 2, see Fig. 6) and 6 particles as an example to illustrate how to enumerate Fock states and write down the many-body Hamiltonian for the exact- diagonalization calculation.

First, we give all sites a serial number, which are illustrated in Fig. 6. cnc_{n}^{\dagger} (cnc_{n}) is the creation (annihilation) operator of a boson on the nn-th site, where n=0,1,,23n=0,1,...,23. For hard-core bosons, the basis is formed by {ci6ci5ci4ci3ci2ci1|vac}\{c_{i_{6}}^{\dagger}c^{\dagger}_{i_{5}}c^{\dagger}_{i_{4}}c^{\dagger}_{i_{3}}c^{\dagger}_{i_{2}}c^{\dagger}_{i_{1}}|\textrm{vac}\rangle\} where i1<i2<i3<i4<i5<i6{i_{1}<i_{2}<i_{3}<i_{4}<i_{5}<i_{6}} and |vac|\textrm{vac}\rangle is the vacuum state. The total number of base vectors is a binomial coefficient of sites and particles, which is 134596 in our case. We give all the occupation structure a serial number from 1 to 134596. The mapping rule is as follows

{ci6ci5ci4ci3ci2ci1|vac}i1<i2<i3<i4<i5<i6:\displaystyle\{c^{\dagger}_{i_{6}}c^{\dagger}_{i_{5}}c^{\dagger}_{i_{4}}c^{\dagger}_{i_{3}}c^{\dagger}_{i_{2}}c^{\dagger}_{i_{1}}|\textrm{vac}\rangle\}_{i_{1}<i_{2}<i_{3}<i_{4}<i_{5}<i_{6}}: \displaystyle\longrightarrow
n1=24i123n2=24i222i1n3=24i322i2n4=24i422i3n5=24i522i4\displaystyle\sum_{n_{1}=24-i_{1}}^{23}\sum_{n_{2}=24-i_{2}}^{22-i_{1}}\sum_{n_{3}=24-i_{3}}^{22-i_{2}}\sum_{n_{4}=24-i_{4}}^{22-i_{3}}\sum_{n_{5}=24-i_{5}}^{22-i_{4}}
(n15)(n24)(n33)(n42)(n51)+n6n5\displaystyle{n_{1}\choose 5}{n_{2}\choose 4}{n_{3}\choose 3}{n_{4}\choose 2}{n_{5}\choose 1}+n_{6}-n_{5} (43)
Refer to caption
Figure 6: 24 lattice sites for exact diagonalization and the enumeration rule.

Besides, the twisted boundary condition should also be satisfied. To achieve this boundary conditions, the creation operators should satisfy cmb|vac=eiθxc0|vacc^{\dagger}_{m_{b}}|\textrm{vac}\rangle=e^{i\theta_{x}}c^{\dagger}_{0}|\textrm{vac}\rangle and cnb|vac=eiθyc0|vacc^{\dagger}_{n_{b}}|\textrm{vac}\rangle=e^{i\theta_{y}}c^{\dagger}_{0}|\textrm{vac}\rangle where nbn_{b} and mbm_{b} are positions outside the zone in Fig. 6. Based on the twisted boundary condition, we map the tunneling out of this region back into the region of interest. We write out the many-body Hamiltonian based on this enumeration rule. To visualize the Hamiltonian, we mark the states connected by NN, NNN, NNNN, NNNN hopping terms respectively in Fig. 7.

Refer to caption
Figure 7: (a-d) The non-zero elements of the NN, NNN, NNNN, NNNNN contributions are marked respectively. nznz is the number of non-zero elements.

References

  • Stormer et al. (1999) Horst L. Stormer, Daniel C. Tsui,  and Arthur C. Gossard, “The fractional quantum Hall effect,” Rev. Mod. Phys. 71, S298–S305 (1999).
  • Cooper (2020) N.R. Cooper, “Fractional quantum Hall states of bosons: Properties and prospects for experimental realization,” in Fractional Quantum Hall Effects: New Developments (World Scientific Press, 2020) Chap. 10, pp. 487–521.
  • Wen (2017) Xiao-Gang Wen, “Colloquium: Zoo of quantum-topological phases of matter,” Rev. Mod. Phys. 89, 041004 (2017).
