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

Pseudopotentials for Two-dimentional Ultracold Scattering in the Presence of Synthetic Spin-orbit-coupling

Christiaan R. Hougaard, Brendan C. Mulkerin, Xia-Ji Liu, Hui Hu, Jia Wang Centre for Quantum and Optical Science, Swinburne University of Technology, Melbourne 3122, Australia.
Abstract

We derive a pseudopotential in two dimensions (2D) with the presence of a 2D Rashba spin-orbit-coupling (SOC), following the same spirit of frame transformation in [Phys. Rev. A 95, 020702(R) (2017)]. The frame transformation correctly describes the non-trivial phase accumulation and partial wave couplings due to the presence of SOC and gives rise to a different pseudopotential than the free-space one, even when the length scale of SOC is significantly larger than the two-body potential range. As an application, we apply our pseudopotential with the Lippmann-Schwinger equation to obtain an analytical scattering matrix. To demonstrate the validity, we compare our results with a numerical scattering calculation of finite-range potential and shows perfect agreement over a wide range of scattering energy and SOC strength. Our results also indicate that the differences between our pseudopotential and the original free-space pseudopotential are essential to reproduce scattering observables correctly.

Modeling the fundamental two-body interactions is one of the critical steps in investigating the complex quantum physics of a many-body system. In particular, for systems with short-range interactions at low energies such as ultracold quantum gases, the two-body interaction can be replaced by a zero-range pseudopotential, giving the same wavefunction outside the original potential. One only needs energy-dependent scattering lengths obtained via partial expansion of two-body scattering to characterize the strength of such pseudopotential. For example, in many cases, s-wave scattering dominates at near-zero temperature, and the Fermi pseudopotential (Fermi, 1934; Huang and Yang, 1957) gives a highly accurate description of the behavior of degenerated quantum gases. For higher temperature beyond the Wigner-threshold regime, generalizing the Fermi pseudopotential with an energy-dependent s-wave scattering length can give quantitative descriptions (Bolda et al., 2002), when the contribution from the higher-partial wave is negligible. However, further generalization is necessary near resonances of higher-partial waves, vanishes of s-wave scattering or when spin-orbit effects couples higher partial waves. These regimes become particularly interesting in the field of quantum gases, where interactions can be engineered at will via Feshbach resonances (Chin et al., 2010) and synthetic spin-orbit-couplings (SOC) (Lin et al., 2011). Even back at the initial development of pseudopotential in the 1950s, Yang and Huang have already made an early attempt to tackle the generalization for higher partial waves (Huang and Yang, 1957). However, they made an algebraic mistake in their original work leading to an incorrect prefactor that is discovered and corrected much later (Stock et al., 2005; Derevianko, 2005; Idziaszek and Calarco, 2006). With the corrected prefactor, the mean-field energy shift of interacting fermions in a trap accurately matches experimental measurements (Roth and Feldmeier, 2001).

These pioneer works mentioned so far all focused on pseudopotentials in three-dimensional (3D) space. Extensions to lower dimensions have been only of academic interest until an exciting development in the area of ultracold quantum gases recently: the creation of low-dimensional quantum gases. Confining quantum gases into lower dimensions can strongly enhance the quantum correlation of the system, leading to qualitative changes (Bloch et al., 2008). Experimental realization of quasi-one-dimensional (quasi-1D) bosons allowed the verification of the fermionization of 1D Bose gases in the Tonks–Girardeau regime (Paredes et al., 2004; Kinoshita et al., 2004). More recently, tightly confining the motion of atoms in one direction creates quasi-two-dimensional (quasi-2D) quantum gases that have several applications in investigating intrigued physics, such as the observation of the BKT phase (Hadzibabic et al., 2004; Cladé et al., 2009), measurement of the equation of state (Rath et al., 2010) and super-Efimov physics (Nishida et al., 2013; Gao et al., 2015). The experimental realization of quasi-2D quantum gas motivates an elegant derivation of 2D pseudopotential for arbitrary partial waves (Kanjilal and Blume, 2006).

Another invaluable development in ultracold quantum gases in recent years is the realization of synthetic gauge fields that can be used to simulate electromagnetic interactions in systems of neutral particles. Artificial gauge fields can also couple a particle’s canonical momentum with its (pseudo)spin degrees of freedom (Lin et al., 2011; Zhang et al., 2014; Zhai, 2015), providing an essential ingredient, namely SOC, for the study of topological insulators (Dalibard et al., 2011; Goldman et al., 2014). The interplay between SOC and short-range interactions might lead to new quantum behaviors and phases, and soon attracts a lot of interest. In cold-atom systems, several experimental techniques have been developed to realize SOC such as lattice shaking (Struck et al., 2012) and Raman coupling (Lin et al., 2011). While the Raman laser scheme has already achieved one-dimensional SOC (an equal mixture of Rashba and Dresselhaus SOC) (Lin et al., 2011; Cheuk et al., 2012; Wang et al., 2012; Zhang et al., 2012a; Qu et al., 2013; Khamehchi et al., 2017), SOC with higher symmetry such as 2D and 3D isotropic ones are more closely related to the cases in condensed-matter physics. On the other hand, 2D Rashba (isotropic) SOC, which is our main focus in this work, is more experimentally accessible than 3D with less problem from heating (Huang et al., 2016; Meng et al., 2016).

In previous theoretical studies, people usually directly apply the original free-space pseudopotentials (obtained from two-body scattering without SOC) in the presence of SOC (Zhang et al., 2014; Zhai, 2015). The justification base on the argument that the characteristic wave-length of synthetic SOC, in reality, is much larger than the inverse of the range of short-range interaction. Therefore the effects of SOC were assumed to have no impact on behaviors of wavefunctions at short inter-particle distances, and hence the original pseudopotential remains valid from a perturbation point of view. Nevertheless, in a foresighted study, Cui points out that the presence of SOC at short distances intrinsically mixes different partial waves via the couplings of spin, which might lead to non-trivial influence on the short-range wavefunction (Cui, 2012). Via some numerical investigations, Cui concludes that free-space pseudopotentials, especially for higher partial waves such as pp-wave, are not satisfactory. Ever since, several studies have carefully calculated two-body scattering with the presence of 3D (Zhang et al., 2012b; Yu, 2012; Zhang et al., 2013a; Duan et al., 2013; Wang and Greene, 2015; Guan and Blume, 2016; Wang et al., 2018) or 2D (Zhang et al., 2012c, 2013b) SOC, paving the way for designing a pseudopotential model. One particularly enlightening study carried out by Guan and Blume reveals that a frame transformation approach (that we will detail later) can correctly calculate the scattering phase accumulated at short distances modified by SOC (Guan and Blume, 2017). However, a proper pseudopotential that includes the nonperturbative effects of SOC at short-range and correctly reproduces scattering observables is still missing. In this Rapid Communication, we derive an analytical form of the pseudopotential in 2D with the presence of 2D Rashba SOC, following the same spirit of frame transformation in Ref. (Guan and Blume, 2017). To verify the validity, we apply the Lippmann-Schwinger equation to obtain the analytical scattering matrix and compare it with a numerical scattering calculation with finite-range potential.

We first give a brief review of 2D pseudopotential in the free-space without the presence of SOC. We consider two identical particles (n=1,2n=1,2) of mass mm confined in a 2D xx-yy plane with position vectors 𝐫n\mathbf{r}_{n}. Seperating out the center-of-mass (COM) motion, the Hamiltonian of the relative motion is given by Hfs=𝐩2/2μ2b+U(ρ)H^{{\rm fs}}=\mathbf{p}^{2}/2\mu_{2b}+U(\rho), where μ2b=m/2\mu_{2b}=m/2 is the two-body reduced mass, 𝐫={ρ,ϕ}\mathbf{r}=\{\rho,\phi\} is the relative position in polar coordinates, and 𝐩=i{ρ,ρ1ϕ}\mathbf{p}=-i\hbar\{\partial_{\rho},\rho^{-1}\partial_{\phi}\} is the relative momentum in 2D. We also assume the potential U(ρ)U(\rho) is isotropic and short-range, i.e., vanishes beyond a small radius ρ0\rho_{0}. The isotropic symmetry allows the wavefunction to be expanded as Ψfs(𝐫)=mRmfs(ρ)Φm(ϕ)\Psi^{{\rm fs}}(\mathbf{r})=\sum_{m_{\ell}}R_{m_{\ell}}^{{\rm fs}}(\rho)\Phi_{m_{\ell}}(\phi), where Φm(ϕ)=eimϕ/2π\Phi_{m_{\ell}}(\phi)=e^{im_{\ell}\phi}/\sqrt{2\pi}, Rmfs(ρ)R_{m_{\ell}}^{{\rm fs}}(\rho) satisfies the radial Schrödinger equation

