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

thanks: [email protected]

Majorana corner pairs in a two-dimensional ss-wave cold atomic superfluid

Ya-Jie Wu Department of Physics, The University of Texas at Dallas, Richardson, Texas 75080-3021, USA School of Science, Xi’an Technological University, Xi’an 710032, China    Xi-Wang Luo Department of Physics, The University of Texas at Dallas, Richardson, Texas 75080-3021, USA    Junpeng Hou Department of Physics, The University of Texas at Dallas, Richardson, Texas 75080-3021, USA    Chuanwei Zhang Department of Physics, The University of Texas at Dallas, Richardson, Texas 75080-3021, USA
Abstract

We propose a method to prepare Majorana pairs at the corners of imprinted defects on a two-dimensional cold atom optical lattice with ss-wave superfluid pairing. Different from previous proposals that manipulate the effective Dirac masses, our scheme relies on the sign flip of the spin-orbit coupling at the corners, which can be tuned in experiments by adjusting the angle of incident Raman lasers. The Majorana corner pairs are found to be located at the interface between two regimes with opposite spin orbit coupling strengths in an anticlockwise direction and are robust against certain symmetry-persevered perturbations. Our work provides a new way for implementing and manipulating Majorana pairs with existing cold-atom techniques.

I Introduction

Majorana zero modes (MZMs) have attracted great attention in past decades owing to their non-Abelian exchange statistics and potential applications as topologically protected qubits Kitaev2003 ; Nayak2008 . They also exhibit significant physics in a range of disciplines such as nuclear and particle physics Elliott2015 . Strenuous efforts to search for MZMs are underway in both theories and experiments. In recent years, a variety of schemes to realize Majorana excitations have been proposed Qix2011 ; Hansan2010 ; zhang2008 ; Sato2009 ; Jiang2011 ; Alicea2011 ; Alicea2012 ; Read2000 ; Phong2017 ; Oppen2010 ; Lutchyn2010 ; Jiang2016 ; Sau2010 ; Tewari2012 ; Teemu013 by utilizing pp-wave superconductors (SCs) or superfluid (SFs) Read2000 ; Phong2017 , or SCs and SFs with effective pp-wave paring via spin-orbit coupling (SOC) and ss-wave pairing Hansan2010 ; Qix2011 ; zhang2008 ; Sato2009 ; Jiang2011 . Remarkable experimental progresses have been made in condensed-matter systems Mourik2012 ; Finck2013 ; Nadj2014 ; Xuj2015 ; Wang2016 ; Heq2017 . The experimental realization of SOC in ultracold atomic gases offer another clean platform to explore Majorana physics Lin2011 ; Zhangj2012 ; Wang2012 ; Williams2013 ; Huang2016 ; Meng2016 ; Wuz2016 . In these platforms, the interplay among SOC, Zeeman fields and ss-wave interactions could produce non-Abelian topological superfluids (TSFs) that host Majorana excitations. There have been several tantalizing proposals for realizing and tuning Majorana excitations, for example, by creating topological defects (such as SF vortices or lattice dislocations) or defect chain Lang2014 ; Deng2015 .

The emergence of Majorana excitations can be intuitively understood by the low-energy theory. A pair of MZMs exist at the kinks where the pairing potential or SOC changes the sign (which corresponds to the sign change of Dirac mass or velocity in Jackiw–Rebbi model). In solid-state materials, the SOC kinks are difficult to tune, while the kinks of pairing potentials can be realized through Josephson junctions in superconducting nanowires, as well as the corners and hinges in recently proposed higher-order topological SCs (TSCs) J. Langbehn2017 ; Benalcazar2017 ; Song2017 ; Fang2017 ; Schindler2018 ; Khalaf2017 ; ZhuX2018 ; Khalaf2018 ; Yan2018 ; Wang2018 ; wangyu2018 ; Liu2018 ; Hsu2018 . In particular, for two-dimensional (2D) second-order TSCs, the bulk topology of the 2D system offers 1D edge modes, which have different topologies for adjacent edges due to the change of the pairing sign, leading to zero-energy Majorana Kramers pairs or Majorana modes at corners Yan2018 ; Wang2018 ; wangyu2018 ; Liu2018 . On the other hand, in cold atomic system, the kinks of pairing potentials may be obtained by soliton excitations Xu2014 ; LiuX2015 . However, these cold atomic systems suffer from dynamical instability in the presence of perturbations. Thus, proposals for realizing robust MZMs in atomic systems, through other manners like SOC kinks, are highly in demand.

In this paper, we propose feasible schemes to realize SOC kinks using trapped ultracold fermionic atoms on a 2D optical lattice, and show that our system supports Majorana pairs in a vortex-free configuration. The main results are listed below:

(i) Effective 1D modes on a rectangular geometry would emerge in the 2D system through engineering on-site potential of the rectangle. In the presence of 1D equal Rashba-Dreeshauls (ERD) SOC, Majorana corner pair emerges at the corner of the rectangle with a proper ss-wave pairing.

