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

Quantum phases of spin-orbital-angular-momentum coupled bosonic gases in optical lattices

Rui Cao Department of Physics, National University of Defense Technology, Changsha 410073, P. R. China    Jinsen Han Department of Physics, National University of Defense Technology, Changsha 410073, P. R. China    Jianhua Wu [email protected] Department of Physics, National University of Defense Technology, Changsha 410073, P. R. China    Jianmin Yuan Department of Physics, Graduate School of China Academy of Engineering Physics, Beijing 100193, P. R. China Department of Physics, National University of Defense Technology, Changsha 410073, P. R. China    Lianyi He [email protected] Department of Physics and State Key Laboratory of Low-Dimensional Quantum Physics, Tsinghua University, Beijing 100084, China    Yongqiang Li [email protected] Department of Physics, National University of Defense Technology, Changsha 410073, P. R. China
Abstract

Spin-orbit coupling plays an important role in understanding exotic quantum phases. In this work, we present a scheme to combine spin-orbital-angular-momentum (SOAM) coupling and strong correlations in ultracold atomic gases. Essential ingredients of this setting is the interplay of SOAM coupling and Raman-induced spin-flip hopping, engineered by lasers that couples different hyperfine spin states. In the presence of SOAM coupling only, we find rich quantum phases in the Mott-insulating regime, which support different types of spin defects such as spin vortex and composite vortex with antiferromagnetic core surrounded by the outer spin vortex. Based on an effective exchange model, we find that these competing spin textures are a result of the interplay of Dzyaloshinskii-Moriya and Heisenberg exchange interactions. In the presence of both SOAM coupling and Raman-induced spin-flip hopping, more many-body phases appear, including canted-antiferromagnetic and stripe phases. Our prediction suggests that SOAM coupling could induce rich exotic many-body phases in the strongly interacting regime.

I Introduction

Spin-orbit coupling, the interplay of particle’s spin and orbital degrees of freedom, plays a crucial role in various exotic phenomena in solid-state systems, such as the quantum spin Hall effect Bernevig et al. (2006); Konig et al. (2007); Wunderlich et al. (2005); Kato et al. (2004), topological insulators Hasan and Kane (2010), and topological superconductors Qi and Zhang (2011). Ultracold atomic system, with high controllability degrees of freedom, is also a versatile candidate to investigate these quantum phenomena, by overcoming the problem of their neutrality Dalibard et al. (2011). One of these schemes relies on two-photon Raman transitions between two hyperfine states (pseudospin) of atoms Spielman (2009), which are coupled with the atomic center-of-mass momentum Goldman et al. (2014). Here, propagation directions of laser beams are crucial to determine the type of spin-orbit coupling in ultracold atoms. When two beams counter-propagate, atom’s spin can be coupled with linear momentum of atoms, i.e. spin-linear-momentum coupling Dalibard et al. (2011); Goldman et al. (2014); Cooper et al. (2019); Galitski and Spielman (2013). Rich exotic quantum states have been observed in ultracold atomic gases with spin-linear-momentum coupling Lin et al. (2011); Wang et al. (2012); Cheuk et al. (2012); Wu et al. (2016); Huang et al. (2016); Li et al. (2017).

Another fundamental type of spin-orbit coupling is called SOAM coupling. This coupling can be achieved by a pair of copropagating Laguerre-Gaussian (LG) lasers, where LG beam modes carry different orbital angular momenta along the direction of beam propagation Allen et al. (2016); Liu et al. (2006). The atomic system obtains orbital angular momentum from the copropagating LG beams via Raman transitions among the internal hyperfine states of atoms, whereas the transfer of photon momentum into atoms is suppressed Chen et al. (2018a); Zhang et al. (2019). Within SOAM coupling, several intriguing quantum phases have been predicted theoretically Chen et al. (2016); Vasić and Balaž (2016); Sun et al. (2015); Chen et al. (2020a); Chiu et al. (2020); DeMarco and Pu (2015); Hu et al. (2015); Duan et al. (2020); Chen et al. (2020b, c); Wang et al. (2021); Bidasyuk et al. (2022) and observed experimentally Chen et al. (2018a); Zhang et al. (2019); Chen et al. (2018b); Nie et al. (2022). In these studies, however, interactions between atoms play tiny role in the various quantum phases, and one mainly focus on the weakly interacting regime.

In the paper, we combine SOAM coupling and strong correlations in ultracold gases, and focus on the response of spin degree of freedom to SOAM coupling. To achieve this goal, we propose a setup by introducing a beam with orbital angular momentum in the third direction (zz direction) for a two-component ultracold bosonic gas loaded into a blue-detuned square lattice, as shown in Fig. 1. By controlling the frequency difference between the standing wave in the xx direction and the Raman beam in the zz direction, the two hyperfine states that match the Raman selection rules can be coupled, as shown in Fig 1(b). In this setup, we actually achieve both a SOAM coupling in the zz direction Chen et al. (2018a); Zhang et al. (2019), and a Raman lattice in the xx direction Liu et al. (2013). The competition between SOAM coupling and Raman-assisted spin-flip hopping may give rise to various quantum many-body phases.

This system can be effectively modeled by an extended Bose-Hubbard model for a sufficient deep optical lattice. We specifically consider the case of half filling in the Mott regime. To obtain many-body phases of the system, a bosonic version of real-space dynamical mean-field theory (RBDMFT) is implemented. Various competing phases are obtained in the Mott-insulating regime, including canted-antiferromagnetism, spin-vortex, and composite spin-vortex with nonrotating core. To explain the many-body phases, an effective spin-exchange model is derived, and we attribute these competing spin textures to the interplay of Heisenberg exchange and Dzyaloshinskii-Moriya interactions. Upon increasing the hopping amplitudes, atoms delocalize, and superfluid phases appear, including normal superfluid, rotating superfluid with vortex texture, and stripe superfluid.

The paper is organized as follows: in section II, we introduce our setup with SOAM coupling, and the extended Bose-Hubbard model. In Section III, we give a detailed description of our RBDMFT approach. Section IV covers our results for our model. We summarize with a discussion in Section V.

II Model and Hamiltonian

Refer to caption
Figure 1: (Color online) (a) Sketch of Raman couplings, induced by a plan-wave laser with orbital angular momentum 2l2l in the zz direction, and a standing wave in the xx direction. To achieve a two-dimensional square lattice, both a standing wave in the yy-direction and a strong confinement freezing the motional degree of freedom of the atoms in the zz direction are added. (b) Atomic level diagram coupled by the pairs of the laser beams Ω1,2\Omega_{1,2}.

We consider two-component bosonic gases trapped in a conventional two-dimensional (2D) square lattice. A plane-wave laser with orbital angular momentum 2l2l is added in the zz-direction, as shown in Fig. 1(a). The two spin states are denoted as σ=\sigma=\uparrow and \downarrow, which are coupled by Raman transitions induced by the standing wave with Rabi frequency Ω1(r)\Omega_{1}(r) in the xx direction, and the plane-wave laser with Rabi frequency Ω2(r)\Omega_{2}(r) in the zz direction Deng et al. (2017); Liu et al. (2014, 2009, 2013). In the large detuning limit Ω1,2|Δ|\Omega_{1,2}\ll\left|\Delta\right|, this system can be described by an effective single-particle Hamiltonian (see Appendix A) Liu et al. (2013); DeMarco and Pu (2015)

s\displaystyle\mathcal{H}_{s} =\displaystyle= p22mlmr2Lzσz+l222mr2+Vext(r)\displaystyle\frac{p^{2}}{2m}-\frac{l\hbar}{mr^{2}}L_{z}\sigma_{z}+\frac{l^{2}\hbar^{2}}{2mr^{2}}+V_{\rm{ext}}(r) (1)
+\displaystyle+ δ2σz+Ω(r)coskxσx,\displaystyle\frac{\delta}{2}\sigma_{z}+\Omega^{\prime}(r)\cos kx\cdot\sigma_{x},

where LzL_{z} denotes orbital-angular-momentum operator of atoms along the zz direction, δ\delta the effective Zeeman field, and Ω(r)coskx\Omega^{\prime}(r)\cos kx the periodic Raman field with Ω(r)=Ω1(r)Ω2(r)Δ\Omega^{\prime}(r)=\frac{\Omega_{1}(r)\Omega_{2}(r)}{\Delta} being the effective Raman Rabi coupling. VextV_{\rm{ext}} denotes the external trap potential in the xyx-y plane, and in the following we choose an isotropic hard-wall box potential, which has already been realized experimentally Mazurenko et al. (2017); Navon et al. (2021).