[22μ2b(2ρ2+1ρρm2ρ2)+U(ρ)E]Rmfs(ρ)=0\left[-\frac{\hbar^{2}}{2\mu_{2b}}\left(\frac{\partial^{2}}{\partial\rho^{2}}+\frac{1}{\rho}\frac{\partial}{\partial\rho}-\frac{m_{\ell}^{2}}{\rho^{2}}\right)+U(\rho)-E\right]R_{m_{\ell}}^{{\rm fs}}(\rho)=0 (1)

and adopts an asymptotic form Rmfs(ρ)Jm(kρ)tan[δm(k)]Nm(kρ)R_{m_{\ell}}^{{\rm fs}}(\rho)\propto J_{m_{\ell}}(k\rho)-\tan[\delta_{m_{\ell}}(k)]N_{m_{\ell}}(k\rho) for ρ>ρ0\rho>\rho_{0}. Here E=2k2/2μ2bE=\hbar^{2}k^{2}/2\mu_{2b} and Jm(kτr)J_{m_{\ell}}(k_{\tau}r) and Ym(kτr)Y_{m_{\ell}}(k_{\tau}r) are the Bessel functions of the first and second kind respectively. δm(k)\delta_{m_{\ell}}(k) are the energy-dependent phase shifts, satisying threshold law tan[δ0(k)]1/logk\tan[\delta_{0}(k)]\propto 1/\log k and tan[δm(k)]1/k2|m|\tan[\delta_{m_{\ell}}(k)]\propto 1/k^{2|m_{\ell}|} for |m|1|m_{\ell}|\geq 1. Reference (Kanjilal and Blume, 2006) shows that replacing U(ρ)U(\rho) by a pseudopotential Vmfs(ρ)V_{m_{\ell}}^{{\rm fs}}(\rho) can give the same asymptotic wavefunction, and hence reproduce the low-energy observables of the original finite-range potential. The explicit form of Vmfs(ρ)V_{m_{\ell}}^{{\rm fs}}(\rho) in free space is given by

Vmfs(ρ,k)=2μ2btan[δm(k)]cmk2mρm[δ(ρs)2πρO^m(ρ,k)]s0+,V_{m_{\ell}}^{{\rm fs}}(\rho,k)=-\frac{\hbar^{2}}{\mu_{2b}}\frac{\tan\left[\delta_{m_{\ell}}(k)\right]}{c_{m_{\ell}}k^{2m_{\ell}}\rho^{m_{\ell}}}\left[\frac{\delta(\rho-s)}{2\pi\rho}\hat{O}_{m_{\ell}}(\rho,k)\right]_{s\rightarrow 0^{+}}, (2)

where cm=(2m)!/[Γ(m+1)]222mc_{m_{\ell}}=(2m_{\ell})!/[\Gamma(m_{\ell}+1)]^{2}2^{2m_{\ell}} with Γ()\Gamma(\cdot) being the gamma function. The form of delta shell δ(ρs)\delta(\rho-s) of radius ss approaches to a contact potential δ(ρ)\delta(\rho) in the limit s0s\rightarrow 0, and allows us to deal with the divergence of the regularized operator rigorously. The regularized operator reads as

O^m={21tan[δ0(k)]f0(k,ρ)ρρ;m=021tan[δm(k)]fm(k,ρ)2mρ2mρm;m>0\hat{O}_{m_{\ell}}=\left\{\begin{array}[]{ll}{\frac{2}{1-\tan\left[\delta_{0}(k)\right]f_{0}(k,\rho)}\frac{\partial}{\partial\rho}\rho}&{;m_{\ell}=0}\\ {\frac{2}{1-\tan\left[\delta_{m_{\ell}}(k)\right]f_{m_{\ell}}(k,\rho)}\frac{\partial^{2m_{\ell}}}{\partial\rho^{2m_{\ell}}}\rho^{m_{\ell}}}&{;m_{\ell}>0}\end{array}\right. (3)

where

fm(k,ρ)={2π[1+γ+log(12kρ)];m=02π[n=02m112mnψ¯+log(12kρ)];m>0.f_{m_{\ell}}(k,\rho)=\left\{\begin{array}[]{ll}{\frac{2}{\pi}\left[1+\gamma+\log\left(\frac{1}{2}k\rho\right)\right]}&{;m_{\ell}=0}\\ {\frac{2}{\pi}\left[\sum_{n=0}^{2m_{\ell}-1}\frac{1}{2m_{\ell}-n}-\bar{\psi}+\log\left(\frac{1}{2}k\rho\right)\right]}&{;m_{\ell}>0}\end{array}\right.. (4)

Here 2ψ¯(m)=ψ(1)+ψ(m+1)2\bar{\psi}(m_{\ell})=\psi(1)+\psi(m_{\ell}+1), where ψ\psi denotes the digamma function. For m<0m_{\ell}<0, the pseudopotential in Eq. (2) takes the same form but with mm_{\ell} replaced by |m||m_{\ell}|. In contrast to the 3D pseudopotential, the tan[δm(k)]\tan\left[\delta_{m_{\ell}}(k)\right] dependence in the denominator of the regularized operators originates from the fact that ρρN0(kρ)\frac{\partial}{\partial\rho}\rho N_{0}(k\rho) and 2mρ2mρmNm(kρ)\frac{\partial^{2m_{\ell}}}{\partial\rho^{2m_{\ell}}}\rho^{m_{\ell}}N_{m_{\ell}}(k\rho) does not vanishes at ρ0\rho\rightarrow 0.

Now we consier the effects of SOC, where each particle feels a 2D Rashba SOC described by HSO(n)=kSO𝐩n𝐬n/mH_{{\rm SO}}^{(n)}=k_{{\rm SO}}\mathbf{p}_{n}\cdot\mathbf{s}_{n}/m, with 𝐩n\mathbf{p}_{n} and 𝐬n\mathbf{s}_{n} being the 2D momentum and spin operator of particle nn respectively. Following the spirit of Refs. (Duan et al., 2013; Wang and Greene, 2015; Guan and Blume, 2016, 2017; Wang et al., 2018), we focus on the scattering in the COM frame, where the relative Hamiltonian can be written as Hrel=Hfs+VSOH_{{\rm rel}}=H^{{\rm fs}}+V^{{\rm SO}} with VSO=kSO𝚺𝐩/2μ2bV^{{\rm SO}}=k_{{\rm SO}}\mathbf{\Sigma}\cdot\mathbf{p}/2\mu_{2b} describing the SOC effect. and 𝚺=𝐬𝟏𝐬𝟐\mathbf{\Sigma}=\mathbf{s_{1}-\mathbf{s_{2}}} is the relative spin operator. kSOk_{{\rm SO}} defines the strength of SOC coupling, and gives an energy scale ESO=2kSO2/2mE_{{\rm SO}}=\hbar^{2}k_{{\rm SO}}^{2}/2m.

A formal way to solve the corresponding relative Schrödinger equation is to formulate it as a multichannel problem by expanding the τ\tau’th independent solution as