(ii) Each edge of the rectangular defect is characterized by a 1D topological SF in the BDI class, and Majorana pair exists at the interface of two adjacent edges which have different signs of SOC (clockwise or anticlockwise along the defective geometry). Our system is in analog with the higher-order TSCs except that the low-energy 1D model is induced by the defect rectangle rather than the bulk topology, and the MZMs are induced by SOC kinks rather than pairing kinks between two adjacent boundaries Yan2018 ; Wang2018 . Our system is more concise and experimentally friendly since no tricky unconventional pairings like dd-wave or s±s_{\pm}-wave are required.

(iii) Our system can be realized with currently already established experimental techniques in cold atoms, including 1D ERD SOC by Raman lasers Lin2011 ; Zhangj2012 , single-site addressing in 2D optical lattices Peter2009 ; Bakr2009 ; Bakr2010 ; Sherson2010 ; Weitenberg2011 , and tunable ss-wave interaction through Feshbach resonance Thorsten2006 ; Chin2010 .

(iv) The Majorana corner pair is robust against shape deformation of the defective rectangle, even to the extent of a defective loop. In addition, the SOC direction dictates on which corners the Majorana pair resides. Therefore the incident direction of Raman lasers can be used to manipulate the Majorana pairs.

The paper is organized as follows. In Sec. II we first introduce the Hamiltonian with 1D ERD SOC on optical lattices, and obtain the phase diagram consisting of metal and ss-wave SF phases. In Sec. III, we study the case with a defective rectangle, and find Majorana pairs emerge at the corners. We extend the discussion to a ring-shaped geometry in Sec. IV, where the Majorana pairs arise naturally due to soft domain walls of SOC. Finally, we make conclusions and discussions in Sec. V.

II Spin-orbit-coupled ss-wave superfluids on 2D optical lattices

We utilize atomic hyperfine states as the pseudospin states |\left|\uparrow\right\rangle and |\left|\downarrow\right\rangle, as illustrated in Fig. 1. The SOC is synthesized by two counter-propagating Raman lasers coupling the two hyperfine states. The single atom motion in 2D real space is described by the Hamiltonian H^0,a=k22m0+δ2σz+(Ωe2ik0r||+h.c.)\hat{H}_{0,a}=\frac{\vec{k}^{2}}{2m_{0}}+\frac{\delta}{2}\sigma_{z}+(\Omega e^{2i\vec{k}_{0}\cdot\vec{r}}\left|\downarrow\right\rangle\left\langle\uparrow\right|+h.c.), where the reduced Planck constant \hbar has been set to be 11, ΩΩ1Ω2\Omega\propto\Omega_{1}\Omega_{2}^{\ast} is the strength of Raman coupling, and k0=k0,xex+k0,yey\vec{k}_{0}=k_{0,x}\vec{e}_{x}+k_{0,y}\vec{e}_{y} is the wave vector of the Raman laser. The off-diagonal terms correspond to a spin flip process accompanied by a momentum transfer of 2k02\vec{k}_{0}, describing the SOC effects. The detuning term reads δ=ωzδω\delta=\omega_{z}-\delta\omega, where ωz>0\omega_{z}>0 is the energy difference between these two hyperfine states, and δω\delta\omega denotes the frequency difference between two Raman laser beams. In the following, we assume that other hyperfine levels are far off-resonance under the two-phonon process, for example, by quadratic Zeeman shift. The Hamiltonian is first transformed by a unitary matrix, namely, H^0,b=UH^0,aU1\hat{H}_{0,b}=U\hat{H}_{0,a}U^{-1}, where U=U=diag(eik0r,eik0r)(e^{-i\vec{k}_{0}\cdot\vec{r}},e^{i\vec{k}_{0}\cdot\vec{r}}). We then perform another pseudo-spin rotations U~=eiπ4σzeiπ4σy\tilde{U}=e^{-i\frac{\pi}{4}\sigma_{z}}e^{-i\frac{\pi}{4}\sigma_{y}} to obtain H^0=U~H^0,bU~\hat{H}_{0}=\tilde{U}\hat{H}_{0,b}\tilde{U}^{\dagger}, which can be written as

H^0=12m0[(kx+k0,xσy)2+(ky+k0,yσy)2]+δ2σyΩσz.\hat{H}_{0}=\frac{1}{2m_{0}}\left[\left(k_{x}+k_{0,x}\sigma_{y}\right)^{2}+\left(k_{y}+k_{0,y}\sigma_{y}\right)^{2}\right]+\frac{\delta}{2}\sigma_{y}-\Omega\sigma_{z}. (1)

The Raman transition produces a desired ERD SOC. In the following, we assume δ=0\delta=0 for convenience.

Refer to caption
Figure 1: Illustration of system setup on a 2D optical lattice, where two counter propagating Raman beams are incident with angle θ\theta. A dip potential is applied on a rectangle geometry (indicated by the blue and red curves) through single-site addressing. Under the configuration θ=π/4\theta=\pi/4, the sign of effective SOC is positive (++)/negative (-) on the blue/red lines in edge coordinate along the arrows. Two spheres (encircled by the red dashed oval) at the interface denote the Majorana corner pair (MCP). The inset above shows the level diagram.
Refer to caption
Figure 2: (a) SF order parameter Δs\Delta_{s} versus interaction UU and chemical potential μ\mu. A phase transition occurs between metal (M) and superfluid (SF) phases. (b) Similar as panel (a) but plotted with Bogliubov quasiparticle energy gap EgE_{g}. In both panels, we take tx=ty=1t_{x}=t_{y}=1, tsox=tsoy=2t_{sox}=t_{soy}=2, hz=1.4h_{z}=1.4.