For a sufficiently deep blue-detuned (Δ>0\Delta>0) optical lattice, the single-particle states at each site can be approximated by the lowest-band Wannier function ω(𝒙𝑹j)\omega\left(\boldsymbol{x}-\boldsymbol{R}_{j}\right). In this approximation, the single-particle Hamlitonian (1) can be cast into a tight-binding model

0\displaystyle\mathcal{H}_{0} =\displaystyle= i,j,σ(tσciσcjσ+itij(cicjcicj)+H.c.)\displaystyle-\sum_{\langle i,j\rangle,\sigma}\left({t_{\sigma}c_{i\sigma}^{\dagger}c_{j\sigma}+it_{ij}(c_{i\uparrow}^{\dagger}c_{j\uparrow}-c_{i\downarrow}^{\dagger}c_{j\downarrow})+{\rm H.c.}}\right) (2)
+\displaystyle+ ix(1)ixΩ(cixcix+1cixcix1+H.c.)\displaystyle\underset{i_{x}}{\sum}{(-1)^{i_{x}}\Omega\left(c_{i_{x}\uparrow}^{\dagger}c_{i_{x+1}\downarrow}-c_{i_{x}\uparrow}^{\dagger}c_{i_{x-1}\downarrow}+{\rm H.c.}\right)}
+\displaystyle+ iVtrapiniσ+mz(nini),\displaystyle\sum_{i}{V^{i}_{\rm trap}n_{i\sigma}+m_{z}\left(n_{i\uparrow}-n_{i\downarrow}\right)},

where i,j\langle i,j\rangle denotes nearest neighbors between sites ii and jj, and ixi_{x} is the site in the xx direction. ciσc_{i\sigma}^{\dagger} and ciσc_{i\sigma} are creation and annihilation operators for site ii and spin σ\sigma, respectively. tσt_{\sigma} denotes conventional hopping amplitudes between nearest neighbors, mzm_{z} the Zeeman field, niσ=ciσciσn_{i\sigma}=c^{\dagger}_{i\sigma}c_{i\sigma} the local density, and VtrapiV^{i}_{\rm trap} the external trap with the contribution from centrifugal potential l222mr2\frac{l^{2}\hbar^{2}}{2mr^{2}} being absorbed. tijt_{ij} is the nearest-neighbor hopping induced by SOAM coupling, favoring hopping along the azimuthal direction (see Appendix B) Wu et al. (2004); Bhat et al. (2006a, b, 2007)

tij\displaystyle t_{ij} =\displaystyle= 𝑑𝒙ω(𝒙𝑹i)(lmr2Lz)ω(𝒙𝑹j)\displaystyle\int{d{\boldsymbol{x}}}\omega^{\ast}\left(\boldsymbol{x}-\boldsymbol{R}_{i}\right)\left(\frac{l\hbar}{mr^{2}}L_{z}\right)\omega\left(\boldsymbol{x}-\boldsymbol{R}_{j}\right) (3)
\displaystyle\approx (xiyjxjyir2)tsoc,\displaystyle\left(\frac{x_{i}y_{j}-x_{j}y_{i}}{r^{\prime 2}}\right)t_{soc},

where tsoc=l2dm𝑑xω(xd)xω(x)t_{soc}=-\frac{l\hbar^{2}}{dm}\int dx\omega^{\ast}(x-d)\partial_{x}\omega(x) with dd being lattice constant. Here, (xi,yi)(x_{i},y_{i}) are the coordinates of the iith site with the origin at the trap center, and rr^{\prime} denotes the lattice spacing between the midpoint of sites ii, jj, and the trap center. The Raman-assisted nearest-neighbor spin-flip hopping along the xx direction

Ω=𝑑𝒙Ω(r)ω(𝒙𝑹i)|coskx|ω(𝒙𝑹j),\displaystyle\Omega=\int{d{\boldsymbol{x}}}\Omega^{\prime}(r)\omega^{\ast}(\boldsymbol{x}-\boldsymbol{R}_{i})\left|\cos kx\right|\omega(\boldsymbol{x}-\boldsymbol{R}_{j}), (4)

where the Raman-assisted onsite spin-flip hopping is zero, since atoms are symmetrically localized at the nodes for the blue-detuned lattice potential Liu et al. (2013); Deng et al. (2017).

For a deep lattice, interaction effects should be included. The ss-wave contact interaction is given by

int\displaystyle\mathcal{H}_{\rm int} =\displaystyle= i,σσ12Uσσniσ(niσδσσ),\displaystyle\underset{i,\sigma\sigma^{\prime}}{\sum}\frac{1}{2}U_{\sigma\sigma^{\prime}}n_{i\sigma}\left(n_{i\sigma^{\prime}}-\delta_{\sigma\sigma^{\prime}}\right), (5)

where U,U_{{\uparrow\uparrow},{\downarrow\downarrow}} and UU_{\uparrow\downarrow} denote the intra- and interspecies interactions, respectively. Additionally, we limit present study to the situations in which the interactions are repulsive and two hyperfine components are miscible with U=U1.01UU_{\uparrow\uparrow}=U_{\downarrow\downarrow}\equiv 1.01U_{\uparrow\downarrow} and tt=tt\equiv t_{\uparrow}=t_{\downarrow}, which is a good approximation for two-hyperfine-state mixtures of a 87Rb gas Weld et al. (2009). Thus, the total Hamiltonian of our system reads

=0+intiσμσniσ,\displaystyle\mathcal{H}=\mathcal{H}_{0}+\mathcal{H}_{\rm int}-\underset{i\sigma}{\sum}\mu_{\sigma}n_{i\sigma}, (6)

where μσ\mu_{\sigma} is the chemical potential for component σ\sigma. Due to the competition between SOAM coupling and Raman-induced hopping, it is expected that various many-body phases develop in the strongly interacting many-body system described by Eq. (6). To resolve these quantum phases, we apply real-space bosonic dynamical mean-field theory (RBDMFT), to obtain the complete phase diagrams. In the following, we set U1U_{\uparrow\downarrow}\equiv 1 and optical lattice spacing d1d\equiv 1 as the units of energy and length, respectively. We focus on the lower filling case with filling ni=ni+ni=1n_{i}=n_{i\uparrow}+n_{i\downarrow}=1 in the Mott regime (the total particle number N=ini=330N=\sum_{i}n_{i}=330), and the lattice size Nlat=24×24N_{\rm lat}=24\times 24.

III Method

To resolve the long-range order, we utilize bosonic dynamical mean-field theory (BDMFT) to calculate many-body ground states of the system described by Eq. (6). By neglecting non-local contributions to the self-energy within BDMFT Georges et al. (1996), the NN-site lattice problem can be mapped to NN single-impurity models interacting with two baths, which correspond to condensing and normal bosons, respectively Li et al. (2018, 2011, 2013); Byczuk and Vollhardt (2008). By a self-consistency condition, we can finally obtain the physical information of the NN-site model. Note here that, in a real-space system without lattice-translational symmetry, the self-energy is lattice-site dependent, i.e. Σi,j=Σiδij\Sigma_{i,j}=\Sigma_{i}\delta_{ij} with δij\delta_{ij} being a Kronecker delta, which motivates us to utilize a real-space version of BDMFT Snoek et al. (2008); Chatterjee et al. (2019); Helmes et al. (2008).