  • Bloch et al. (2008) Immanuel Bloch, Jean Dalibard,  and Wilhelm Zwerger, “Many-body physics with ultracold gases,” Rev. Mod. Phys. 80, 885–964 (2008).
  • Cooper et al. (2019) N. R. Cooper, J. Dalibard,  and I. B. Spielman, “Topological bands for ultracold atoms,” Rev. Mod. Phys. 91, 015005 (2019).
  • Sørensen et al. (2005) Anders S. Sørensen, Eugene Demler,  and Mikhail D. Lukin, “Fractional quantum Hall states of atoms in optical lattices,” Phys. Rev. Lett. 94, 086803 (2005).
  • Hafezi et al. (2007) M. Hafezi, A. S. Sørensen, E. Demler,  and M. D. Lukin, “Fractional quantum Hall effect in optical lattices,” Phys. Rev. A 76, 023613 (2007).
  • Cooper and Dalibard (2013a) Nigel R. Cooper and Jean Dalibard, “Reaching fractional quantum Hall states with optical flux lattices,” Phys. Rev. Lett. 110, 185301 (2013a).
  • Wang et al. (2011) Yi-Fei Wang, Zheng-Cheng Gu, Chang-De Gong,  and D. N. Sheng, “Fractional quantum Hall effect of hard-core bosons in topological flat bands,” Phys. Rev. Lett. 107, 146803 (2011).
  • Neupert et al. (2011) Titus Neupert, Luiz Santos, Claudio Chamon,  and Christopher Mudry, “Fractional quantum Hall states at zero magnetic field,” Phys. Rev. Lett. 106, 236804 (2011).
  • Hudomal et al. (2019) Ana Hudomal, Nicolas Regnault,  and Ivana Vasić, “Bosonic fractional quantum Hall states in driven optical lattices,” Phys. Rev. A 100, 053624 (2019).
  • Sun et al. (2011) Kai Sun, Zhengcheng Gu, Hosho Katsura,  and S. Das Sarma, “Nearly flatbands with nontrivial topology,” Phys. Rev. Lett. 106, 236803 (2011).
  • Tang et al. (2011) Evelyn Tang, Jia-Wei Mei,  and Xiao-Gang Wen, “High-temperature fractional quantum Hall states,” Phys. Rev. Lett. 106, 236802 (2011).
  • Grushin et al. (2014) Adolfo G. Grushin, Álvaro Gómez-León,  and Titus Neupert, “Floquet fractional Chern insulators,” Phys. Rev. Lett. 112, 156801 (2014).
  • Cooper and Dalibard (2013b) Nigel R. Cooper and Jean Dalibard, “Reaching fractional quantum Hall states with optical flux lattices,” Phys. Rev. Lett. 110, 185301 (2013b).
  • Miyake et al. (2013) Hirokazu Miyake, Georgios A. Siviloglou, Colin J. Kennedy, William Cody Burton,  and Wolfgang Ketterle, “Realizing the Harper Hamiltonian with laser-assisted tunneling in optical lattices,” Phys. Rev. Lett. 111, 185302 (2013).
  • Aidelsburger et al. (2013) M. Aidelsburger, M. Atala, M. Lohse, J. T. Barreiro, B. Paredes,  and I. Bloch, “Realization of the Hofstadter Hamiltonian with ultracold atoms in optical lattices,” Phys. Rev. Lett. 111, 185301 (2013).
  • Jotzu et al. (2014) Gregor Jotzu, Michael Messer, Rémi Desbuquois, Martin Lebrat, Thomas Uehlinger, Daniel Greif,  and Tilman Esslinger, “Experimental realization of the topological Haldane model with ultracold fermions,” Nature 515, 237–240 (2014).
  • Aidelsburger et al. (2015) Monika Aidelsburger, Michael Lohse, Christian Schweizer, Marcos Atala, Julio T Barreiro, Sylvain Nascimbène, NR Cooper, Immanuel Bloch,  and Nathan Goldman, “Measuring the Chern number of Hofstadter bands with ultracold bosonic atoms,” Nature Physics 11, 162–166 (2015).