ΨτSO(𝐫)=νRντSO(ρ)Aν(Ω),\Psi_{\tau}^{{\rm SO}}(\mathbf{r})=\sum_{\nu}R_{\nu\tau}^{{\rm SO}}(\rho)A_{\nu}(\Omega), (5)

where the channel functions Aν(Ω)Ω|νA_{\nu}(\Omega)\equiv\langle\Omega|\nu\rangle are functions of Ω\Omega that includes all degrees of freedom except for ρ\rho. Due to the azimutual symmetry, total angular momentum (along zz-axis) mjm_{j} is a good quantum number that equals to m+mSm_{\ell}+m_{S}. Here m1m_{1}, m2m_{2} and mS=m1+m2m_{S}=m_{1}+m_{2} are quantum number of the projection of the operator 𝐬1\mathbf{s}_{1}, 𝐬2\mathbf{s}_{2} and 𝐒=𝐬1+𝐬2\mathbf{S}=\mathbf{s}_{1}+\mathbf{\mathbf{s}}_{2} to the quantization zz-axis respectively. Defining the total spin basis |χ|(s1,s2),S,mS\left|\chi\right\rangle\equiv\left|(s_{1,}s_{2}),S,m_{S}\right\rangle as usual, we choose the channel functions being Aν(Ω)=imΦm(ϕ)|χA_{\nu}(\Omega)=i^{m_{\ell}}\Phi_{m_{\ell}}(\phi)\left|\chi\right\rangle with m+mS=mjm_{\ell}+m_{S}=m_{j} and S+m+s1+s2S+m_{\ell}+s_{1}+s_{2} being even/odd for bosons/fermions respectively. The subindex ν\nu collectively represents quantum numbers {m,S,mS;mj}\{m_{\ell},S,m_{S};m_{j}\}. Here we omit quantum numbers s1s_{1} and s2s_{2} in the channel index notations since they are the same for all channels. At ρ>ρ0,\rho>\rho_{{\rm 0}}, wavefuctions can be expressed as a linear combination of non-interacting (but with SOC) regular and irregular solution R¯SO=F¯G¯𝒦¯\underline{R}^{{\rm SO}}=\underline{F}-\underline{G}\underline{\mathcal{K}}, where R¯SO\underline{R}^{{\rm SO}} is the matrix form of raidal solutions RντSO(ρ)R_{\nu\tau}^{{\rm SO}}(\rho). (Through out this paper, underline implies matrix form.) The matrix elements of regular solution F¯\underline{F} can be written as Fντ(ρ)=NτCντkτJm(kτρ)F_{\nu\tau}(\rho)=N_{\tau}C_{\nu\tau}\sqrt{k_{\tau}}J_{m_{\ell}}(k_{\tau}\rho), where CντC_{\nu\tau}, kτk_{\tau} and NτN_{\tau} can be obtained by diagonalizing the non-interacting Hamiltonian using the same procedure as Ref. (Wang and Greene, 2015). The corresponding irregular solution can be obtained as Gντ(ρ)=NτCντkτYm(kτρ)G_{\nu\tau}(\rho)=N_{\tau}C_{\nu\tau}\sqrt{k_{\tau}}Y_{m_{\ell}}(k_{\tau}\rho). The scattering KK matrix 𝒦¯\underline{\mathcal{K}} determines scattering observables and is related to the more familiar SS matrix by 𝒮¯=(I¯+i𝒦¯)(I¯i𝒦¯)1\underline{\mathcal{S}}=(\underline{I}+i\underline{\mathcal{K}})(\underline{I}-i\underline{\mathcal{K}})^{-1}. Our goal is to replace the potential U(ρ)U(\rho) by a potential V¯(mj)(ρ)\underline{V}^{(m_{j})}(\rho) that acts only at ρ=0\rho=0, and gives the same asymptotic wavefunction and hence the same KK matrix. Here the underline indicates V¯(mj)(ρ)\underline{V}^{(m_{j})}(\rho) is a matrix and not nessesary diagonal due to the presene of SOC.

To derive this pseudopotential, we follow the spirit of Ref. (Guan and Blume, 2017), and apply a frame transformation approach. Defining a unitary transformation 𝒰1=exp(ikSO𝚺𝐫/2)\mathcal{U}_{1}=\exp(-ik_{{\rm SO}}\mathbf{\Sigma}\cdot\mathbf{r}/2\hbar), the “rotated” Hamiltonian Htemp𝒰11Hrel𝒰1=Hfs+ϵtemp+𝒪(ρ)H^{{\rm temp}}\equiv\mathcal{U}_{1}^{-1}H_{{\rm rel}}\mathcal{U}_{1}=H^{{\rm fs}}+\mathcal{\epsilon^{{\rm temp}}}+\mathcal{O}(\rho) is introduced as an intermediate step. Here we neglect terms of order ρ\rho and higher denoted by 𝒪(ρ)\mathcal{O}(\rho), since the pseudopotential will only act at ρ=0\rho=0. The constant term ϵtemp\mathcal{\epsilon^{{\rm temp}}} is given by ϵtemp=ESO(Σx2+Σy2+SzLz)/22\mathcal{\epsilon^{{\rm temp}}}=-E_{{\rm SO}}(\Sigma_{x}^{2}+\Sigma_{y}^{2}+S_{z}L_{z})/2\hbar^{2}, where Σx\Sigma_{x} (Σy\Sigma_{y}) are the xx(yy)-component of 𝚺\mathbf{\Sigma}, SzS_{z} is the zz-component of total spin operator 𝐒\mathbf{S}, and Lz=iϕL_{z}=-i\hbar\partial_{\phi} is the 2D angular momentum operator. For two spin-1/2 particles, this operator expanded by channel functions Aν(Ω)A_{\nu}(\Omega) gives a diagonal matrix ϵ¯temp\mathcal{\underline{\epsilon}^{{\rm temp}}} . In contrary, for higher spins, ϵ¯temp\mathcal{\underline{\epsilon}^{{\rm temp}}} is in general not diagonal, where the only non-zero matrix elements are the ones couples channels with the same mSm_{S} and hence the same mm_{\ell}. Therefore, one can introduc another ρ\rho-indepdent unitary transformation 𝒰2\mathcal{U}_{2} that is block-diagonal in mm_{\ell} subspaces and satisfies ϵ¯=𝒰21ϵ¯temp𝒰2\underline{\epsilon}=\mathcal{U}_{2}^{-1}\underline{\epsilon}^{{\rm temp}}\mathcal{U}_{2} is diagonal. Automatically, 𝒰21Hfs𝒰2=Hfs\mathcal{U}_{2}^{-1}H^{{\rm fs}}\mathcal{U}_{2}=H^{{\rm fs}} is also diagonal. Therefore, we find a unitary transformation 𝒰=𝒰2𝒰1\mathcal{U}=\mathcal{U}_{2}\mathcal{U}_{1} that leads to a set of uncoupled radial Schrödinger equation that is at least valid near the origin ρ=0\rho=0, which is given by,

[22μ2b(2ρ2+1ρρm2ρ2)+U(ρ)Eν]R~ννSO(ρ)=0,\left[-\frac{\hbar^{2}}{2\mu_{2b}}\left(\frac{\partial^{2}}{\partial\rho^{2}}+\frac{1}{\rho}\frac{\partial}{\partial\rho}-\frac{m_{\ell}^{2}}{\rho^{2}}\right)+U(\rho)-E_{\nu}\right]\tilde{R}_{\nu\nu}^{{\rm SO}}(\rho)=0, (6)

where EνE+ϵνE_{\nu}\equiv E+\epsilon_{\nu} with ϵν\epsilon_{\nu} being the diagonal matrix elements of SOC-induced energy shift ϵ¯\underline{\epsilon}. Comparing with the free-space Schrödinger equation in Eq. (1), a pseudopotential V¯~(mj)(ρ,k)\underline{\tilde{V}}^{(m_{j})}(\rho,k) with diagonal matrix elements V~νν(mj)(ρ,k)=Vmfs(ρ,kν)\tilde{V}_{\nu\nu}^{(m_{j})}(\rho,k)=V_{m_{\ell}}^{{\rm fs}}(\rho,k_{\nu}) where kν2/2μ2b=Eνk_{\nu}^{2}/2\mu_{2b}=E_{\nu} can reproduce the wavefunction R~ννSO(ρ)\tilde{R}_{\nu\nu}^{{\rm SO}}(\rho) in the rotated frame. The pseudopotential in the original frame can therefore be obtained by an inverse rotation:

V¯(mj)(ρ,k)=𝒰V¯~(mj)(ρ,k)𝒰1.\underline{V}^{(m_{j})}(\rho,k)=\mathcal{U}\tilde{\underline{V}}^{(m_{j})}(\rho,k)\mathcal{U}^{-1}. (7)

Once we obtained V¯(mj)(ρ,k)\underline{V}^{(m_{j})}(\rho,k) for all mjm_{j} (or up to a cut-off mjm_{j} in practice), the total pseudopotential can be expressed as

V(ρ,k)ΨSO(𝐫)=mjννAν(Ω)ρ𝑑ρV¯νν(mj)(ρ)×dΩAν(Ω)ΨSO(𝐫),\begin{gathered}V(\rho,k)\Psi^{{\rm SO}}(\mathbf{r})=\sum_{m_{j}}\sum_{\nu\nu^{\prime}}A_{\nu}(\Omega)\int\rho^{\prime}d\rho^{\prime}\underline{V}_{\nu\nu^{\prime}}^{(m_{j})}(\rho^{\prime})\\ \times\int d\Omega^{\prime}A_{\nu^{\prime}}^{*}(\Omega^{\prime})\Psi^{{\rm SO}}(\mathbf{r}^{\prime})\end{gathered}, (8)

which might be helpful for applications where mjm_{j} is not a good quantum number.