1D system suffers from strong quantum fluctuations, which could eliminate ss-wave SF order, together with the Majorana modes. This motivates us to investigate the physics in 2D, where quasi-long-range SF order exists below the Berezinskii-Kosterlitz-Thouless (BKT) transition temperature. We concentrate on the lowest (nearly) degenerate bands for constructing a tight-binding model. From Eq. (1), through the operator Ψ^(r)=i,σc^i,σψσ(rri)\hat{\Psi}\left(\vec{r}\right)=\sum_{i,\sigma}\hat{c}_{i,\sigma}\psi_{\sigma}\left(\vec{r}-\vec{r}_{i}\right) with ψσ(rri)\psi_{\sigma}\left(\vec{r}-\vec{r}_{i}\right) the Wannier function at site ii, we obtain a second-quantization formula:

H^0s\displaystyle\hat{H}_{0s} =\displaystyle= i(txc^ic^i+extyc^ic^i+eyitsoxc^iσyc^i+ex\displaystyle\sum_{i}\left(-t_{x}\hat{c}_{i}^{\dagger}\hat{c}_{i+e_{x}}-t_{y}\hat{c}_{i}^{\dagger}\hat{c}_{i+e_{y}}-it_{sox}\hat{c}_{i}^{\dagger}\sigma_{y}\hat{c}_{i+e_{x}}\right. (2)
itsoyc^iσyc^i+ey)+h.c.hzc^iσzc^iμc^ic^i,\displaystyle\left.-it_{soy}\hat{c}_{i}^{\dagger}\sigma_{y}\hat{c}_{i+e_{y}}\right)+\mathrm{h.c.}-h_{z}\hat{c}_{i}^{\dagger}\sigma_{z}\hat{c}_{i}-\mu\hat{c}_{i}^{\dagger}\hat{c}_{i},

where tx=t0cos(k0cosθ)t_{x}=t_{0}\cos\left(k_{0}\cos\theta\right), ty=t0cos(k0sinθ)t_{y}=t_{0}\cos\left(k_{0}\sin\theta\right), tsox=t0sin(k0cosθ)t_{sox}=t_{0}\sin\left(k_{0}\cos\theta\right), tsoy=t0sin(k0sinθ)t_{soy}=t_{0}\sin\left(k_{0}\sin\theta\right), and the bare hopping strength reads t0=𝑑rψσ(rri)(kx2+ky22m0+Vlat)ψσ(rri+1)t_{0}=-\int d\vec{r}\psi_{\sigma}^{\ast}\left(\vec{r}-\vec{r}_{i}\right)\left(\frac{k_{x}^{2}+k_{y}^{2}}{2m_{0}}+V_{\mathrm{lat}}\right)\psi_{\sigma}\left(\vec{r}-\vec{r}_{i+1}\right). Here, we have chosen the basis c^i=(c^i,,c^i,)T\hat{c}_{i}=\left(\hat{c}_{i,\uparrow},\hat{c}_{i,\downarrow}\right)^{T} and denoted hz=Ωh_{z}=\Omega. The incident angle θ\theta is illustrated in Fig. 1. The lattice spacing is set to be a=1a=1. Hereafter, we set t0=1/cos[k0sin(π/4)]=1/cos(2π/4)t_{0}=1/\cos\left[k_{0}\sin\left(\pi/4\right)\right]=1/\cos(\sqrt{2}\pi/4) for convenience. When θ=π/4\theta=\pi/4, we have approximately tsox=tsoy2tx=2tyt_{sox}=t_{soy}\approx 2t_{x}=2t_{y}.

We consider an attractive SU(2)(2)-invariant interaction H^int=iUn^i,n^i,\hat{H}_{int}=-\sum_{i}U\hat{n}_{i,\uparrow}\hat{n}_{i,\downarrow} and study the superfluid phase under mean-field approach by solving the ss-wave superfluid order parameter Δs=Uci,ci,\Delta_{s}=U\left\langle c_{i,\downarrow}c_{i,\uparrow}\right\rangle self-consistently. The Bogoliubov de Genns (BdG) Hamiltonian in the Nambu basis Ψk=(ck,ck,ck,ck)T\Psi_{k}=\left(c_{k\uparrow},c_{k\downarrow},c_{-k\downarrow}^{\dagger},-c_{-k\uparrow}^{\dagger}\right)^{T} is described by H^s=kΨkH(k)Ψk\hat{H}_{s}=\sum_{k}\Psi_{k}^{\dagger}H\left(k\right)\Psi_{k} with

H(k)=(ϵk+γkσy)τzhzσz+Δsτx,H\left(k\right)=\left(\epsilon_{k}+\gamma_{k}\sigma_{y}\right)\tau_{z}-h_{z}\sigma_{z}+\Delta_{s}\tau_{x}, (3)

where ϵk=2(txcoskx+tycosky)μ\epsilon_{k}=-2\left(t_{x}\cos k_{x}+t_{y}\cos k_{y}\right)-\mu, γk=2(tsoxsinkx+tsoysinky)\gamma_{k}=2\left(t_{sox}\sin k_{x}+t_{soy}\sin k_{y}\right), σ\sigma and τ\tau are Pauli matrices acting on the spin and particle-hole spaces, respectively. By minimizing free energy with respect to the order parameter Δs\Delta_{s} and chemical potential μ\mu, we may derive the following self-consistent equations