In RBDMFT, our challenge is to solve the single-impurity model, and the physics of site ii is given by the local effective action 𝒮imp(i)\mathcal{S}^{(i)}_{imp}. Following the standard derivation Georges et al. (1996), we can write down the effective action for impurity site ii, which is described by

𝒮imp(i)\displaystyle\mathcal{S}_{imp}^{(i)} =\displaystyle= 0β𝑑τ𝑑τσσ𝒄σ(i)(τ)𝓖σσ(i)(ττ)1𝒄σ(i)(τ)+0β𝑑τ12σσUσσn(i)σ(τ)(n(i)σ(τ)δσσ)\displaystyle-\int_{0}^{\beta}{d}\tau d\tau^{\prime}\underset{\sigma\sigma^{\prime}}{\sum}\boldsymbol{c}_{\sigma}^{\left(i\right)}\left(\tau\right)^{\dagger}\boldsymbol{\mathcal{G}}_{\sigma\sigma^{\prime}}^{\left(i\right)}\left(\tau-\tau^{\prime}\right)^{-1}\boldsymbol{c}_{\sigma^{\prime}}^{\left(i\right)}\left(\tau^{\prime}\right)+\int_{0}^{\beta}{d}\tau\frac{1}{2}\underset{\sigma\sigma^{\prime}}{\sum}U_{\sigma\sigma^{\prime}}{n^{\left(i\right)}}_{\sigma}\left(\tau\right)\left({n^{\left(i\right)}}_{\sigma^{\prime}}\left(\tau\right)-\delta_{\sigma\sigma^{\prime}}\right)
+\displaystyle+ 1z0βdτ(i,j,σtσ(𝒄σ(i)(τ)(ϕj,σ(i)(τ)ϕj,σ(i)(τ)))+itij(𝒄(i)(τ)(ϕj,(i)(τ)ϕj,(i)(τ))𝒄(i)(τ)(ϕj,(i)(τ)ϕj,(i)(τ)))\displaystyle\frac{1}{z}\int_{0}^{\beta}{d}\tau\left(-\underset{\langle i,j\rangle,\sigma}{\sum}t_{\sigma}\left(\boldsymbol{c}_{\sigma}^{\left(i\right)}\left(\tau\right)\left(\begin{array}[]{c}\phi_{j,\sigma}^{\left(i\right)}\left(\tau\right)^{\ast}\\ \phi_{j,\sigma}^{\left(i\right)}\left(\tau\right)\\ \end{array}\right)\right)+it_{ij}\left(\boldsymbol{c}_{\uparrow}^{\left(i\right)}\left(\tau\right)\left(\begin{array}[]{c}\phi_{j,\uparrow}^{\left(i\right)}\left(\tau\right)^{\ast}\\ -\phi_{j,\uparrow}^{\left(i\right)}\left(\tau\right)\\ \end{array}\right)-\boldsymbol{c}_{\downarrow}^{\left(i\right)}\left(\tau\right)\left(\begin{array}[]{c}\phi_{j,\downarrow}^{\left(i\right)}\left(\tau\right)^{\ast}\\ -\phi_{j,\downarrow}^{\left(i\right)}\left(\tau\right)\\ \end{array}\right)\right)\right.
+\displaystyle+ ix,σσ(1)ixΩ(cσ(ix)(τ)(ϕ(ix)ix+1,σ(τ)ϕ(ix)ix1,σ(τ))+c(ix)σ(τ)(ϕix+1,σ(ix)(τ)ϕi1,σ(ix)(τ)))\displaystyle\left.\underset{i_{x},\sigma\neq\sigma^{\prime}}{\sum}(-1)^{i_{x}}\Omega\left(c_{\sigma}^{\left(i_{x}\right)}\left(\tau\right)^{\ast}\left({\phi^{\left(i_{x}\right)}}_{i_{x}+1,\sigma^{\prime}}\left(\tau\right)-{\phi^{\left(i_{x}\right)}}_{i_{x}-1,\sigma^{\prime}}\left(\tau\right)\right)+{c^{\left(i_{x}\right)}}_{\sigma}\left(\tau\right)\left(\phi_{i_{x}+1,\sigma^{\prime}}^{\left(i_{x}\right)}\left(\tau\right)^{\ast}-\phi_{i-1,\sigma^{\prime}}^{\left(i_{x}\right)}\left(\tau\right)^{\ast}\right)\right)\right.
+\displaystyle+ i,σσVtrap(i)n(i)σ(τ)+mz(n(i)iσ(τ)n(i)iσ(τ))).\displaystyle\left.\underset{i,\sigma\neq\sigma^{\prime}}{\sum}V^{\left(i\right)}_{\rm trap}{n^{\left(i\right)}}_{\sigma}\left(\tau\right)+m_{z}\left({n^{\left(i\right)}}_{i\sigma}\left(\tau\right)-{n^{\left(i\right)}}_{i\sigma^{\prime}}\left(\tau\right)\right)\right). (14)

Here, 𝓖σσ(i)(ττ)\boldsymbol{\mathcal{G}}^{(i)}_{\sigma\sigma^{\prime}}\left(\tau-\tau^{\prime}\right) is a local non-interacting propagator interpreted as a dynamical Weiss mean field which simulates the effects of all other sites. To shorten the formula, the Nambu notation is used 𝒄σ(i)(τ)(cσ(i)(τ),cσ(i)(τ))\boldsymbol{c}^{(i)}_{\sigma}(\tau)\equiv(c^{(i)}_{\sigma}(\tau),c^{(i)}_{\sigma}(\tau)^{\ast}). The parameter zz is the lattice coordination, which is treated as a control parameter within RBDMFT. The terms up to subleading order are included in the effective action. The static bosonic mean-fields are defined in terms of the bosonic operator cjσc_{j\sigma} as

ϕjσ(i)(τ)=cjσ(i)(τ)0,\displaystyle\phi^{(i)}_{j\sigma}\left(\tau\right)=\left<c^{(i)}_{j\sigma}\left(\tau\right)\right>_{0}, (15)

where 0\left<...\right>_{0} means the expectation value in the cavity system without the impurity site.

Instead of solving the effective action directly, we normally turn to the Hamiltonian representation, i.e. Anderson impurity Hamiltonian Hubener et al. (2009); Li et al. (2011). By exactly diagonalizing the Anderson impurity Hamiltonian with a finite number of bath orbitals Georges et al. (1996); bat , we can finally obtain the local propagator

Gσσ,imp(i)(τ,τ)=T𝒄σ(i)(τ)𝒄iσ(i)(τ)𝒮imp(i).\displaystyle{G}^{(i)}_{\sigma\sigma^{\prime},imp}\left(\tau,\tau^{\prime}\right)=-\left<T\boldsymbol{c}^{(i)}_{\sigma}(\tau)\boldsymbol{c}^{(i)}_{i\sigma^{\prime}}(\tau^{\prime})^{\dagger}\right>_{\mathcal{S}^{(i)}_{imp}}. (16)

Next, we utilize the Dyson equation to obtain site-dependent self-energies in the Matsubara frequency representation

Σσσ,imp(i)(iωn)=𝒢σσ(i)(iωn)1Gσσ(i)(iωn).\displaystyle{\Sigma}^{(i)}_{\sigma\sigma^{\prime},imp}(i\omega_{n})={\mathcal{G}}^{(i)}_{\sigma\sigma^{\prime}}(i\omega_{n})^{-1}-{G}^{(i)}_{\sigma\sigma^{\prime}}(i\omega_{n}). (17)

In the framework of RBDMFT, we assume that the impurity self-energy Σimp(iωn)\Sigma_{imp}(i\omega_{n}) is local (momentum-independent) and coincides with lattice self-energy Σlattice(iωn)\Sigma_{lattice}(i\omega_{n}), whose assumption is exact in infinite dimensions and good approximations in higher dimensions Georges et al. (1996). Finally, we employ the Dyson equation in the real-space representation to obtain the interacting lattice Green’s function

𝑮σσ,lattice=1iωn+𝝁𝜺𝚺imp(iωn),\displaystyle\boldsymbol{G}_{\sigma\sigma^{\prime},lattice}=\frac{1}{i\omega_{n}+\boldsymbol{\mu}-\boldsymbol{\varepsilon}-\boldsymbol{\Sigma}_{imp}(i\omega_{n})}, (18)

where boldface quantities denote matrices with site-dependent elements. 𝜺\boldsymbol{\varepsilon} denotes a matrix with the elements being nearest-neighbor hopping amplitudes for a given lattice structure, 𝝁\boldsymbol{\mu} represents the onsite hopping amplitudes with the external trap, and 𝚺imp(iωn)\boldsymbol{\Sigma}_{imp}(i\omega_{n}) denotes the self-energy. The self-consistency RBDMFT loop is closed by the Dyson equation to obtain a new local non-interacting propagator 𝒢σσi{\mathcal{G}^{i}_{\sigma\sigma^{\prime}}}. These processes are repeated until the desired accuracy for superfluid order parameters and noninteracting Green’s functions is obtained.

IV Results

Refer to caption
Figure 2: (Color online) Nearest-neighbor hopping amplitudes tsoc/t{t_{soc}}/{t} as a function of lattice depth VV. In the regime with V>5ERV>5\,E_{R}, tsoc/t{t_{soc}}/{t} scales roughly linearly with the lattice depth VV, where ERE_{R} is the recoil energy. We choose the orbital angular momentum l=1l=1.