For illustration, we consider two spin-1/21/2 fermions in the mj=0m_{j}=0 subspace, and omit the notation of mjm_{j} hereafter unless specify otherwise. The basis are denoted by ν{m,S,mS}={1,1,1},{0,0,0},{1,1,1}\nu\equiv\{m_{\ell},S,m_{S}\}=\{-1,1,1\},\{0,0,0\},\{1,1,-1\}. In this order of the basis, the rotational matrix can be written out explicitely

𝒰¯=[cos2(λρ2)sin(λρ)2sin2(λρ2)sin(λρ)2cos(λρ)sin(λρ2sin2(λρ2)sin(λρ)2cos2(λρ2)],\underline{\mathcal{U}}=\left[\begin{array}[]{ccc}{\cos^{2}\left(\frac{\lambda\rho}{2}\right)}&{-\frac{\sin(\lambda\rho)}{\sqrt{2}}}&{-\sin^{2}\left(\frac{\lambda\rho}{2}\right)}\\ {\frac{\sin(\lambda\rho)}{\sqrt{2}}}&{\cos(\lambda\rho)}&{\frac{\sin(\lambda\rho}{\sqrt{2}}}\\ {-\sin^{2}\left(\frac{\lambda\rho}{2}\right)}&{-\frac{\sin(\lambda\rho)}{\sqrt{2}}}&{\cos^{2}\left(\frac{\lambda\rho}{2}\right)}\end{array}\right], (9)

where λ=kSO/2\lambda=k_{{\rm SO}}/2 is introduced for convenience. After rotation, the SOC-induced energy shift is given by ϵ¯=diag[0,2λ2/μ2b,0]\underline{\epsilon}={\rm diag}[0,\hbar^{2}\lambda^{2}/\mu_{2b},0], where diag[]{\rm diag[\cdot]} represents a diagonal matrix. The energy shift determines the pseudopotential in the rotated frame as V¯~(ρ,k)=diag[V1fs(ρ,kp),V0fs(ρ,ks),V1fs(ρ,kp)]\tilde{\underline{V}}(\rho,k)={\rm diag}[V_{1}^{{\rm fs}}(\rho,k_{p}),V_{0}^{{\rm fs}}(\rho,k_{s}),V_{1}^{{\rm fs}}(\rho,k_{p})], where ks=k2+2λ2k_{s}=\sqrt{k^{2}+2\lambda^{2}} and kp=kk_{p}=k. We then apply Eq. (7) to obtain the pseudopotential in the original frame. We can write the the pseudopotential as a summation of ss- and pp- wave contribution:

V¯=2μ2b{δ(ρs)2πρ[tanδs(ks)O¯s+tanδp(kp)kp2O¯p]}s0,\underline{V}=-\frac{\hbar^{2}}{\mu_{2b}}\left\{\frac{\delta(\rho-s)}{2\pi\rho}\left[\tan\delta_{s}(k_{s})\underline{O}_{s}+\frac{\tan\delta_{p}(k_{p})}{k_{p}^{2}}\underline{O}_{p}\right]\right\}_{s\rightarrow 0}, (10)

where

O¯s=[000λ2O^0ρ1ρO^0λ2O^0ρ000],\underline{O}_{s}=\left[\begin{array}[]{ccc}0&0&0\\ -\frac{\lambda}{\sqrt{2}}\hat{O}_{0}\rho&\frac{1}{\rho}\hat{O}_{0}&-\frac{\lambda}{\sqrt{2}}\hat{O}_{0}\rho\\ 0&0&0\end{array}\right], (11)

and

O¯p=[1ρO^1(1λ2ρ24)λρ2O^1ρ1ρO^1λ2ρ24λ2O^1(1λ2ρ22)λ2O^1ρλ2O^1(1λ2ρ22)1ρO^1λ2ρ24λρ2O^1ρ1ρO^1(1λ2ρ24)].\underline{O}_{p}=\left[\begin{array}[]{ccc}\frac{1}{\rho}\hat{O}_{1}\left(1-\frac{\lambda^{2}\rho^{2}}{4}\right)&\frac{\lambda}{\rho\sqrt{2}}\hat{O}_{1}\rho&-\frac{1}{\rho}\hat{O}_{1}\frac{\lambda^{2}\rho^{2}}{4}\\ \frac{\lambda}{\sqrt{2}}\hat{O}_{1}\left(1-\frac{\lambda^{2}\rho^{2}}{2}\right)&\lambda^{2}\hat{O}_{1}\rho&\frac{\lambda}{\sqrt{2}}\hat{O}_{1}\left(1-\frac{\lambda^{2}\rho^{2}}{2}\right)\\ -\frac{1}{\rho}\hat{O}_{1}\frac{\lambda^{2}\rho^{2}}{4}&\frac{\lambda}{\rho\sqrt{2}}\hat{O}_{1}\rho&\frac{1}{\rho}\hat{O}_{1}\left(1-\frac{\lambda^{2}\rho^{2}}{4}\right)\end{array}\right]. (12)
Refer to caption
Figure 1: Results near ss-wave resonances with λrvdW=0.01\lambda r_{{\rm vdW}}=0.01. The free-space scattering phase-shifts are obtained from a Lennard-Jones model potential with parameter r0=0.58rvdWr_{0}=0.58r_{{\rm vdW}} that leads to as(0)21.777rvdWa_{s}(0)\approx 21.777r_{{\rm vdW}}. (a) and (b) partial cross-sections. (c) and (d) KK matrix elements. The curves represents analytical results, and the symbols are obtained by numerical scattering calculations.
Refer to caption
Figure 2: Results near pp-wave resonances with λrvdW=0.01\lambda r_{{\rm vdW}}=0.01. The free-space scattering phase-shifts are obtained from a Lennard-Jones model potential with parameter r0=0.552981rvdWr_{0}=0.552981r_{{\rm vdW}} that leads to Ap8.577×105rvdW2A_{p}\approx-8.577\times 10^{5}r_{{\rm vdW}}^{2}. (a) and (b) partial cross-sections. (c) and (d) KK matrix elements. The curves represents analytical results, and the symbols are obtained by numerical scattering calculations.

Terms of higher order of ρ\rho can be ignored with the consideration that the pseudopotential only contribute to KK matrix with terms proportional to Fντ(s)Vνν(s,k)Fντ(s)s0F_{\nu^{\prime}\tau^{\prime}}^{*}(s)V_{\nu\nu^{\prime}}(s,k)F_{\nu\tau}(s)_{s\rightarrow 0} and Fντ(s)Vνν(s,k)Gντ(s)s0F_{\nu^{\prime}\tau^{\prime}}^{*}(s)V_{\nu\nu^{\prime}}(s,k)G_{\nu\tau}(s)_{s\rightarrow 0} [see Eqs. (15) below]. Comparing with the the free-space pseudopotential diag[V1fs(k,ρ),V0fs(k,ρ),V1fs(k,ρ)]{\rm diag}[V_{1}^{{\rm fs}}(k,\rho),V_{0}^{{\rm fs}}(k,\rho),V_{1}^{{\rm fs}}(k,\rho)], there are two important differences. One is the SOC-induced energy shift leads to a different ss-wave phase shift δs(ks)\delta_{s}(k_{s}). The other is the non-diagonal terms rised from the rotational transformation, which describes the intrinsical partial waves mixing at short distances induced by SOC. As we will see, both of these differences play significant roles in producing scattering observables correctly.