1\displaystyle 1 =\displaystyle= U2Nlν=±,ktanh(βξk,ν/2)ξk,ν(1+νhz2gk),\displaystyle\frac{U}{2N_{l}}\sum_{\nu=\pm,k}\frac{\tanh\left(\beta\xi_{k,\nu}/2\right)}{\xi_{k,\nu}}\left(1+\frac{\nu h_{z}^{2}}{g_{k}}\right), (4)
nf\displaystyle n_{f} =\displaystyle= 11Nlν=±,kϵktanh(βξk,ν,/2)ξk,ν(1+νmk2gk)\displaystyle 1-\frac{1}{N_{l}}\sum_{\nu=\pm,k}\epsilon_{k}\frac{\tanh\left(\beta\xi_{k,\nu,}/2\right)}{\xi_{k,\nu}}\left(1+\frac{\nu m_{k}^{2}}{g_{k}}\right) (5)

where nfn_{f} is the particle filling factor, NlN_{l} is the number of lattice sites, and other parameters are defined as β=1/(kBT)\beta=1/\left(k_{B}T\right) with kBk_{B} the Boltzmann constant and TT the temperature, ξk,ν=±=ϵk2+Δs2+mk2+2νgk\xi_{k,\nu=\pm}=\sqrt{\epsilon_{k}^{2}+\Delta_{s}^{2}+m_{k}^{2}+2\nu g_{k}}, gk=ϵk2mk2+Δs2hz2g_{k}=\sqrt{\epsilon_{k}^{2}m_{k}^{2}+\Delta_{s}^{2}h_{z}^{2}}, and mk=hz2+|γk|2m_{k}=\sqrt{h_{z}^{2}+\left|\gamma_{k}\right|^{2}}. By numerically solving equations (4) and (5), we obtain phase diagrams at zero temperature for pairing order Δs\Delta_{s} and quasiparticle energy gap EgE_{g} in Fig. 2 (a) and (b), respectively. Fig. 2 (a) confirms the phase transition from a metal (M) phase to an ss-wave SF. From panel (b), we find a finite gap for Bogoliubov quasiparticle excitations in proper parameter region in the SF phase. The energy gap also survives on a finite-size sample and could protect Majorana modes from lower extended states.

III Majorana corner pairs on a topological defective rectangular geometry

Given a proper local dip potential, the defect chain enjoys a non-trivial topology, belonging to the BDI class. It can be characterized through a winding number, which is discussed in Appendix A. Similarly, with a local dip μd\mu_{d}, we can get a defect rectangle in the 2D optical lattice as illustrated in Fig. 1, where SOC domain walls (anticlockwise or clockwise) naturally arise at two corners. In the following, we will first focus on the continuum limit to explore the nature of the emerged Majorana pairs, supplemented with self-consistent numerical calculations on a 2D optical lattice.

We assume that with appropriate μd\mu_{d}, the defect rectangle enters the TSF phase while the rest part remains trivial. As a result, we could assume that the topological defect rectangle is isolated from the 2D bulk. The numerics performed on a 2D optical lattice with an imprinted defect-rectangle also supports this assumption later. From Eq. (3), the low-energy Hamiltonian expands around k=(0,0)\vec{k}=\left(0,0\right) on edges m=\mathrm{m}=I, II, III, IV (see Fig. 1) and is then given by

Hm=tmkm2τz+2ts,mkmσyτzμmτzhzσz+Δs,mτx,H_{\mathrm{m}}=t_{\mathrm{m}}k_{\mathrm{m}}^{2}\tau_{z}+2t_{s,\mathrm{m}}k_{\mathrm{m}}\sigma_{y}\tau_{z}-\mu_{\mathrm{m}}\tau_{z}-h_{z}\sigma_{z}+\Delta_{s,\mathrm{m}}\tau_{x}, (6)

where tI=tIII=tyt_{\mathrm{I}}=t_{\mathrm{III}}=t_{y}, tII=tIV=txt_{\mathrm{II}}=t_{\mathrm{IV}}=t_{x}, kI=kIII=kyk_{\mathrm{I}}=k_{\mathrm{III}}=k_{y}, kII=kIV=kxk_{\mathrm{II}}=k_{\mathrm{IV}}=k_{x}, ts,I=ts,III=2tsoyt_{s,\mathrm{I}}=t_{s,\mathrm{III}}=2t_{soy} and ts,II=ts,IV=2tsoxt_{s,\mathrm{II}}=t_{s,\mathrm{IV}}=2t_{sox}. The on-site chemical potential is μm=μ+2(tx+ty)μd\mu_{\mathrm{m}}=\mu+2\left(t_{x}+t_{y}\right)-\mu_{d} with μd\mu_{d} the dip potential, and Δs,m\Delta_{s,\mathrm{m}} the ss-wave pairing on each edge. Without loss of generality, we set incident angle of Raman lasers θ=π4\theta=\frac{\pi}{4} such that tx=ty=tt_{x}=t_{y}=t, tsox=tsoy=tsot_{sox}=t_{soy}=t_{\mathrm{so}}, μm=μedge=μ+4tμd\mu_{\mathrm{m}}=\mu_{\mathrm{edge}}=\mu+4t-\mu_{d}, and assume the ss-wave SF order parameter is nearly uniform on the four edges Δs,m=Δedge\Delta_{s,\mathrm{m}}=\Delta_{\mathrm{edge}}. For later convenience, we take an “edge coordinate” ss, in which we take the anticlockwise direction as positive. In such a coordinate, the low-energy edge Hamiltonian reads