  • Aidelsburger et al. (2011) M. Aidelsburger, M. Atala, S. Nascimbène, S. Trotzky, Y.-A. Chen,  and I. Bloch, “Experimental realization of strong effective magnetic fields in an optical lattice,” Phys. Rev. Lett. 107, 255301 (2011).
  • Wintersperger et al. (2020) Karen Wintersperger, Christoph Braun, F Nur Ünal, André Eckardt, Marco Di Liberto, Nathan Goldman, Immanuel Bloch,  and Monika Aidelsburger, “Realization of an anomalous Floquet topological system with ultracold atoms,” Nature Physics 16, 1058–1063 (2020).
  • Beeler et al. (2013) M. C. Beeler, R. A. Williams, K. Jimenez-Garcia, L. J. LeBlanc, A. R. Perry,  and I. B. Spielman, “The spin Hall effect in a quantum gas,” Nature 498, 201–204 (2013).
  • Stuhl et al. (2015) B.K. Stuhl, H.-I. Lu, L.M. Aycock, D. Genkina,  and I.B. Spielman, “Visualizing edge states with an atomic bose gas in the quantum Hall regime,” Science 349, 1514–1518 (2015).
  • Lin et al. (2009) Y.-J. Lin, Rob L. Compton, Karina Jiménez-García, James V. Porto,  and Ian B. Spielman, “Synthetic magnetic fields for ultracold neutral atoms,” Nature 462, 628–632 (2009).
  • Haldane (1988) F. D. M. Haldane, “Model for a quantum Hall effect without Landau levels: Condensed-matter realization of the ”parity anomaly”,” Phys. Rev. Lett. 61, 2015–2018 (1988).
  • Chin et al. (2010) Cheng Chin, Rudolf Grimm, Paul Julienne,  and Eite Tiesinga, “Feshbach resonances in ultracold gases,” Rev. Mod. Phys. 82, 1225–1286 (2010).
  • Goldman and Dalibard (2014) N. Goldman and J. Dalibard, “Periodically driven quantum systems: Effective Hamiltonians and engineered gauge fields,” Phys. Rev. X 4, 031027 (2014).
  • Eckardt and Anisimovas (2015) André Eckardt and Egidijus Anisimovas, “High-frequency approximation for periodically driven quantum systems from a Floquet-space perspective,” New Journal of Physics 17, 093039 (2015).
  • Rahav et al. (2003) Saar Rahav, Ido Gilary,  and Shmuel Fishman, “Effective Hamiltonians for periodically driven systems,” Phys. Rev. A 68, 013820 (2003).
  • Bukov et al. (2015) Marin Bukov, Luca D’Alessio,  and Anatoli Polkovnikov, “Universal high-frequency behavior of periodically driven systems: from dynamical stabilization to Floquet engineering,” Advances in Physics 64, 139–226 (2015).
  • Parameswaran et al. (2013) Siddharth A. Parameswaran, Rahul Roy,  and Shivaji L. Sondhi, “Fractional quantum Hall physics in topological flat bands,” Comptes Rendus Physique 14, 816–839 (2013).
  • Marzari et al. (2012) Nicola Marzari, Arash A. Mostofi, Jonathan R. Yates, Ivo Souza,  and David Vanderbilt, “Maximally localized Wannier functions: Theory and applications,” Rev. Mod. Phys. 84, 1419–1475 (2012).
  • Zhai (2021) Hui Zhai, Ultracold Atomic Physics (Cambridge University Press, 2021).
  • Courteille et al. (1998) Ph. Courteille, R. S. Freeland, D. J. Heinzen, F. A. van Abeelen,  and B. J. Verhaar, “Observation of a Feshbach resonance in cold atom scattering,” Phys. Rev. Lett. 81, 69–72 (1998).
  • Claussen et al. (2003) N. R. Claussen, S. J. J. M. F. Kokkelmans, S. T. Thompson, E. A. Donley, E. Hodby,  and C. E. Wieman, “Very-high-precision bound-state spectroscopy near a Rb85{}^{85}\mathrm{Rb} Feshbach resonance,” Phys. Rev. A 67, 060701(R) (2003).