To verify the validity of the pseudopotential, we apply the Lippmann-Schwinger equation to calulate the KK matrix. The Lippmann-Schwinger equation is the integral form of the Schrödinger equation Ψτ(𝐫)=Ψ0(𝐫)+G(𝐫,𝐫)V(𝐫)Ψτ(𝐫)𝑑𝐫\Psi_{\tau}(\mathbf{r})=\Psi_{0}(\mathbf{r})+\int G(\mathbf{r},\mathbf{r^{\prime}})V(\mathbf{r^{\prime}})\Psi_{\tau}(\mathbf{r}^{\prime})d\mathbf{r}^{\prime}, or equivalently in the matrix form

R¯SO(ρ)=F¯(ρ)+𝒢¯(ρ,ρ)V¯(ρ)R¯(ρ)ρ𝑑ρ.\underline{R}^{{\rm SO}}(\rho)=\underline{F}(\rho)+\int\underline{\mathcal{G}}(\rho,\rho^{\prime})\underline{V}(\rho^{\prime})\underline{R}(\rho^{\prime})\rho^{\prime}d\rho^{\prime}. (13)

Here 𝒢¯(ρ,ρ)\underline{\mathcal{G}}(\rho,\rho^{\prime}) is the matrix representation of the Green’s function 𝒢(𝐫,𝐫)=ννAν(Ω)𝒢¯νν(ρ,ρ)Aν(Ω)\mathcal{G}(\mathbf{r},\mathbf{r^{\prime}})=\sum_{\nu\nu^{\prime}}A_{\nu}(\Omega)\underline{\mathcal{G}}_{\nu\nu^{\prime}}(\rho,\rho^{\prime})A_{\nu^{\prime}}^{*}(\Omega), which is given by

𝒢¯(ρ,ρ)=π{F¯(ρ)G(ρ),ρ<ρG¯(ρ)F¯(ρ),ρ>ρ.\underline{\mathcal{G}}\left(\rho,\rho^{\prime}\right)=\pi\left\{\begin{array}[]{ll}{\underline{F}(\rho)G^{\dagger}\left(\rho^{\prime}\right),}&{\rho<\rho^{\prime}}\\ {\underline{G}(\rho)\underline{F}^{\dagger}\left(\rho^{\prime}\right),}&{\rho>\rho^{\prime}}\end{array}.\right. (14)

The Green’s function approach becomes very helpful when the potential can be replaced by a pseudopotential V¯(ρ)δ(ρ)\underline{V}(\rho^{\prime})\propto\delta(\rho^{\prime}), and the KK matrix can then be obtained by 𝒦¯=(I+¯)1𝒜¯\underline{\mathcal{K}}=(I+\underline{\mathcal{B}})^{-1}\mathcal{\underline{A}}, where

𝒜¯=πρ𝑑ρF¯(ρ)V¯(ρ)F¯(ρ),¯=πρ𝑑ρF¯(ρ)V¯(ρ)G¯(ρ).\begin{gathered}\underline{\mathcal{A}}=-\pi\int\rho^{\prime}d\rho^{\prime}\underline{F}(\rho^{\prime})\underline{V}(\rho^{\prime})\underline{F}(\rho^{\prime}),\\ \underline{\mathcal{B}}=-\pi\int\rho^{\prime}d\rho^{\prime}\underline{F}(\rho^{\prime})\underline{V}(\rho^{\prime})\underline{G}(\rho^{\prime}).\end{gathered} (15)

Noticing it is a special property of 2D that \mathcal{B} does not vanish. As a direct consequence, the KK matrix in general cannot be written as a summation of ss- and pp-wave contribution in contrast to the 3D case as shown in Eq. (11) of Ref. (Guan and Blume, 2017).

For illustration, we focus on the case of two spin-1/21/2 fermions in the mj=0m_{j}=0 subspace. The regular solution F¯\underline{F} and irregular solution G¯\underline{G} can be determined by the coefficient CντC_{\nu\tau} in a matrix form

C¯=[1/21/21/21/21/201/21/21/2],\underline{C}=\left[\begin{array}[]{ccc}{-1/2}&{-1/2}&{1/\sqrt{2}}\\ {-1/\sqrt{2}}&{1/\sqrt{2}}&{0}\\ {1/2}&{1/2}&{1/\sqrt{2}}\end{array}\right], (16)

where the the column index τ\tau corresponding to cannonical momentum {kτ}{k1,k2,k3}={kb+λ,kbλ,k}\{k_{\tau}\}\equiv\{k_{1},k_{2},k_{3}\}=\{k_{b}+\lambda,k_{b}-\lambda,k\} and normalization {2Nτ2/μ2b}={1/kb,1/kb,1/k}\{\hbar^{2}N_{\tau}^{2}/\mu_{2b}\}=\left\{1/k_{b},1/k_{b},1/k\right\} where kb=λ2+k2k_{b}=\sqrt{\lambda^{2}+k^{2}}. One can identify τ={1,2,3}\tau=\{1,2,3\} corresponds to three different configurations |,|-,-\rangle, |+,+|+,+\rangle and |,+|-,+\rangle, where - (++) indicates the helicity, i.e. whether the spin is anti-parallel/parallel to the direction of current (Duan et al., 2013).

Inserting F¯\underline{F} and G¯\underline{G} into Eq. (15) gives 𝒜¯\underline{\mathcal{A}} and ¯\underline{\mathcal{B}} that determines 𝒦¯\underline{\mathcal{K}}. We find that the KK matrix is block-diagonal and can be expressed as

𝒦¯=[𝒦¯(+)00𝒦¯()],\mathcal{\underline{K}}=\left[\begin{array}[]{cc}\underline{\mathcal{K}}^{(+)}&0\\ 0&\underline{\mathcal{K}}^{(-)}\end{array}\right], (17)

where 𝒦¯()=tan[δp(kp)]\underline{\mathcal{K}}^{(-)}=\tan[\delta_{p}(k_{p})], and 𝒦¯(+)\underline{\mathcal{K}}^{(+)} is a 2×22\times 2 matrix. The block-diagonal structure can be understood by studying the 𝒫𝒯\mathcal{\mathcal{PT}}-symmetry of νFντ(ρ)Aν(Ω)\sum_{\nu}F_{\nu\tau}(\rho)A_{\nu}(\Omega) in the mj=0m_{j}=0 subspace, where 𝒫\mathcal{P} is defined as ϕϕ+π\phi\rightarrow\phi+\pi and 𝒯\mathcal{T} is defined as |sn,mn|sn,mn\left|s_{n},m_{n}\right\rangle\rightarrow\left|s_{n},-m_{n}\right\rangle for both n=1,2n=1,2. Defining 𝒫𝒯[Fντ(ρ)Aν(Ω)]=Πτ[Fντ(ρ)Aν(Ω)]\mathcal{P}\mathcal{T}[\sum F_{\nu\tau}(\rho)A_{\nu}(\Omega)]=\Pi_{\tau}[\sum F_{\nu\tau}(\rho)A_{\nu}(\Omega)], one finds that Πτ=+1\Pi_{\tau}=+1 /1-1 for τ={1,2}\tau=\{1,2\}/{3}\{3\}, leading to the block-diagonal structure. [For mj0m_{j}\neq 0 subspaces, sign(mj){\rm sign}(m_{j}) and 𝒫𝒯\mathcal{P}\mathcal{T} cannot simutaneously be good quantum numbers]. The expressions of matrix elements of 𝒦¯(+)\underline{\mathcal{K}}^{(+)} are analytical but quite cumbersome, and hence we only give the full expression in the supplemental materials and illustrate them in Fig. 1 and Fig. 2 as two numerical examples near ss- and pp-wave resonances respectively. In these two examples, the energy-dependent phase-shifts δs(k)\delta_{s}(k) and δp(k)\delta_{p}(k) are obtained from a free-space scattering calculation with Lenard-Jones potential U(ρ)=C6ρ6(1r06ρ6)U(\rho)=-\frac{C_{6}}{\rho^{6}}\left(1-\frac{r_{0}^{6}}{\rho^{6}}\right), where C6C_{6} defines a length scale rvdW=(2μ2bC6/2)1/4/2r_{{\rm vdW}}=(2\mu_{2b}C_{6}/\hbar^{2})^{1/4}/2 and r0r_{0} controls short-range physics and is used to tune zero-energy scattering phase shifts. The analytical results shows a good agreement with a full numerical scatteirng calculation with the representation of SOC, using a similar procedure as Ref. (Wang et al., 2018). The technical details are shown in the Supplemental Material.