Hedge=tτz2s2iα(s)σyτzsμedgeτzhzσz+Δedgeτx,H_{\mathrm{edge}}=-t\tau_{z}\frac{\partial^{2}}{\partial s^{2}}-i\alpha\left(s\right)\sigma_{y}\tau_{z}\frac{\partial}{\partial s}-\mu_{\mathrm{edge}}\tau_{z}-h_{z}\sigma_{z}+\Delta_{\mathrm{edge}}\tau_{x}, (7)

with α(s)=2tso\alpha\left(s\right)=-2t_{so}, 2tso2t_{so}, 2tso2t_{so}, 2tso-2t_{so} for edge I-IV respectively. Remarkably, while the terms Δedge\Delta_{\mathrm{edge}} and μedge\mu_{\mathrm{edge}} remain the same on the four edges, the effective coupling α(s)\alpha\left(s\right) changes sign at two of four corners (the corner between the edges I (III) and II (IV)), forming two SOC domain walls as illustrated in Fig. 1. This will give rise to a Majorana pair if hz2>μedge2+Δedge2h_{z}^{2}>\mu_{\mathrm{edge}}^{2}+\Delta_{\mathrm{edge}}^{2}. Specifically, at the corner between edge I and II (corner s=0s=0 in our coordinate), two orthogonal wave functions for MCMs are given by

Ψ0,±=C±eη±|s|(eiϕ±2|y+σ|y±τ+eiϕ±2|yσ|yτ).\Psi_{0,\pm}=C_{\pm}e^{-\eta_{\pm}\left|s\right|}\left(e^{i\frac{\phi_{\pm}}{2}}\left|y_{+}\right\rangle_{\sigma}\left|y_{\pm}\right\rangle_{\tau}+e^{-i\frac{\phi_{\pm}}{2}}\left|y_{-}\right\rangle_{\sigma}\left|y_{\mp}\right\rangle_{\tau}\right). (8)

Here, C±C_{\pm} are normalization constants and eiϕ±=hz[i(αη±Δedge)(tη±2+μedge)](tη±2+μedge)2+(αη±Δedge)2e^{i\phi_{\pm}}=\frac{h_{z}\left[i\left(\alpha\eta_{\pm}\mp\Delta_{\mathrm{edge}}\right)-\left(t\eta_{\pm}^{2}+\mu_{\mathrm{edge}}\right)\right]}{\left(t\eta_{\pm}^{2}+\mu_{\mathrm{edge}}\right)^{2}+\left(\alpha\eta_{\pm}\mp\Delta_{\mathrm{edge}}\right)^{2}}, where η±=122ϰ3+δ124ϰ3aδ+ζ±>0\eta_{\pm}=\frac{1}{2}\sqrt{-\frac{2\varkappa}{3}+\delta}\mp\frac{1}{2}\sqrt{-\frac{4\varkappa}{3a}-\delta+\zeta_{\pm}}>0, ζ±=2d/(a2c/(3a)+δ\zeta_{\pm}=\mp 2d/(a\sqrt{-2c/(3a)+\delta}, δ=23δ13aδ2+4δ13+δ223+δ2+4δ13+δ22332a3\delta=\frac{\sqrt[3]{2}\delta_{1}}{3a\sqrt[3]{\delta_{2}+\sqrt{-4\delta_{1}^{3}+\delta_{2}^{2}}}}+\frac{\sqrt[3]{\delta_{2}+\sqrt{-4\delta_{1}^{3}+\delta_{2}^{2}}}}{3\sqrt[3]{2a}}, δ1=ϰ2+12ae\delta_{1}=\varkappa^{2}+12ae, δ2=2ϰ3+27ad272aϰe\delta_{2}=2\varkappa^{3}+27ad^{2}-72a\varkappa e, a=t2a=t^{2}, ϰ=α2+2tμedge\varkappa=\alpha^{2}+2t\mu_{\mathrm{edge}}, d=2αΔedged=-2\alpha\Delta_{\mathrm{edge}} and e=Δedge2hz2+μedge2e=\Delta_{\mathrm{edge}}^{2}-h_{z}^{2}+\mu_{\mathrm{edge}}^{2}. The vectors |y±σ\left|y_{\pm}\right\rangle_{\sigma} and |y±τ\left|y_{\pm}\right\rangle_{\tau} are eigenstates of operators σy\sigma_{y} and τy\tau_{y}, respectively. Following similar approach, we could also find two Majorana modes at the corner between edges III and IV (see Appendix B for details). We emphasize that as long as the four edges are in the TSF phase, the very existence of Majorana pairs is robust against the fluctuations of chemical potential and SF order parameter.