IV.1 Spin-orbital-angular-momentum coupling

Before exploring the whole model, described by Eq. (6), we first discuss the competition between conventional nearest-neighbor hopping tt and orbital-angular-momentum-induced hopping tsoct_{soc}. As shown in Fig. 2, the orbital-angular-momentum-induced hopping can be the order of the conventional one even for l=1l=1, where the hopping amplitudes are obtained from band-structure simulations ban . By neglecting the Raman-induced spin-flip hopping Ω\Omega, Eq. (6) is reduced to

H\displaystyle H =\displaystyle= i,j,σ(tciσcjσ+itij(ci,cj,ci,cj,)+H.c.)\displaystyle-\sum_{\langle i,j\rangle,\sigma}\left(tc_{i\sigma}^{\dagger}c_{j\sigma}+it_{ij}(c_{i,\uparrow}^{\dagger}c_{j,\uparrow}-c_{i,\downarrow}^{\dagger}c_{j,\downarrow})+{\rm H.c.}\right)
+\displaystyle+ i,σσ12Uσσniσ(niσδσσ)+Vtrapini,σμiσniσ,\displaystyle\underset{i,\sigma\sigma^{\prime}}{\sum}\frac{1}{2}U_{\sigma\sigma^{\prime}}n_{i\sigma}\left(n_{i\sigma^{\prime}}-\delta_{\sigma\sigma^{\prime}}\right)+V^{i}_{\rm trap}n_{i,\sigma}-\mu_{i\sigma}n_{i\sigma},

with mz=0m_{z}=0.

Refer to caption
Figure 3: (Color online) (a) Many-body phase diagram of two-component ultracold bosonic gases in a square lattice in the presence of SOAM coupling, described by Eq. (IV.1). The system favors three Mott phases with ferromagnetism (MII\rm MI_{I}), spin-vortex (MIII\rm MI_{II}), and composite spin-vortex with antiferromagnetic core (MIIII\rm MI_{III}). In the superfluid regime, two quantum phases appear, denoted as conventional superfluid (SFI)(\rm SF_{I}) and rotating superfluid (SFII)(\rm SF_{II}). Inset: winding number ww as a function of tsoct_{soc} with t=0.015t=0.015. (b) Contour plots of spin textures for phases MIII\rm MI_{II} and MIIII\rm MI_{III} in the Mott-insulating regime. The interactions U=U=1.01UU_{\uparrow\uparrow}=U_{\downarrow\downarrow}=1.01U_{\uparrow\downarrow}.
Refer to caption
Figure 4: (Color online) Many-body phase diagrams of two-hyperfine-state mixtures of a 87Rb gas in a square lattice, described by Eq. (6), in the presence of SOAM coupling and Raman-induced spin-flip hopping for orbital angular momenta l=1l=1 (a) and l=2l=2 (b). The system favors Mott phases with ferromagnetic (MII\rm MI_{I}), vortex (MIII\rm MI_{II}), and canted-antiferromagnetic (MIIV\rm MI_{IV}) orders, and superfluid phases with vortex (SFII\rm SF_{II}) and stripe (SFIII\rm SF_{III}) textures. The interactions U/U=U/U=1.01U_{\uparrow\uparrow}/U_{\uparrow\downarrow}=U_{\downarrow\downarrow}/U_{\uparrow\downarrow}=1.01, and the effective Zeeman field mz=0m_{z}=0.

As shown in Fig. 3, a many-body phase diagram is shown as a function of hopping amplitudes tt and tsoct_{soc} for interactions U=U=1.01UU_{\uparrow\uparrow}=U_{\downarrow\downarrow}=1.01U_{\uparrow\downarrow}, based on RBDMFT. To distinguish quantum phases, we introduce superfluid order parameters ϕσ\phi_{\sigma}, pseudospin operators Siz=12(bi,bi,bi,bi,)S^{z}_{i}=\frac{1}{2}({b}^{\dagger}_{i,\uparrow}b_{i,\uparrow}-{b}^{\dagger}_{i,\downarrow}b_{i,\downarrow}), Six=12(bi,bi,+bi,bi,)S^{x}_{i}=\frac{1}{2}({b}^{\dagger}_{i,\uparrow}b_{i,\downarrow}+{b}^{\dagger}_{i,\downarrow}b_{i,\uparrow}) and Siy=12i(bi,bi,bi,bi,)S^{y}_{i}=\frac{1}{2i}(b^{\dagger}_{i,\uparrow}b_{i,\downarrow}-{b}^{\dagger}_{i,\downarrow}b_{i,\uparrow}), and winding number w=12π𝒞arg(MiMj)w=\frac{1}{2\pi}\sum_{\mathcal{C}}{\rm arg}(M^{\ast}_{i}M_{j}) with Mi=Six+iSiyM_{i}=S^{x}_{i}+iS^{y}_{i} and 𝒞\mathcal{C} being a closed loop around the center of the trap Zhou et al. (2020); Zhang et al. (2017). We observe five different quantum phases. When tsoctt_{soc}\ll t, the system demonstrates a ferromagnetic phase (MII\rm MI_{I}), identical with the system without SOAM coupling. With the growth of tsoct_{soc}, a spin-vortex phase appears in the Mott-insulating regime with w0w\neq 0 (MIII\rm MI_{II}), since the growth of tsoct_{soc} is equivalent to the growth of orbital angular momentum ll. As shown in Fig. 3(b), SOAM-induced spin rotation appears around the trap center with winding number w=1w=1, indicating that the spin rotates slowly with the corresponding response being mainly around the center of the trap. The physical reason is that the SOAM-induced hopping is site-dependent, and pronounced around the trap center, as indicated by Eq. (3). Further increasing tsoct_{soc}, the winding number ww grows as well, as shown in the inset of Fig. 3(a), and finally we observe the whole system rotating in the regime tsoctt_{soc}\gg t (MIIII\rm MI_{III}). Interestingly, this spin-vortex phase is actually a composite vortex defect, which supports a nonrotating core of antiferromagnetic spin texture, with the nearest-neighbor spins being antiparallel in the trap center, as shown in Fig. 3(b).

To understand the underlying physics in the Mott regime, we treat the hopping as perturbations and derive an effective exchange model at half filling. With defining the projection operators 𝒫\mathcal{P} and 𝒬=1𝒫\mathcal{Q}=1-\mathcal{P}, we can project the system into the Hilbert space consisting of both singly occupied sites and the states being at least one site with double occupation, and obtain an effective exchange model Heff=𝒫Ht𝒬(1𝒬HU𝒬E)𝒬Ht𝒫H_{eff}=-\mathcal{P}H_{t}\mathcal{Q}(\frac{1}{\mathcal{Q}H_{U}\mathcal{Q}-E})\mathcal{Q}H_{t}\mathcal{P} Pinheiro et al. (2013); Pinheiro (2016); Mila and Schmidt (2011). The effective exchange model is finally given by:

Heff\displaystyle H_{eff} =\displaystyle= i,jJzSizSjz+J(SixSjx+SiySjy)\displaystyle\sum_{\langle i,j\rangle}J_{z}S_{i}^{z}S_{j}^{z}+J\left(S_{i}^{x}S_{j}^{x}+S_{i}^{y}S_{j}^{y}\right) (20)
+\displaystyle+ D(Si×Sj)z.\displaystyle D\left(S_{i}\times S_{j}\right)_{z}.

Here, Jz=4(2U1U)(t2+tij2)J_{z}=-4(\frac{2}{U}-\frac{1}{U_{\uparrow\downarrow}})(t^{2}+t^{2}_{ij}), J=4(t2tij2)UJ=-\frac{4(t^{2}-t^{2}_{ij})}{U_{\uparrow\downarrow}}, and D=8ttijUD=-\frac{8tt_{ij}}{U_{\uparrow\downarrow}}. The details of derivation are given in the Appendix C.

In the absence of SOAM coupling, this effective model is reduced to the conventional XXZ model, where the system prefers ferromagnetic and anti-ferromagnetic orders Duan et al. (2003); Kuklov and Svistunov (2003); Altman et al. (2003); Isacsson et al. (2005). In the presence of SOAM coupling, a Dzyaloshinskii-Moriya term Fert et al. (2013); Luo et al. (2014) appears in the zz direction. This term competes with the normal Heisenberg exchange interactions, resulting spin-vortex defects in the Mott-insulating regime. Interestingly, the Heisenberg exchange term JJ also depends on the SOAM-induced hopping tsoct_{soc}, and dominates in the regime tsoctt_{soc}\gg t, resulting an antiferromagnetic texture. This texture is consistent with our numerical results, as shown in Fig. 3(b). We remark that the SOAM-induced Dzyaloshinskii-Moriya term DD preserves rotational symmetry with spin texture rotating along the azimuthal direction, in contrast to the spin-linear-momentum coupling by breaking lattice-translational symmetry Cole et al. (2012); He et al. (2015); Stanescu et al. (2008); Graß et al. (2011); Mandal et al. (2012).