Near ss-wave resonance, pp-wave scattering is negligible and the scattering matrix can be simplified as

𝒦¯(s)=T0[k1kkk2],\underline{\mathcal{K}}^{(s)}=T_{0}\left[\begin{array}[]{cc}k_{1}&-k\\ -k&k_{2}\end{array}\right], (18)

where T0=tan[δs(ks)]/2kbαsT_{0}=\tan[\delta_{s}(k_{s})]/2k_{b}\alpha_{s} and αs=1+2πtan[δs(ks)](logkks+λkbtanh1λkb)\alpha_{s}=1+\frac{2}{\pi}\tan[\delta_{s}(k_{s})]\left(\log\frac{k}{k_{s}}+\frac{\lambda}{k_{b}}\tanh^{-1}\frac{\lambda}{k_{b}}\right). The SS matrix can be obtained via 𝒮¯(s)=(I+i𝒦¯(s))(Ii𝒦¯(s))1\underline{\mathcal{S}}^{(s)}=(I+i\mathcal{\underline{\mathcal{K}}}^{(s)})(I-i\mathcal{\underline{K}}^{(s)})^{-1} and determines the scattering cross-section σττ(s)=2|𝒮¯ττ(s)δττ|2/kτ\sigma_{\tau\tau^{\prime}}^{(s)}=2|\mathcal{\underline{S}}_{\tau\tau^{\prime}}^{(s)}-\delta_{\tau\tau^{\prime}}|^{2}/k_{\tau} that are read as σ11(s)=σ21(s)=k18T02/(1+4T02kb2)\sigma_{11}^{(s)}=\sigma_{21}^{(s)}=k_{1}8T_{0}^{2}/(1+4T_{0}^{2}k_{b}^{2}) and σ12(s)=σ22(s)=k28T02/(1+4T02kb2)\sigma_{12}^{(s)}=\sigma_{22}^{(s)}=k_{2}8T_{0}^{2}/(1+4T_{0}^{2}k_{b}^{2}). As a result, σ21(s)/σ12(s)=k1/k2>1\sigma_{21}^{(s)}/\sigma_{12}^{(s)}=k_{1}/k_{2}>1 indicates that particles are preferentially scattered into the lower energy helicity “-” state. The validity of Eq. (18) near ss-wave resonances is varified in Fig. 1. In the zero-energy limit, λσ11(s)4/{1+[γ+2πlog(λas(2λ))]2}\lambda\sigma_{11}^{(s)}\rightarrow 4/\left\{1+[\gamma+\frac{2}{\pi}\log\left(\lambda a_{s}(\sqrt{2}\lambda)\right)]^{2}\right\}, where as(k)a_{s}(k) is the generalized energy-dependent ss-wave scattering length defined by cot[δs(k)]=2πlog[ksas(k)]+γ\cot[\delta_{s}(k)]=\frac{2}{\pi}\log[k_{s}a_{s}(k)]+\gamma. The rescaled cross-section therefore reach maximum when as(ks)a_{s}(k_{s}) equals to areseπγ/2/λa_{{\rm res}}\equiv e^{-\pi\gamma/2}/\lambda. In comparison, if we replace U(ρ)U(\rho) directly by diag[0,V0fs(k,ρ),0]{\rm diag}[0,V_{0}^{{\rm fs}}(k,\rho),0], the free-space pseudopotential with ss-wave only, and apply the Lippmann-Schwinger equation, the obtained KK matrix will obey the same formula Eq. (18) with ksk_{s} replaced by kk. Consequently, the rescaled cross-section reaches maximum when as(0)=aresa_{s}(0)=a_{{\rm res}}. Figure 3 shows such comparison, where one can see the SOC-induced energy shift that leads to kskk_{s}\neq k is crucial to chareterize the two-body scattering correctly, especially near the maximum of λσ11\lambda\sigma_{11}.

Refer to caption
Figure 3: Scattering results at zero scattering energy for the same Lennard-Jones potential of Fig. 1. (a) Scaled partial cross-section λσ11\lambda\sigma_{11} as a function of λ\lambda. (b) KK matrix element 𝒦11\mathcal{K}_{11} as a function of λ\lambda. The blue solid curves are the analytical results, and the red dashed curves are determined by the ss-wave only approximation Eq. (18), which are indistinguishable to the solid curves on the scale shown. The purple dash-dotted curves are calculated using the free-space pseudopotential directly. The green crosses are results from a numerical calculation using the same Lennard-Jones potential with the presence of SOC. (c) The blue solid curve shows rvdW/[as(ks)ares]r_{{\rm vdW}}/[a_{s}(k_{s})-a_{{\rm res}}] as a function of λ\lambda, while the purple dash-dotted curve shows rvdW/[as(0)ares]r_{{\rm vdW}}/[a_{s}(0)-a_{{\rm res}}].
Refer to caption
Figure 4: Scattering results at zero scattering energy for the same Lennard-Jones potential of Fig. 2. (a) Scaled partial cross-section λσ11\lambda\sigma_{11} as a function of λ\lambda. (b) KK matrix element 𝒦11\mathcal{K}_{11} as a function of λ\lambda. The blue solid curves are the analytical results, and the yellow dashed curves are determined by the ss-wave only approximation Eq. (18). The red crosses are results from a numerical calculation using the same Lennard-Jones potential with the presence of SOC.

Near pp-wave resonances, the pp-wave phase shift can no longer be neglected. Neverthuless, a simplified fromula can be obtained in the low-energy limit k0k\rightarrow 0, where tan[δs(ks)]Astan[δs(2λ)]\tan[\delta_{s}(k_{s})]\rightarrow-A_{s}\equiv-\tan[\delta_{s}(\sqrt{2}\lambda)] and tan[δp(kp)]Apk2\tan[\delta_{p}(k_{p})]\rightarrow-A_{p}k^{2}, and the KK matrix is given by,

limk0𝒦¯(+)=1d[b11b12k/λb21k/λb22k2/λ2].\lim_{k\rightarrow 0}\underline{\mathcal{K}}^{(+)}=\frac{1}{d}\left[\begin{array}[]{cc}b_{11}&b_{12}k/\lambda\\ b_{21}k/\lambda&b_{22}k^{2}/\lambda^{2}\end{array}\right]. (19)

Here, dd and bττb_{\tau^{\prime}\tau} are all constants, which is given by d=12(log2)As/π2λ2Ap/π+4λ2ApAs[(log2)1]/π2d=1-2(\log\sqrt{2})A_{s}/\pi-2\lambda^{2}A_{p}/\pi+4\lambda^{2}A_{p}A_{s}[(\log\sqrt{2})-1]/\pi^{2}, b11=As(1/2λ2Ap/π)b_{11}=-A_{s}(1/2-\lambda^{2}A_{p}/\pi), b12=b21=As(1/2+λ2Ap/π)b_{12}=b_{21}=A_{s}(1/2+\lambda^{2}A_{p}/\pi) and b22=As/4λ2Apλ2ApAs(3/2log2)/πb_{22}=-A_{s}/4-\lambda^{2}A_{p}-\lambda^{2}A_{p}A_{s}(3/2-\log 2)/\pi. When |λ2Ap|1|\lambda^{2}A_{p}|\gtrsim 1, pp-wave scattering gives a significant contribution as shown in Fig. 4, where Eq. (18) is no longer valid. Neverthuless, the threshold laws for cross-section and KK matrix elements are valid for all situations as shown in Figs. 1 and 2. Interestingly, the elastic scattering rate k1σ11\propto k_{1}\sigma_{11} that determines thermalization remains constant in the zero-energy limit, in contrast to the vanishing 1/(logk)21/(\log k)^{2} rate without the presence of SOC. We also remark here that, using the free-space pseudopotential including the pp-wave contribution diag[V1fs(k,ρ),V0fs(k,ρ),V1fs(k,ρ)]{\rm diag}[V_{1}^{{\rm fs}}(k,\rho),V_{0}^{{\rm fs}}(k,\rho),V_{1}^{{\rm fs}}(k,\rho)] will wrongly give a vanishingly small KK matrix due to a log(s)|s0\log(s)|_{s\rightarrow 0} term in the denumerator of all the matrix elements of 𝒦¯(+)\underline{\mathcal{K}}^{(+)}, reflecting the importance of the non-diagonal terms in the pseudopotnetial.

In summary, we have derived a pseudopotential in the COM frame with the presence of SOC in 2D using a frame-transformation approach. Different than the free-space pseudopotential, the s -wave scattering phase-shift changes due to a SOC-induced energy shift. The frame-transformation also introduces non-diagonal terms, which are also essential to reproduce two-body scattering observables. We applied this pseudopotential with the Lippmann-Schwinger equation to obtain the analytical scattering matrix and compare it with a numerical scattering calculation with finite-range potential. Our pseudopotential is valid even near ss- or pp-wave resonances as long as λrvdW1\lambda r_{{\rm vdW}}\ll 1, which is usually well satisfied in ultracold quantum gases. Our results indicate that, if we consider, if we consider ss-wave only (which usually implies near ss-wave resonances), and the energy-dependency of as(k)a_{s}(k) is very weak (which usually implies a very broad resonance) so that as(2λ)as(0)a_{s}(\sqrt{2}\lambda)\approx a_{s}(0), the free-space pseudopotential can give a good approximation, which gives the valid regime of previous studies in Refs. (Zhang et al., 2012c, 2013b). On the other hand, if the energy-dependency of as(k)a_{s}(k) is strong or pp-wave interaction is nonnegligible, our pseudopotential has to be adopted to reproduce two-body scattering. Our approach can also be easily applied in 3D and reproduce Eq. (11) of Ref. (Guan and Blume, 2017), which we will pursuit elsewhere. Our results are also useful for investigating universal relations and Tan’s contacts for SOC quantum gases in 2D (Cai-Xia Zhang, 2019) and might eventually be applied in many-body physics studies.