Refer to caption
Figure 3: (a) Self-consistent ss-wave pairing order parameters in real space. (b) Energy level diagrams. (c,d) The real-space distributions of MZMs. For all panels, we choose tx=ty=1,tsox=tsoy=2t_{x}=t_{y}=1,t_{sox}=t_{soy}=2, μ=7.95\mu=-7.95, μd=4.0\mu_{d}=-4.0, h=1.4h=1.4, U=6.3U=6.3. The incident angle is θ=π/4\theta=\pi/4 in (a)-(c) and θ=π/4\theta=-\pi/4 in (d).

With the above understanding of continuum systems, we now proceed to study the discrete cases on an optical lattice shown in Fig. 1. The total Hamiltonian now becomes

H^BdG=H^s+iμdc^ic^i,\hat{H}_{BdG}=\hat{H}_{s}+\sum_{i\in\square}\mu_{d}\hat{c}_{i}^{\dagger}\hat{c}_{i}, (9)

where ii\in\square enumerate each site with the dip potential (a rectangular geometry in this case). The local ss-wave superfluid order parameter in real space is determined in a self-consistent manner Jiang2016 , as well as the quasiparticle energy spectra and wave functions. On the defect rectangle, the system is topological once hz>μ~2+Δs2h_{z}>\sqrt{\tilde{\mu}^{2}+\Delta_{s}^{2}} and μ~=μ+2tx+2tyμd\tilde{\mu}=\mu+2t_{x}+2t_{y}-\mu_{d}. In our self-consistent numerical calculations, we take the lattice sizes nx=ny=30n_{x}=n_{y}=30, and the defect rectangle is given by nxd=nyd=22n_{x}^{d}=n_{y}^{d}=22. The SF order parameter Δs,i\Delta_{s,i} is shown in Fig. 3 (a), which has a constant phase across the entire system. Fig. 3 (b) shows the quasiparticle energy spectrum, where four Majorana bound states (two Majorana corner pairs) exist in the energy gap. A small energy splitting is observed as a result of finite-size effect. Fig. 3 (c) shows the density distribution of the bounded Majorana corner states, which clearly demonstrates its localization at the corners of the defect rectangle. The Majorana corner pairs are robust against the perturbations of chemical potential and SF order parameter that preserve chiral symmetry. We have confirmed this point by numerical calculations.

The incident angle of Raman lasers can change the SOC and the nearest-neighbor hopping, and thus alter the Majorana bound states. Figs. 4 (a) and (b) illustrate the corresponding phase diagram with respect to μ\mu-θ\theta and μd\mu_{d}-θ\theta, where Majorana corner pairs exist in the topological region (T). It is found that the Majorana pairs is also robust to certain variation of the incident angle θ\theta. We remark that if the sign of θ\theta is reversed, the Majorana pairs appear at another two corners (the interfaces of II-III and I-IV) as shown in Fig. 3(d), which can be compared with Fig. 3(c). Thus, our proposed setup provides better tunability for manipulating Majorana bound states.

Refer to caption
Figure 4: Phase diagram for 2D system with the defect rectangle. In region “T”, Majorana bound states exist at corners. In (a), μd=4.0\mu_{d}=-4.0, hz=1.4h_{z}=1.4, U=6.3U=6.3 are used. In (b), μ=7.8\mu=-7.8, hz=1.4h_{z}=1.4, U=6.3U=6.3 are used.

IV Majorana corner pairs on a ring geometry

In this section, we study the case with a ring-shaped defect line. Here, the optical lattice is removed and we focus on the low-energy effective 1D model for simplicity. The effective model is illustrated in Fig. 5 (a) and we find that soft domain walls of SOC naturally arise on the ring, which leads to the emergence of Majorana corner pairs.

Without loss of generality, we assume the momentum kick by Raman lasers is along the xx direction. Under a spin-rotation σy\sigma_{y}\rightarrow σx\sigma_{x} with k0,y=0k_{0,y}=0 in Eq. (1), the SOC has the form α0kxσx\alpha_{0}k_{x}\sigma_{x}, where the coupling constant is given by the ratio of laser wavevector and atomic mass, i.e., α0=k0/m0\alpha_{0}=k_{0}/m_{0}. Hence, in the continuum limit, the effective Hamiltonian reads

=k22m0τz+α0kxσxτzμτz+hzσz+Δsτx,\mathcal{H}=\frac{k^{2}}{2m_{0}}\tau_{z}+\alpha_{0}k_{x}\sigma_{x}\tau_{z}-\mu\tau_{z}+h_{z}\sigma_{z}+\Delta_{s}\tau_{x}, (10)

where Δs\Delta_{s} is an ss-wave SF order. For simplicity, we set Δs\Delta_{s} to be real. The relation hz>μ2+Δs2h_{z}>\sqrt{\mu^{2}+\Delta_{s}^{2}} holds in the topological regions.