With the increase of hopping amplitudes, atoms delocalize and the superfluid phase appears. We characterize the superfluid phase with superfluid order parameters ϕσ\phi_{\sigma}. In the superfluid region, we observe two quantum many-body phases, with one being a phase with phase rotating (SFII\rm SF_{II}), and the other with conventional phase (SFI\rm SF_{I}).

Refer to caption
Figure 5: (Color online) Phase transitions as a function of Raman-induced spin-flip hopping for different lattice depths V=14ERV=14\,E_{R} (t0.011t\approx 0.011) (a) and V=11.5ERV=11.5\,E_{R} (t0.022t\approx 0.022) (b). Inset: spin structure factor (upper) and real-space spin texture (lower) for different phases (a), and local phases of superfluid order parameter for the spin-\uparrow component (b). The interactions U/U=U/U=1.01U_{\uparrow\uparrow}/U_{\uparrow\downarrow}=U_{\downarrow\downarrow}/U_{\uparrow\downarrow}=1.01, and the orbital angular momentum l=2l=2.

IV.2 Interplay of spin-orbital-angular-momentum coupling and Raman-induced spin-flip hopping

Now we turn to study the whole system, described by Eq. (6), and focus on the stability of spin-vortex texture in the strongly interacting regime. Generally, SOAM coupling preserves rotational symmetry and favors spin-vortex defects Chen et al. (2020b, a), whereas the one-dimensional Raman-induced spin-flip hopping prefers the stripe phase and breaks translational symmetry Wang et al. (2010); Zhai (2015). It is expected that more exotic many-body phases appear, due to the competition between SOAM coupling and Raman-induced spin-flip hopping. Here, we choose two hyperfine states of a 87Rb gas as examples, where all the Hubbard parameters are obtained from band-structure simulations ban . To emphasize the influence of SOAM coupling, we consider the orbital angular momenta l=1l=1 and l=2l=2. Note here that ttsoct\approx t_{soc} for orbital angular momentum l=1l=1 in the deep lattice, as shown in Fig. 2.

Rich phases are found in Fig. 4, including Mott-insulating phases with ferromagnetic (MII\rm MI_{I}), vortex (MIII\rm MI_{II}), and canted-antiferromagnetic (MIIV\rm MI_{IV}) orders Cooley et al. (2020), and superfluid phases with vortex (SFII\rm SF_{II}) and stripe (SFIII\rm SF_{III}) patterns. In the limit Ωtsoc\Omega\ll t_{soc}, the many-body phases develop spin-vortex textures (MIII\rm MI_{II} and SFII\rm SF_{II}), whereas the system prefers density-wave orders (MIIV\rm MI_{IV} and SFIII\rm SF_{III}) in the limit Ωtsoc\Omega\gg t_{soc}. This conclusion is consistent with our general discussion above, as a result of the interplay of SOAM coupling and Raman-induced spin-flip hopping. Note here that the region of the spin-vortex phase is enlarged for larger orbital angular momentum, as shown in Fig. 4(b), indicating large opportunity for observing this spin texture for larger orbital angular momentum.

To characterize these different phases, we choose winding number, real-space spin texture, spin-structure factor Sq=|Sieiqri|S_{\vec{q}}=\left|\vec{S}_{i}e^{i\vec{q}\cdot\vec{r}_{i}}\right|  Cole et al. (2012), and local phase of superfluid order parameter, as shown in Fig. 5. Here, we choose the orbital angular momentum l=2l=2, and different lattice depths V=14ERV=14\,E_{R} with hopping t0.011t\approx 0.011 [Fig. 5(a)], and V=11.5ERV=11.5\,E_{R} with t0.022t\approx 0.022 [Fig. 5(b)]. For small Ω\Omega, a spin-vortex phase (MIII\rm MI_{II}) develops with winding number w=4w=4, as shown in Fig. 5(a). Increasing Ω\Omega, the spin texture changes to ferromagnetic (MII\rm MI_{I}) and canted-antiferromagnetic (MIIV\rm MI_{IV}) textures with vanishing winding number, which are characterized both by magnetic spin-structure factor SqS_{\vec{q}} and real-space spin texture, as shown in the inset of Fig. 5(a). We remark here that the MIIV\rm MI_{IV} phase possesses spin-density-wave in SzS_{z}, ferromagnetic order in SxS_{x}, and antiferromagnetic order in SyS_{y}. The local phase of superfluid order parameter is shown in Fig. 5(b). For small Ω\Omega, a nonzero winding number of the local phase develops in the vortex superfluid (SFII\rm SF_{II}). With Ω\Omega larger, we find the local phase demonstrates a stripe order instead (SFIII\rm SF_{III}).

Refer to caption
Figure 6: (Color online) Many-body phase diagram of two-hyperfine-state mixtures of a 87Rb gas in a square lattice with the depth V=9ERV=9\,E_{R} (t0.045)(t\approx 0.045), as a function of Raman-induced spin-flip hopping Ω\Omega and effective magnetic field mzm_{z}. Inset: local phases of superfluid order parameters for different phases. The interactions U/U=U/U=1.01U_{\uparrow\uparrow}/U_{\uparrow\downarrow}=U_{\downarrow\downarrow}/U_{\uparrow\downarrow}=1.01, and the orbital angular momentum l=2l=2.

To understand the physical phenomena in the Mott regime, we derive an effective exchange model of the system at half filling, described by Eq. (6),

Heff\displaystyle H_{eff} =\displaystyle= ix,jxJzSixzSjxz+JxSixxSjxx+JySixySjxy\displaystyle\underset{\langle i_{x},j_{x}\rangle}{\sum}J^{\prime}_{z}S_{i_{x}}^{z}S_{j_{x}}^{z}+J^{\prime}_{x}S_{i_{x}}^{x}S_{j_{x}}^{x}+J^{\prime}_{y}S_{i_{x}}^{y}S_{j_{x}}^{y} (21)
+\displaystyle+ iy,jyJzSiyzSjyz+J(SiyxSjyx+SiyySjyy)\displaystyle\underset{\langle i_{y},j_{y}\rangle}{\sum}J_{z}S_{i_{y}}^{z}S_{j_{y}}^{z}+J\left(S_{i_{y}}^{x}S_{j_{y}}^{x}+S_{i_{y}}^{y}S_{j_{y}}^{y}\right)
+\displaystyle+ i,jD(Si×Sj)z,\displaystyle\sum_{\langle i,j\rangle}D\left(S_{i}\times S_{j}\right)_{z},

where Jz=4(2(t2+tij2)U+Ω2U(t2+tij2)U2Ω2U)J_{z}^{\prime}=-4\left(\frac{2\left(t^{2}+t_{ij}^{2}\right)}{U}+\frac{\Omega^{2}}{U_{\uparrow\downarrow}}-\frac{\left(t^{2}+t_{ij}^{2}\right)}{U_{\uparrow\downarrow}}-\frac{2\Omega^{2}}{U}\right), Jx=4((t2tij2)U+Ω2U)J_{x}^{\prime}=-4\left(\frac{\left(t^{2}-t_{ij}^{2}\right)}{U_{\uparrow\downarrow}}+\frac{\Omega^{2}}{U_{\uparrow\downarrow}}\right), and Jy=4((t2tij2)UΩ2U)J_{y}^{\prime}=-4\left(\frac{\left(t^{2}-t_{ij}^{2}\right)}{U_{\uparrow\downarrow}}-\frac{\Omega^{2}}{U_{\uparrow\downarrow}}\right). We observe that the Raman-induced spin-flip hopping does not influence Dzyaloshinskii-Moriya interactions, but induce an anisotropy for Heisenberg exchange interactions in the xx and yy directions. When Ω\Omega is large enough, the Raman-induced hopping can induce Jy,zJ^{\prime}_{y,z} to be positive and JxJ^{\prime}_{x} negative. It indicates that a spin-density-wave and canted-antiferromagnetic order develops for large Ω\Omega, consistent with our numerical results, as shown in Fig. 5(a). In the intermediate regime of Ω\Omega, the JxJ^{\prime}_{x} term dominates and a ferromagnetic order appears. For small Ω\Omega, the effective model reduces to Eq. (20), and the spin vortex pattern dominates.