Supplemental Material

.1 Full analytical expression of KK matrix

Here, we give the full analytical expression of 𝒦¯(+)\underline{\mathcal{K}}^{(+)},

𝒦¯(+)=[𝒦11𝒦12𝒦21𝒦22]=1D[B11B12B21B22],\underline{\mathcal{K}}^{(+)}=\left[\begin{array}[]{cc}\mathcal{K}_{11}&\mathcal{K}_{12}\\ \mathcal{K}_{21}&\mathcal{K}_{22}\end{array}\right]=\frac{1}{D}\left[\begin{array}[]{cc}B_{11}&B_{12}\\ B_{21}&B_{22}\end{array}\right], (20)

where

D=1+2πts(logkks+λkbtanh1λkb)+2πtp(λ2k2λkbtanh1λkb)+4π2tstp[logk2logk3(logk)2]+4π2tstp[λ2k2(logkks+λkbtanh1λkb1)+(2logkks)(λkbtanh1λkb)],\begin{gathered}D=1+\frac{2}{\pi}t_{s}\left(\log\frac{k}{k_{s}}+\frac{\lambda}{k_{b}}\tanh^{-1}\frac{\lambda}{k_{b}}\right)+\frac{2}{\pi}t_{p}\left(\frac{\lambda^{2}}{k^{2}}-\frac{\lambda}{k_{b}}\tanh^{-1}\frac{\lambda}{k_{b}}\right)+\frac{4}{\pi^{2}}t_{s}t_{p}\left[\log k_{2}\log k_{3}-(\log k)^{2}\right]\\ +\frac{4}{\pi^{2}}t_{s}t_{p}\left[\frac{\lambda^{2}}{k^{2}}\left(\log\frac{k}{k_{s}}+\frac{\lambda}{k_{b}}\tanh^{-1}\frac{\lambda}{k_{b}}-1\right)+\left(2-\log\frac{k}{k_{s}}\right)\left(\frac{\lambda}{k_{b}}\tanh^{-1}\frac{\lambda}{k_{b}}\right)\right],\end{gathered} (21)
B11=tsk12kb+tpk22kb+tstpπ(λkb+λk2λkblogkkslogkksk22),B_{11}=t_{s}\frac{k_{1}}{2k_{b}}+t_{p}\frac{k_{2}}{2k_{b}}+\frac{t_{s}t_{p}}{\pi}\left(\frac{\lambda}{k_{b}}+\frac{\lambda}{k_{2}}-\frac{\lambda}{k_{b}}\log\frac{k}{k_{s}}-\log\frac{kk_{s}}{k_{2}^{2}}\right), (22)
B12=tsk2kb+tpk2kb+2πtstpk2kb(logkks+λ2k2),B_{12}=-t_{s}\frac{k}{2k_{b}}+t_{p}\frac{k}{2k_{b}}+\frac{2}{\pi}t_{s}t_{p}\frac{k}{2k_{b}}\left(\log\frac{k}{k_{s}}+\frac{\lambda^{2}}{k^{2}}\right), (23)
B21=tsk2kb+tpk2kb+2πtstpk2kb(logkks+λ2k2),B_{21}=-t_{s}\frac{k}{2k_{b}}+t_{p}\frac{k}{2k_{b}}+\frac{2}{\pi}t_{s}t_{p}\frac{k}{2k_{b}}\left(\log\frac{k}{k_{s}}+\frac{\lambda^{2}}{k^{2}}\right), (24)
B22=tsk22kb+tpk12kbtstpπ(λkb+λk1λkblogkks+logkksk12),B_{22}=t_{s}\frac{k_{2}}{2k_{b}}+t_{p}\frac{k_{1}}{2k_{b}}-\frac{t_{s}t_{p}}{\pi}\left(\frac{\lambda}{k_{b}}+\frac{\lambda}{k_{1}}-\frac{\lambda}{k_{b}}\log\frac{k}{k_{s}}+\log\frac{kk_{s}}{k_{1}^{2}}\right), (25)

with the notations tstan[δs(ks)]t_{s}\equiv\tan\left[\delta_{s}(k_{s})\right], tptan[δp(kp)]t_{p}\equiv\tan[\delta_{p}(k_{p})], ks=k2+2λ2k_{s}=\sqrt{k^{2}+2\lambda^{2}}, kp=kk_{p}=k, kb=k2+λ2k_{b}=\sqrt{k^{2}+\lambda^{2}}, k1=kb+λk_{1}=k_{b}+\lambda and k2=kbλk_{2}=k_{b}-\lambda.

.2 Numerical method

We carry out a numerical calculation with finite range Lennard-Jones potentials U(ρ)U(\rho) to verify our analytical results, using a similar procedure as Ref. (Wang et al., 2018). We expand the rescaled Hamiltonian hρ1/2Hrelρ1/2h\equiv\rho^{1/2}H_{{\rm rel}}\rho^{-1/2} with the channel functions that leads to a set of coupled differential equations for u¯=ρ1/2R¯SO\underline{u}=\rho^{1/2}\underline{R}^{{\rm SO}}: h¯u¯=Eu¯\underline{h}\underline{u}=E\underline{u}, where h¯=h¯fs+V¯SO\underline{h}=\underline{h}^{{\rm fs}}+\underline{V}^{{\rm SO}} has matrix elements

hννfs=[2μ2b(ρ2m214ρ2)+U(ρ)]δννh_{\nu^{\prime}\nu}^{{\rm fs}}=\left[-\frac{\hbar^{2}}{\mu_{2b}}\left(\partial_{\rho}^{2}-\frac{m_{\ell}^{2}-\frac{1}{4}}{\rho^{2}}\right)+U(\rho)\right]\delta_{\nu\text{\textquoteright}\nu} (26)

and

VννSO=2ksoc2μ2b[Σχχ(+)pmm()+Σχχ()pmm(+)],V_{\nu^{\prime}\nu}^{{\rm SO}}=\frac{\hbar^{2}k_{\mathrm{soc}}}{2\mu_{2b}}\left[\Sigma_{\chi^{\prime}\chi}^{(+)}p_{m_{\ell}^{\prime}m_{\ell}}^{(-)}+\Sigma_{\chi^{\prime}\chi}^{(-)}p_{m_{\ell}^{\prime}m_{\ell}}^{(+)}\right], (27)

where pmm(±)=(ρ1/2ρ±m/ρ)δm,m±1p_{m_{\ell}^{\prime}m_{\ell}}^{(\pm)}=\mp(\partial_{\rho}-1/2\rho\pm m_{\ell}/\rho)\delta_{m_{\ell}^{\prime},m_{\ell}\pm 1}. Here we define Σχχ(±)=χ|s1±s2±|χ/2\Sigma_{\chi^{\prime}\chi}^{(\pm)}=\left\langle\chi^{\prime}\left|s_{1}^{\pm}-s_{2}^{\pm}\right|\chi\right\rangle/2 with sn+s_{n}^{+} ( sns_{n}^{-} ) beign the raising (lowering) operator for the spin state of the nn’th particle, which can be obtained explicitly as