In a polar coordinate (ρ,ϕ)(\rho,\phi), the above Hamiltonian becomes a function of polar angle ϕ\phi on a ring with given radii ρ\rho, i.e.,

(ϕ)=ηϕ2τz+iα~0sinϕϕσxτzμτz+hzσz+Δsτx,\mathcal{H}\left(\phi\right)\mathcal{=}-\eta\partial_{\phi}^{2}\tau_{z}+i\tilde{\alpha}_{0}\sin\phi\frac{\partial}{\partial\phi}\sigma_{x}\tau_{z}-\mu^{\prime}\tau_{z}+h_{z}\sigma_{z}+\Delta_{s}\tau_{x}, (11)

where η=12m0ρ2\eta=\frac{1}{2m_{0}\rho^{2}}, α~0=α0ρ\tilde{\alpha}_{0}=\frac{\alpha_{0}}{\rho} and μ=μk022m0\mu^{\prime}=\mu-\frac{k_{0}^{2}}{2m_{0}} (more details are discussed in Appendix 7). The Hamiltonian (ϕ)\mathcal{H}\left(\phi\right) has particle-hole symmetry 𝒫(ϕ)𝒫1=(ϕ)\mathcal{PH}\left(\phi\right)\mathcal{P}^{-1}=-\mathcal{H}\left(-\phi\right) where 𝒫=σyτy𝒦\mathcal{P}=\sigma_{y}\tau_{y}\mathcal{K} and 𝒦\mathcal{K} denotes the complex conjugation. It also preserves a generalized time-reversal symmetry 𝒯(ϕ)𝒯1=(ϕ)\mathcal{TH}\left(\phi\right)\mathcal{T}^{-1}=\mathcal{H}\left(-\phi\right) with 𝒯=σz𝒦\mathcal{T}=\sigma_{z}\mathcal{K}. The combination of 𝒫\mathcal{P} and 𝒯\mathcal{T} leads to the chiral symmetry 𝒞\mathcal{C}: 𝒞(ϕ)𝒞1=(ϕ)\mathcal{CH}\left(\phi\right)\mathcal{C}^{-1}=-\mathcal{H}\left(\phi\right), with 𝒞=𝒫𝒯=iσxτy\mathcal{C}=\mathcal{PT}=i\sigma_{x}\tau_{y}. Therefore, the Hamiltonian belongs to BDI\mathrm{BDI} class and can be characterized by a \mathbb{Z} topological invariant Altland1997 ; Schnyder2008 .

Refer to caption
Figure 5: (a) Illustration of a ring trap in which the atoms are confined. Two counter-propagating Raman lasers with wavevector ±k0\pm\vec{k}_{0} couple the atomic hyperfine states. The notations ++ and - indicate the sign of SOC. Two spheres encircled by the red dashed circle denote a MCP. (b) The SOC term α(ϕ)\alpha\left(\phi\right) plotted to varying polar angle ϕ\phi. (c) The eigenspectrum computed through plane-wave expansion. (d) The particle density ρMZ\rho_{{}_{MZ}} of Majorana zero modes versus the angle ϕ\phi. The parameters are chosen as α~0=0.04\tilde{\alpha}_{0}=0.04, η=0.005,μ=0.2,hz=1.4,Δs=0.8\eta=0.005,\mu^{\prime}=0.2,h_{z}=-1.4,\Delta_{s}=0.8.

In Eq. (11), the SOC α(ϕ)=α~0sinϕ\alpha\left(\phi\right)=\tilde{\alpha}_{0}\sin\phi changes sign at  ϕ=0,π\phi=0,\pi, as shown in Fig. 5 (b). Specifically, we have α(ϕ)>\alpha\left(\phi\right)> 0 if ϕ(0,π)\phi\in\left(0,\pi\right) and α(ϕ)<0\alpha\left(\phi\right)<0 if ϕ(π,2π)\phi\in\left(\pi,2\pi\right). Hence, the system can be divided into two segments. Both belong to the BDI\mathrm{BDI} class but possess opposite topological invariant. The interfaces are determined by ϕ=0\phi=0 and π\pi, corresponding to two “soft” domain walls in the sense that the SOC term changes smoothly across these two points. From Eq. (11), the Hamiltonian (ϕ)\mathcal{H}\left(\phi\right) is invariant under a 2nπ2n\pi rotation if nn is an integer. Therefore, to solve the eigenvalues of (ϕ)\mathcal{H}\left(\phi\right), we assume the following trial solution,

Φ(ϕ)=(ua(ϕ),ub(ϕ),uc(ϕ),ud(ϕ))T,\Phi\left(\phi\right)=\left(\begin{array}[]{cccc}u_{a}\left(\phi\right),&u_{b}\left(\phi\right),&u_{c}\left(\phi\right),&u_{d}\left(\phi\right)\end{array}\right)^{T}, (12)

where uν=a,b,c,d(ϕ)=mνmeimϕu_{\nu=a,b,c,d}\left(\phi\right)=\sum_{m}\nu_{m}e^{im\phi} and mm is an integer. By solving the Schrödinger equation (ϕ)Φ(ϕ)=EΦ(ϕ)\mathcal{H}\left(\phi\right)\Phi\left(\phi\right)=E\Phi\left(\phi\right), the eigenvalues are obtained as shown in Fig. 5 (c). See Appendix 7 for more details. It is clear that four Majorana modes emerge (with an numerical error about E104E\approx 10^{-4}). One Majorana corner pair consisting of two Majorana modes localizes at ϕ=0\phi=0, and the other pair localizes at ϕ=π\phi=\pi, as illustrated in Fig. 5 (a). This is also demonstrated by the particle density distribution ρMZ\rho_{{}_{MZ}} of Majorana modes, as shown in Fig. 5 (d). We remark that a toroidal Bose-Einstein condensate has been created in an all-optical trap Ramanathan2011 . We expect our scheme could be reached with similar techniques and additional Raman lasers.