In a realistic system, one can tune the balance of the two-spin components, which actually acts as an effective magnetic field mzm_{z}. Here, we can control the chemical potential difference of the two components to study the effect of the magnetic field. In Fig. 6, we fix the depth of optical lattice V=9ERV=9\,E_{R} with hopping t0.045t\approx 0.045, and study the many-body phase diagram as a function of the effective magnetic field mzm_{z} and Raman-induced spin-flip hopping Ω\Omega. When the effective magnetic field is large and negative, the spin-\downarrow component supports a vortex structure, indicated by the local phase of the superfluid order parameter, as shown in inset of Fig. 6, or vice versa. For large enough Raman coupling, the two components are mixed, and the system supports a stripe pattern in the xx direction. We remake here that the phase diagram is similar to the one achieved by the SOAM experiments in continuous space Zhang et al. (2019), where the difference is vortex-antivortex pair phase for larger Raman coupling, instead of stripe order Chen et al. (2020a, b), since we essentially include a Raman lattice in the xx direction.

V conclusion and discussion

In summary, we propose a scheme to investigate spin-orbital-angular-momentum coupling in strongly interacting bosonic gases in a two-dimensional square lattice. Using real-space dynamical mean-field theory, we obtain various quantum phases, including spin-vortex defect, composite vortex, canted-antiferromagnetic, and ferromagnetic insulating phases. Based on effective exchange models, we find that the spin-vortex texture is a result of the Dzyaloshinskii-Moriya interaction, induced by spin-orbital-angular-momentum coupling. Due to the competition of Dzyaloshinskii-Moriya and Heisenberg exchange interactions, various spin textures develop. In the superfluid, we find three quantum phases with conventional, stripe and vortex orders, characterized by the local phase of superfluid order parameters. Our study would be helpful to identify interesting many-body phases in future experiments.

VI acknowledgments

We acknowledge helpful discussions with Kaijun Jiang and Keji Chen. This work is supported by the National Natural Science Foundation of China under Grants No.12074431, 11774428, and 11974423, and National Key RD program, Grant No. 2018YFA0306503. We acknowledge the Beijing Super Cloud Computing Center (BSCC) for providing HPC resources that have contributed to the research results reported within this paper.

VII Appendix

VII.1 Single-particle Hamiltonian

In our scheme, the Raman transition is Λ\Lambda-type configuration with Ω1(r)=Ω1cos(kx)\Omega_{1}(r)=\Omega_{1}{\rm cos}(kx) along the xx direction and Ω2(r)=Ω2e2r2/ρ2ei2lϕ\Omega_{2}(r)=\Omega_{2}e^{-2r^{2}/\rho^{2}}e^{-i2l\phi} along the zz direction. In the regime |Δ|Ω1,2\left|\Delta\right|\gg\Omega_{1,2}, the single-photon transition between the ground and excited states is suppressed. We can adiabatically remove the excited state, and the system is effectively regarded as two-ground-state mixtures coupled by two-photon Raman processes. Including the two-photon Raman processes, we can obtain an effective spin-1/21/2 Hamiltonian

H=(p22m+δ2Ω(r)coskxei2lϕΩ(r)coskxei2lϕp22mδ2),\displaystyle H=\left(\begin{matrix}\frac{p^{2}}{2m}+\frac{\delta}{2}&\Omega^{\prime}\left(r\right)\cos kx\cdot e^{-i2l\phi}\\ \Omega^{\prime}\left(r\right)\cos kx\cdot e^{i2l\phi}&\frac{p^{2}}{2m}-\frac{\delta}{2}\\ \end{matrix}\right), (S1)

where δ\delta is a two-photon detuning, and Ω(r)=Ω1Ω2e2r2/ρ2/Δ\Omega^{\prime}(r)=\Omega_{1}\Omega_{2}e^{-2r^{2}/\rho^{2}}/\Delta denotes the Raman Rabi frequency. After introducing the unity transformation to the single-particle wave function

U\displaystyle U =\displaystyle= (eilϕ00eilϕ),\displaystyle\left(\begin{matrix}e^{-il\phi}&0\\ 0&e^{il\phi}\\ \end{matrix}\right), (S2)

and Pauli matrix σ\sigma, we finally obtain

H\displaystyle H =\displaystyle= 22mrr(rr)+δ2σz+(Lzlσz)2mr22+Ω(r)coskxσx,\displaystyle-\frac{\hbar^{2}}{2mr}\frac{\partial}{\partial_{r}}\left(r\frac{\partial}{\partial_{r}}\right)+\frac{\delta}{2}\sigma_{z}+\frac{\left(L_{z}-l\hbar\sigma_{z}\right)}{2mr^{2}}^{2}+\Omega^{\prime}\left(r\right)\cos kx\cdot\sigma_{x},

where Lz=izL_{z}=-i\hbar\partial_{z} is the orbital angular momentum along the zz axis. Normally, the plan-wave laser is a LG beam with the intensity being suppressed near the trap center, which can influence experimental observations. Here, we instead consider a Gaussian-type Raman beam with orbital angular momentum, where such a Gaussian beam can be obtained by a quarter-wave plate Chen et al. (2020c); Rafayelyan and Brasselet (2017). The waist of the plane-wave laser is set to ρ=20\rho=20.

VII.2 Orbital angular momentum in the Wannier basis

The angular momentum tijt_{ij} is given by

Lz\displaystyle\mathcal{H}_{L_{z}} =\displaystyle= 𝑑xΨ(x)lmr2LzΨ(x)\displaystyle\int dx\Psi^{\dagger}(x)\frac{l\hbar}{mr^{2}}L_{z}\Psi(x) (S4)
=\displaystyle= lm𝑑xΨ(x)1r2LzΨ(x).\displaystyle\frac{l\hbar}{m}\int dx\Psi^{\dagger}(x)\frac{1}{r^{2}}L_{z}\Psi(x).

For a sufficient deep lattice, the field operator Ψ(x)\Psi(x) can be expanded in the lowest-band Wannier basis ω(𝒙𝑹i)\omega\left(\boldsymbol{x}-\boldsymbol{R}_{i}\right). Eq. (S4) can be rewritten as

Lz\displaystyle\mathcal{H}_{L_{z}} =\displaystyle= lmi,jcicj𝑑x3ω(𝒙𝑹i)1r2Lzω(𝒙𝑹j).\displaystyle\frac{l\hbar}{m}\sum_{\langle i,j\rangle}c^{\dagger}_{i}c_{j}\int{dx^{3}}\omega^{\ast}\left(\boldsymbol{x}-\boldsymbol{R}_{i}\right)\frac{1}{r^{2}}L_{z}\omega\left(\boldsymbol{x}-\boldsymbol{R}_{j}\right). (S5)

For nearest neighbors ii and jj, we have the relation that

1r12Ki,j<𝑑x3ω(𝒙𝑹i)1r2Lzω(𝒙𝑹j)<1r22Ki,j,\displaystyle\frac{1}{r_{1}^{2}}K_{i,j}<\int{dx^{3}}\omega^{\ast}\left(\boldsymbol{x}-\boldsymbol{R}_{i}\right)\frac{1}{r^{2}}L_{z}\omega\left(\boldsymbol{x}-\boldsymbol{R}_{j}\right)<\frac{1}{r_{2}^{2}}K_{i,j},

where r1=max(ri,rj)r_{1}={\rm max}(r_{i},r_{j}), and r2=min(ri,rj)r_{2}={\rm min}(r_{i},r_{j}), with rir_{i} (rjr_{j}) being the distance between site ii (jj) and the trap center. For simplicity, we take the distance between the midpoint of sites ii, jj, and the trap center as the approximation of rr, and denote it by rr^{\prime}. Eq. (S4) reads

Lz\displaystyle\mathcal{H}_{L_{z}} \displaystyle\approx lmr2i,jcicjKi,j,\displaystyle\frac{l\hbar}{mr^{\prime 2}}\sum_{\langle i,j\rangle}c^{\dagger}_{i}c_{j}K_{i,j}, (S7)

with Ki,j=𝑑x3ω(𝒙𝑹i)Lzω(𝒙𝑹j)=i𝑑x3ω(𝒙𝑹i)(xyyx)ω(𝒙𝑹j)K_{i,j}=\int{dx^{3}}\omega^{\ast}\left(\boldsymbol{x}-\boldsymbol{R}_{i}\right)L_{z}\omega\left(\boldsymbol{x}-\boldsymbol{R}_{j}\right)=-i\hbar\int{dx^{3}}\omega^{\ast}\left(\boldsymbol{x}-\boldsymbol{R}_{i}\right)(x\partial_{y}-y\partial_{x})\omega\left(\boldsymbol{x}-\boldsymbol{R}_{j}\right). Generally, the Wannier function can be factorized into the xx- and yy-dependent parts for the deep lattice, and we finally obtain

Refer to caption
Figure S1: (Color online) Schematic for a 4×44\times 4 lattice. CC is the trap center, and rir_{i} (rjr_{j}) is the distance between site ii (jj) and the trap center CC, with dd being the lattice constant. rr^{\prime} is the distance between the midpoint of sites ii, jj, and the trap center CC.
Ki,j\displaystyle K_{i,j} =\displaystyle= i(dxω(xxi)xω(xxj)dyω(yyi)yω(yyj)\displaystyle-i\hbar\left(\int dx\omega^{\ast}(x-x_{i})x\omega(x-x_{j})\int dy\omega^{\ast}(y-y_{i})\partial_{y}\omega(y-y_{j})\right. (S8)
\displaystyle- dxω(xxi)xω(xxj)dyω(yyi)yω(yyj)).\displaystyle\left.\int dx\omega^{\ast}(x-x_{i})\partial_{x}\omega(x-x_{j})\int dy\omega^{\ast}(y-y_{i})y\omega(y-y_{j})\right).