  • Fukui et al. (2005) Takahiro Fukui, Yasuhiro Hatsugai,  and Hiroshi Suzuki, “Chern numbers in discretized Brillouin zone: Efficient method of computing (spin) Hall conductances,” Journal of the Physical Society of Japan 74, 1674–1677 (2005).
  • Viebahn et al. (2021) Konrad Viebahn, Joaquín Minguzzi, Kilian Sandholzer, Anne-Sophie Walter, Manish Sajnani, Frederik Görg,  and Tilman Esslinger, “Suppressing dissipation in a Floquet-Hubbard system,” Phys. Rev. X 11, 011057 (2021).
  • Niu et al. (1985) Qian Niu, D. J. Thouless,  and Yong-Shi Wu, “Quantized Hall conductance as a topological invariant,” Phys. Rev. B 31, 3372–3377 (1985).
  • Kjäll and Moore (2012) Jonas A. Kjäll and Joel E. Moore, “Edge excitations of bosonic fractional quantum Hall phases in optical lattices,” Phys. Rev. B 85, 235137 (2012).
  • Wen (1990) X. G. Wen, “Chiral Luttinger liquid and the edge excitations in the fractional quantum Hall states,” Phys. Rev. B 41, 12838–12844 (1990).
  • Estienne et al. (2015) B. Estienne, N. Regnault,  and B. A. Bernevig, “Correlation lengths and topological entanglement entropies of unitary and nonunitary fractional quantum Hall wave functions,” Phys. Rev. Lett. 114, 186801 (2015).
  • Repellin and Goldman (2019) C. Repellin and N. Goldman, “Detecting fractional Chern insulators through circular dichroism,” Phys. Rev. Lett. 122, 166801 (2019).
  • Price and Cooper (2012) H. M. Price and N. R. Cooper, “Mapping the Berry curvature from semiclassical dynamics in optical lattices,” Phys. Rev. A 85, 033620 (2012).
  • Račiūnas et al. (2018) Mantas Račiūnas, F. Nur Ünal, Egidijus Anisimovas,  and André Eckardt, “Creating, probing, and manipulating fractionally charged excitations of fractional Chern insulators in optical lattices,” Phys. Rev. A 98, 063621 (2018).
  • Marzari and Vanderbilt (1997) Nicola Marzari and David Vanderbilt, “Maximally localized generalized Wannier functions for composite energy bands,” Phys. Rev. B 56, 12847–12865 (1997).
  • Pizzi et al. (2020) Giovanni Pizzi, Valerio Vitale, Ryotaro Arita, Stefan Blügel, Frank Freimuth, Guillaume Géranton, Marco Gibertini, Dominik Gresch, Charles Johnson, Takashi Koretsune, Julen Ibañez-Azpiroz, Hyungjun Lee, Jae-Mo Lihm, Daniel Marchand, Antimo Marrazzo, Yuriy Mokrousov, Jamal I Mustafa, Yoshiro Nohara, Yusuke Nomura, Lorenzo Paulatto, Samuel Poncé, Thomas Ponweiser, Junfeng Qiao, Florian Thöle, Stepan S Tsirkin, Małgorzata Wierzbowska, Nicola Marzari, David Vanderbilt, Ivo Souza, Arash A Mostofi,  and Jonathan R Yates, “Wannier90 as a community code: new features and applications,” Journal of Physics: Condensed Matter 32, 165902 (2020).
  • Ibañez Azpiroz et al. (2013) Julen Ibañez Azpiroz, Asier Eiguren, Aitor Bergara, Giulio Pettini,  and Michele Modugno, “Tight-binding models for ultracold atoms in honeycomb optical lattices,” Phys. Rev. A 87, 011602 (2013).
  • Sträter and Eckardt (2016) Christoph Sträter and André Eckardt, “Interband Heating Processes in a Periodically Driven Optical Lattice,” Zeitschrift Naturforschung Teil A 71, 909–920 (2016)arXiv:1604.00850 [cond-mat.quant-gas] .
  • Weinberg et al. (2015) M. Weinberg, C. Ölschläger, C. Sträter, S. Prelle, A. Eckardt, K. Sengstock,  and J. Simonet, “Multiphoton interband excitations of quantum gases in driven optical lattices,” Phys. Rev. A 92, 043621 (2015).