Σχχ(±)=m1m2Cs1m1;s2m2SmS(a1±Cs1m1±1;s2m2SmSa2±Cs1m1;s2m2±1SmS)/2\begin{array}[]{cc}\Sigma_{\chi^{\prime}\chi}^{(\pm)}=&\sum_{m_{1}m_{2}}C_{s_{1}m_{1};s_{2}m_{2}}^{Sm_{S}}(a_{1\pm}C_{s_{1}m_{1}\pm 1;s_{2}m_{2}}^{S^{\prime}m_{S}^{\prime}}\\ &-a_{2\pm}C_{s_{1}m_{1};s_{2}m_{2}\pm 1}^{S^{\prime}m_{S}^{\prime}})/2\end{array} (28)

with an±=sn(sn+1)mn(mn±1)a_{n\pm}=\sqrt{s_{n}(s_{n}+1)-m_{n}(m_{n}\pm 1)} and Cs1m1;s2m2SmSC_{s_{1}m_{1};s_{2}m_{2}}^{Sm_{S}} being Clebsch-Gordan coefficients. For numerical calculations, one can solve the multichannel radial Schrödinger equations by propagating wavefunction matrix u¯\underline{u} or equivalently the logarithmic derivative matrix ¯=u¯u¯1\underline{\mathcal{L}}=\underline{u}^{\prime}\underline{u}^{-1} to a large enough distances ρmaxρ0\rho_{{\rm max}}\gg\rho_{0}. The KK matrix can be obtained via 𝒦¯=(¯g¯g¯)1(¯f¯f¯)\underline{\mathcal{K}}=(\underline{\mathcal{L}}\underline{g}-\underline{g}^{\prime})^{-1}(\underline{\mathcal{L}}\underline{f}-\underline{f}^{\prime}), where f¯=ρF¯\underline{f}=\sqrt{\rho}\underline{F} and g¯=ρG¯\underline{g}=\sqrt{\rho}\underline{G}. The numerical results are shown as symbols in all figures in the main text, showing perfect agreement in the range of kk and λ\lambda considered.

References

  • Fermi (1934) E. Fermi, Nuovo Cimento 11, 157 (1934).
  • Huang and Yang (1957) K. Huang and C. N. Yang, Phys. Rev. 105, 767 (1957).
  • Bolda et al. (2002) E. L. Bolda, E. Tiesinga,  and P. S. Julienne, Phys. Rev. A 66, 013403 (2002).
  • Chin et al. (2010) C. Chin, R. Grimm, P. Julienne,  and E. Tiesinga, Rev. Mod. Phys. 82, 1225 (2010).
  • Lin et al. (2011) Y.-J. Lin, K. Jiménez-García,  and I. B. Spielman, Nature(London) 471, 83 (2011).
  • Stock et al. (2005) R. Stock, A. Silberfarb, E. L. Bolda,  and I. H. Deutsch, Phys. Rev. Lett. 94, 023202 (2005).
  • Derevianko (2005) A. Derevianko, Phys. Rev. A 72, 044701 (2005).
  • Idziaszek and Calarco (2006) Z. Idziaszek and T. Calarco, Phys. Rev. Lett. 96, 013201 (2006).
  • Roth and Feldmeier (2001) R. Roth and H. Feldmeier, Phys. Rev. A 64, 043603 (2001).
  • Bloch et al. (2008) I. Bloch, J. Dalibard,  and W. Zwerger, Rev. Mod. Phys. 80, 885 (2008).
  • Paredes et al. (2004) B. Paredes, A. Widera, V. Murg, O. Mandel, S. Fölling, I. Cirac, G. V. Shlyapnikov, T. W. Hänsch,  and I. Bloch, Nature(London) 429, 177 (2004).
  • Kinoshita et al. (2004) T. Kinoshita, T. Wenger,  and D. S. Weiss, Science 305, 1125 (2004).
  • Hadzibabic et al. (2004) Z. Hadzibabic, P. Krüger, M. Cheneau, B. Battelier,  and J. Dalibard, Science 305, 1125 (2004).
  • Cladé et al. (2009) P. Cladé, C. Ryu, A. Ramanathan, K. Helmerson,  and W. D. Phillips, Phys. Rev. Lett. 102, 170401 (2009).
  • Rath et al. (2010) S. P. Rath, T. Yefsah, K. J. Günter, M. Cheneau, R. Desbuquois, M. Holzmann, W. Krauth,  and J. Dalibard, Phys. Rev. A 82, 013609 (2010).
  • Nishida et al. (2013) Y. Nishida, S. Moroz,  and D. T. Son, Phys. Rev. Lett. 110, 235301 (2013).
  • Gao et al. (2015) C. Gao, J. Wang,  and Z. Yu, Phys. Rev. A 92, 020504 (2015).
  • Kanjilal and Blume (2006) K. Kanjilal and D. Blume, Phys. Rev. A 73, 060701 (2006).
  • Zhang et al. (2014) J. Zhang, H. Hu, X.-J. Liu,  and H. Pu, Annu. Rev. Cold At. Mol. 2, 81 (2014).
  • Zhai (2015) H. Zhai, Rep. Prog. Phys. 78, 026001 (2015).
  • Dalibard et al. (2011) J. Dalibard, F. Gerbier, G. Juzeliūnas,  and P. Öhberg, Rev. Mod. Phys. 83, 1523 (2011).
  • Goldman et al. (2014) N. Goldman, G. Juzeliunas, P. Öhberg,  and I. B. Spielman, Rep. Prog. Phys. 77, 126401 (2014).
  • Struck et al. (2012) J. Struck, C. Ölschläger, M. Weinberg, P. Hauke, J. Simonet, A. Eckardt, M. Lewenstein, K. Sengstock,  and P. Windpassinger, Phys. Rev. Lett. 108, 225304 (2012).
  • Cheuk et al. (2012) L. W. Cheuk, A. T. Sommer, Z. Hadzibabic, T. Yefsah, W. S. Bakr,  and M. W. Zwierlein, Phys. Rev. Lett. 109, 095302 (2012).
  • Wang et al. (2012) P. Wang, Z.-Q. Yu, Z. Fu, J. Miao, L. Huang, S. Chai, H. Zhai,  and J. Zhang, Phys. Rev. Lett. 109, 095301 (2012).
  • Zhang et al. (2012a) J.-Y. Zhang, S.-C. Ji, Z. Chen, L. Zhang, Z.-D. Du, B. Yan, G.-S. Pan, B. Zhao, Y.-J. Deng, H. Zhai, S. Chen,  and J.-W. Pan, Phys. Rev. Lett. 109, 115301 (2012a).
  • Qu et al. (2013) C. Qu, C. Hamner, M. Gong, C. Zhang,  and P. Engels, Phys. Rev. A 88, 021604 (2013).
  • Khamehchi et al. (2017) M. A. Khamehchi, K. Hossain, M. E. Mossman, Y. Zhang, T. Busch, M. M. Forbes,  and P. Engels, Phys. Rev. Lett. 118, 155301 (2017).
  • Huang et al. (2016) L. Huang, Z. Meng, P. Wang, P. Peng, S.-L. Zhang, L. Chen, D. L. andQi Zhou,  and J. Zhang, Nat. Phys. 12, 540 (2016).
  • Meng et al. (2016) Z. Meng, L. Huang, P. Peng, D. Li, L. Chen, Y. Xu, C. Zhang, P. Wang,  and J. Zhang, Phys. Rev. Lett. 117, 235304 (2016).
  • Cui (2012) X. Cui, Phys. Rev. A 85, 022705 (2012).
  • Zhang et al. (2012b) P. Zhang, L. Zhang,  and Y. Deng, Phys. Rev. A 86, 053608 (2012b).
  • Yu (2012) Z. Yu, Phys. Rev. A 85, 042711 (2012).
  • Zhang et al. (2013a) L. Zhang, Y. Deng,  and P. Zhang, Phys. Rev. A 87, 053626 (2013a).
  • Duan et al. (2013) H. Duan, L. You,  and B. Gao, Phys. Rev. A 87, 052708 (2013).
  • Wang and Greene (2015) S.-J. Wang and C. H. Greene, Phys. Rev. A 91, 022706 (2015).
  • Guan and Blume (2016) Q. Guan and D. Blume, Phys. Rev. A 94, 022706 (2016).
  • Wang et al. (2018) J. Wang, C. R. Hougaard, B. C. Mulkerin,  and X.-J. Liu, Phys. Rev. A 97, 042709 (2018).
  • Zhang et al. (2012c) P. Zhang, L. Zhang,  and W. Zhang, Phys. Rev. A 86, 042707 (2012c).
  • Zhang et al. (2013b) L. Zhang, Y. Deng,  and P. Zhang, Phys. Rev. A 87, 053626 (2013b).
  • Guan and Blume (2017) Q. Guan and D. Blume, Phys. Rev. A 95, 020702 (2017).
  • Cai-Xia Zhang (2019) K. J. Cai-Xia Zhang, Shi-Guo Peng, “Universal relations for spin-orbit-coupled fermi gases in two and three dimensions,”  (2019), arXiv:1907.11433 .