For the discrete lattice system, KijK_{ij} can be written in a product form

Ki,j\displaystyle K_{i,j} =\displaystyle= i(xiyjxjyidα),\displaystyle-i\hbar\left(\frac{x_{i}y_{j}-x_{j}y_{i}}{d}\alpha\right), (S9)

with α=𝑑xω(xd)xω(x)\alpha=\int dx\omega^{\ast}\left(x-d\right)\partial_{x}\omega\left(x\right) and dd being the lattice constant, as shown in Fig. S1. We finally obtain the orbital angular momentum in the Wannier basis

Lz\displaystyle\mathcal{H}_{L_{z}} =\displaystyle= i,jil2mr2xiyjxjyid(cicjcicj)𝑑xω(xd)xω(x)\displaystyle\sum_{\langle i,j\rangle}-i\frac{l\hbar^{2}}{mr^{\prime 2}}\frac{x_{i}y_{j}-x_{j}y_{i}}{d}\left(c^{\dagger}_{i}c_{j}-c_{i}c^{\dagger}_{j}\right)\int dx\omega^{\ast}\left(x-d\right)\partial_{x}\omega\left(x\right) (S10)
=\displaystyle= i,jixiyjxjyir2(l2dm𝑑xω(xd)xω(x))(cicjcicj).\displaystyle\sum_{\langle i,j\rangle}i\frac{x_{i}y_{j}-x_{j}y_{i}}{r^{\prime 2}}\left(-\frac{l\hbar^{2}}{dm}\int dx\omega^{\ast}\left(x-d\right)\partial_{x}\omega\left(x\right)\right)\left(c^{\dagger}_{i}c_{j}-c_{i}c^{\dagger}_{j}\right).

VII.3 Effective exchange model

The system can be described by an effective exchange model in the deep Mott-insulating regime. To derive the effective model, we first divide the Hilbert space according into site occupations for filling n=1n=1. We define the operators 𝒫\mathcal{P} and 𝒬\mathcal{Q} , which denote the projection into the Mott state subspace P\mathcal{H}_{P} and the perpendicular subspace Q\mathcal{H}_{Q}. For a Hamiltonian HH, the Schrödinger equation reads

H|ψ=E|ψ.\displaystyle H\left|\psi\right>=E\left|\psi\right>. (S11)

With the unity operator 1=𝒫+𝒬1=\mathcal{P}+\mathcal{Q}, we obtain

H(𝒫+𝒬)|ψ=E(𝒫+𝒬)|ψ.\displaystyle H(\mathcal{P}+\mathcal{Q})\left|\psi\right>=E(\mathcal{P}+\mathcal{Q})\left|\psi\right>. (S12)

Multiplying by 𝒫\mathcal{P} and 𝒬\mathcal{Q} the left side of Eq. (S12) results in

(𝒫H𝒫+𝒫H𝒬)|ψ\displaystyle(\mathcal{P}H\mathcal{P}+\mathcal{P}H\mathcal{Q})|\psi\rangle =\displaystyle= E𝒫|ψ,\displaystyle E\mathcal{P}|\psi\rangle, (S13)
(𝒬H𝒫+𝒬H𝒬)|ψ\displaystyle(\mathcal{Q}H\mathcal{P}+\mathcal{Q}H\mathcal{Q})|\psi\rangle =\displaystyle= E𝒬|ψ.\displaystyle E\mathcal{Q}|\psi\rangle. (S14)

Eq. (S14) can be rewritten with the projection operator relation 𝒬2=𝒬\mathcal{Q}^{2}=\mathcal{Q},

𝒬|ψ=1E𝒬H𝒬𝒬H𝒫|ψ.\displaystyle\mathcal{Q}\left|\psi\right>=\frac{1}{E-\mathcal{Q}H\mathcal{Q}}\mathcal{Q}H\mathcal{P}|\psi\rangle. (S15)

Inserting (S15) into (S13), we obtain an equation for 𝒫|ψ\mathcal{P}|\psi\rangle,

(𝒫H𝒬1E𝒬H𝒬𝒬H𝒫)𝒫|ψ=E𝒫|ψ.\displaystyle(\mathcal{P}H\mathcal{Q}\frac{1}{E-\mathcal{Q}H\mathcal{Q}}\mathcal{Q}H\mathcal{P})\mathcal{P}|\psi\rangle=E\mathcal{P}|\psi\rangle. (S16)

For Hamiltonian HH, we can divide it into two parts H=Ht+HUH=H_{t}+H_{U}, where HtH_{t} and HUH_{U} are the hopping and interaction parts of HH, respectively. Eq. (S16) can be rewritten as

(𝒫Ht𝒬1E𝒬HU𝒬𝒬Ht𝒬𝒬Ht𝒫)𝒫|ψ=E𝒫|ψ,\displaystyle(\mathcal{P}H_{t}\mathcal{Q}\frac{1}{E-\mathcal{Q}H_{U}\mathcal{Q}-\mathcal{Q}H_{t}\mathcal{Q}}\mathcal{Q}H_{t}\mathcal{P})\mathcal{P}|\psi\rangle=E\mathcal{P}|\psi\rangle, (S17)

with 𝒫HU𝒫\mathcal{P}H_{U}\mathcal{P}, 𝒫HU𝒬\mathcal{P}H_{U}\mathcal{Q} and 𝒫Ht𝒫\mathcal{P}H_{t}\mathcal{P} being zero. Because Et2U0E\sim\frac{t^{2}}{U}\sim 0 in the Mott phase, we take 1E𝒬HU𝒬𝒬Ht𝒬1𝒬HU𝒬𝒬Ht𝒬\frac{1}{E-\mathcal{Q}H_{U}\mathcal{Q}-\mathcal{Q}H_{t}\mathcal{Q}}\approx\frac{1}{-\mathcal{Q}H_{U}\mathcal{Q}-\mathcal{Q}H_{t}\mathcal{Q}}. Using the expansion 1AB=1AΣn=0(B1A)n\frac{1}{A-B}=\frac{1}{A}\Sigma^{\infty}_{n=0}(B\frac{1}{A})^{n}, with A=𝒬HU𝒬A=-\mathcal{Q}H_{U}\mathcal{Q} and B=𝒬Ht𝒬B=\mathcal{Q}H_{t}\mathcal{Q}, we finally obtain

eff\displaystyle\mathcal{H}_{eff} =\displaystyle= 𝒫Ht𝒬1𝒬HU𝒬Σn=0(𝒬Ht𝒬1𝒬HU𝒬)n𝒬Ht𝒫.\displaystyle\mathcal{P}H_{t}\mathcal{Q}\frac{1}{-\mathcal{Q}H_{U}\mathcal{Q}}\cdot\Sigma^{\infty}_{n=0}(\mathcal{Q}H_{t}\mathcal{Q}\frac{1}{\mathcal{Q}H_{U}\mathcal{Q}})^{n}\mathcal{Q}H_{t}\mathcal{P}. (S18)

Normally, we only need to include nearest-neighbor terms in the effective Hamiltonian with n=0n=0, i.e.

eff\displaystyle\mathcal{H}_{eff} =\displaystyle= 𝒫Ht𝒬1𝒬HU𝒬𝒬Ht𝒫.\displaystyle\mathcal{P}H_{t}\mathcal{Q}\frac{1}{-\mathcal{Q}H_{U}\mathcal{Q}}\mathcal{Q}H_{t}\mathcal{P}. (S19)

In the tight-binding regime, the subspace P\mathcal{H}_{P} of the states with half filling for a two-site problem is

P:{|,,|,,|,,|,},\displaystyle\mathcal{H}_{P}:\left\{\left|\uparrow,\uparrow\right>,\left|\uparrow,\downarrow\right>,\left|\downarrow,\downarrow\right>,\left|\downarrow,\downarrow\right>\right\}, (S20)

where |σ,σ\left|\sigma,\sigma^{\prime}\right> denotes the spin state σ\sigma in the left site and σ\sigma^{\prime} in the right one. The subspace P\mathcal{H}_{P} for doubly occupied sites is

Q:{|,|,|}.\displaystyle\mathcal{H}_{Q}:\left\{\left|\uparrow\uparrow\right>,\left|\uparrow\downarrow\right>,\left|\downarrow\downarrow\right>\right\}. (S21)

In a matrix form, (𝒬HU𝒬)1\left(\mathcal{Q}H_{U}\mathcal{Q}\right)^{-1} is given by

(𝒬HU𝒬)1\displaystyle\left(\mathcal{Q}H_{U}\mathcal{Q}\right)^{-1} =\displaystyle= (1U0001U0001U).\displaystyle\left(\begin{matrix}\frac{1}{U}&0&0\\ 0&\frac{1}{U}&0\\ 0&0&\frac{1}{U_{\uparrow\downarrow}}\\ \end{matrix}\right). (S22)

According to Eq. (S19), we obtain the final effective Hamiltonian

eff\displaystyle\mathcal{H}_{eff} =\displaystyle= (4(t2+tij2)U+2Ω2U)(ni,nj,+ni,nj,)(2(t2+tij2)U+4Ω2U)(ni,nj,+ni,nj,)\displaystyle-\left(\frac{4\left(t^{2}+t_{ij}^{2}\right)}{U}+\frac{2\Omega^{2}}{U_{\uparrow\downarrow}}\right)\left(n_{i,\uparrow}n_{j,\uparrow}+n_{i,\downarrow}n_{j,\downarrow}\right)-\left(\frac{2\left(t^{2}+t_{ij}^{2}\right)}{U_{\uparrow\downarrow}}+\frac{4\Omega^{2}}{U}\right)\left(n_{i,\uparrow}n_{j,\downarrow}+n_{i,\downarrow}n_{j,\uparrow}\right)
\displaystyle- 2(t+itij)2Uci,ci,cj,cj,2(titij)2Uci,ci,cj,cj,\displaystyle\frac{2\left(-t+it_{ij}\right)^{2}}{U_{\uparrow\downarrow}}c_{i,\downarrow}^{\dagger}c_{i,\uparrow}c_{j,\uparrow}^{\dagger}c_{j,\downarrow}-\frac{2\left(-t-it_{ij}\right)^{2}}{U_{\uparrow\downarrow}}c_{i,\uparrow}^{\dagger}c_{i,\downarrow}c_{j,\downarrow}^{\dagger}c_{j,\uparrow}
\displaystyle- 2Ω2U(ci,ci,cj,cj,+ci,ci,cj,cj,)(1)ix(2Ω(titij)U+2Ω(titij)U)(cj,cj,+ci,ci,)\displaystyle\frac{2\Omega^{2}}{U_{\uparrow\downarrow}}\left(c_{i,\uparrow}^{\dagger}c_{i,\downarrow}c_{j,\uparrow}^{\dagger}c_{j,\downarrow}+c_{i,\downarrow}^{\dagger}c_{i,\uparrow}c_{j,\downarrow}^{\dagger}c_{j,\uparrow}\right)-(-1)^{i_{x}}\left(\frac{2\Omega\left(-t-it_{ij}\right)}{U}+\frac{2\Omega\left(-t-it_{ij}\right)}{U_{\uparrow\downarrow}}\right)\left(c_{j,\downarrow}^{\dagger}c_{j,\uparrow}+c_{i,\uparrow}^{\dagger}c_{i,\downarrow}\right)
\displaystyle- (1)ix(2Ω(t+itij)U+2Ω(t+itij)U)(cj,cj,+ci,ci,).\displaystyle(-1)^{i_{x}}\left(\frac{2\Omega\left(-t+it_{ij}\right)}{U_{\uparrow\downarrow}}+\frac{2\Omega\left(-t+it_{ij}\right)}{U}\right)\left(c_{j,\uparrow}^{\dagger}c_{j,\downarrow}+c_{i,\downarrow}^{\dagger}c_{i,\uparrow}\right).

Introducing the pseudospin operator as follows

Six\displaystyle S^{x}_{i} =\displaystyle= 12(ci,ci,+ci,ci,)\displaystyle\frac{1}{2}\left(c^{\dagger}_{i,\uparrow}c_{i,\downarrow}+c^{\dagger}_{i,\downarrow}c_{i,\uparrow}\right) (S24)
Siy\displaystyle S^{y}_{i} =\displaystyle= 12i(ci,ci,ci,ci,)\displaystyle\frac{1}{2i}\left(c^{\dagger}_{i,\uparrow}c_{i,\downarrow}-c^{\dagger}_{i,\downarrow}c_{i,\uparrow}\right) (S25)
Siz\displaystyle S^{z}_{i} =\displaystyle= 12(ni,ni,),\displaystyle\frac{1}{2}\left(n_{i,\uparrow}-n_{i,\downarrow}\right), (S26)

Eq. (LABEL:effec1) can be rewritten as

eff\displaystyle\mathcal{H}_{eff} =\displaystyle= 4(2(t2+tij2)U+Ω2U(t2+tij2)U2Ω2U)SizSjz4((t2tij2)U+Ω2U)SixSjx\displaystyle-4\left(\frac{2\left(t^{2}+t_{ij}^{2}\right)}{U}+\frac{\Omega^{2}}{U_{\uparrow\downarrow}}-\frac{\left(t^{2}+t_{ij}^{2}\right)}{U_{\uparrow\downarrow}}-\frac{2\Omega^{2}}{U}\right)S_{i}^{z}S_{j}^{z}-4\left(\frac{\left(t^{2}-t_{ij}^{2}\right)}{U_{\uparrow\downarrow}}+\frac{\Omega^{2}}{U_{\uparrow\downarrow}}\right)S_{i}^{x}S_{j}^{x} (S27)
\displaystyle- 4((t2tij2)UΩ2U)SiySjy8ttijU(Si×Sj)z\displaystyle 4\left(\frac{\left(t^{2}-t_{ij}^{2}\right)}{U_{\uparrow\downarrow}}-\frac{\Omega^{2}}{U_{\uparrow\downarrow}}\right)S_{i}^{y}S_{j}^{y}-\frac{8tt_{ij}}{U_{\uparrow\downarrow}}\left(S_{i}\times S_{j}\right)_{z}
+\displaystyle+ (1)ixΩt(4U+4U)(Six+Sjx)(1)ixΩtij(4U+4U)(SiySjy).\displaystyle(-1)^{i_{x}}\Omega t\left(\frac{4}{U}+\frac{4}{U_{\uparrow\downarrow}}\right)\left(S_{i}^{x}+S_{j}^{x}\right)-(-1)^{i_{x}}\Omega t_{ij}\left(\frac{4}{U}+\frac{4}{U_{\uparrow\downarrow}}\right)\left(S_{i}^{y}-S_{j}^{y}\right).

The result can be easily extended to the case with vanishing Ω=0\Omega=0, and it reads

eff\displaystyle\mathcal{H}_{eff} =\displaystyle= 4(2(t2+tij2)U(t2+tij2)U)SizSjz4(t2tij2)USixSjx\displaystyle-4\left(\frac{2\left(t^{2}+t_{ij}^{2}\right)}{U}-\frac{\left(t^{2}+t_{ij}^{2}\right)}{U_{\uparrow\downarrow}}\right)S_{i}^{z}S_{j}^{z}-\frac{4\left(t^{2}-t_{ij}^{2}\right)}{U_{\uparrow\downarrow}}S_{i}^{x}S_{j}^{x}
\displaystyle- 4(t2tij2)USiySjy8ttijU(Si×Sj)z.\displaystyle\frac{4\left(t^{2}-t_{ij}^{2}\right)}{U_{\uparrow\downarrow}}S_{i}^{y}S_{j}^{y}-\frac{8tt_{ij}}{U_{\uparrow\downarrow}}\left(S_{i}\times S_{j}\right)_{z}.

References