V Discussion and Conclusion

From the effective low-energy theory of TSFs, it is well-known that Majorana modes would emerge if the sign of the Dirac mass changes and most previous proposals are based on this principle. In this paper, we propose an alternative approach to implement Majorana modes (Majorana corner pairs) through tuning the effective SOC. By loading Fermi gases on 2D optical lattices subjected to a 1D ERD SOC, we can find a SF phase under appropriate ss-wave interaction and Zeeman field. Using single-site addressing techniques, we could engineer defective geometries, which are topologically non-trivial, on the 2D optical lattice. From the viewpoint of low-energy theories, a defect rectangle consists of two TSFs characterized by distinct topological invariants whose sign is determined by the sign of SOC in edge coordinate. Obviously, the sign of SOC changes at two corners on the defect rectangle. At the interface of two distinct TSF, a topologically protected Majorana pair naturally arises according to the index theorem. For TSF with 1D ERD SOC on a ring, two soft SOC domain walls exist, and two Majorana pairs also appear near the domain walls. In principle, as long as two effective 1D SFs are topological with different topological invariants w=±1w=\pm 1, the Majorana pair will emerge at the interface. It is robust as long as the perturbations preserve three underlying symmetries (𝒫\mathcal{P}, 𝒯\mathcal{T}, 𝒞\mathcal{C}) of the system.

We emphasize that the Majorana corner pair in the context differs from those in second order TSCs in two dimensions. First, for second order TSCs with time reversal symmetry Yan2018 ; Wang2018 , 1D edge modes evolve from the higher-dimensional bulk of the topological insulators. However, in our scheme, the 1D modes originate from the defect geometry. Second, in a higher-order TSC, a momentum-dependent SC pairing (s±s_{\pm} or dd-wave) leads to Dirac mass kink at the corner of the sample, and then induces the Majorana Kramers pair. In contrary, our proposal utilizes the sign reverse of effective SOC on the edges and lacks Kramers degeneracy.

In summary, we propose a distinct scheme to implement Majorana pairs in an atomic platform. The coordinate of Majorana pair depends on the position of SOC domain wall which can be tuned by the directions of the Raman laser beams. Moreover, our system is free of dynamical instability such that the MZM has a longer lifetime. Our work opens the possibility of implementing robust Majorana pairs and the associated non-Abelian braiding in cold atoms.

Acknowledgements.
This work is supported by Air Force Office of Scientific Research (FA9550-16-1-0387), National Science Foundation (PHY-1806227), and Army Research Office (W911NF-17-1-0128). This work is also supported in part by NSFC under the grant No. 11504285, and the Scientific Research Program Funded by Natural Science Basic Research Plan in Shaanxi Province of China (Program Nos. 2018JQ1058), the Scientific Research Program Funded by Shaanxi Provincial Education Department under the grant No. 18JK0397, and the scholarship from China Scholarship Council (CSC) (Program No. 201708615072).

Appendix A Majorana modes at the ends of topological defect-chain

In this section, we show a topologically non-trivial defective chain can be implemented through on-site potential engineering on 2D optical lattices.

For a 1D system with SOC and SF order, the system is topological if hz>(μ~μd)2+Δs2h_{z}>\sqrt{\left(\tilde{\mu}-\mu_{d}\right)^{2}+\Delta_{s}^{2}} , where μ~=μ+2tx+2ty\tilde{\mu}=\mu+2t_{x}+2t_{y} Lutchyn2010 ; Oppen2010 ; Jiang2016 ; Sau2010 . Reference Jiang2016 shows Majorana fermions may be generated in a 2D optical lattices with 1D ERD SOC along the xx direction. This motivates us to demonstrate the existence of Majorana bound states in a genuine 2D systems with 1D defects, where the SOC lays along ex+ey\vec{e}_{x}+\vec{e}_{y} direction. Through single-site addressing, a potential could be locally applied to a given site.

Imposing a 1D potential dip μd\mu_{d}, we have the following Hamiltonian

H^BdG=H^s+iμdc^ic^i,\hat{H}_{BdG}=\hat{H}_{s}+\sum_{i\in-}\mu_{d}\hat{c}_{i}^{\dagger}\hat{c}_{i}, (13)

where ii\in- denotes the sites (ix,iy)(i_{x},i_{y}) satisfying iy=nyci_{y}=n_{y_{c}}. In self-consistent numerical calculations, we take a lattice with nx=100n_{x}=100, ny=9n_{y}=9 and nyc=5n_{y_{c}}=5 under open boundary conditions, where nxn_{x} and nyn_{y} denote the site number along the xx and yy directions. Figs. 6 (a) and (b) present self-consistent  numerical results. Fig. 6 (b) shows density profile of the zero-energy mode (E104E\approx 10^{-4}). It demonstrates the existence of MZMs even in a genuine 2D system. From self-consistent BdG numerical results, the SF order parameter is almost homogeneous along the xx direction as shown in Fig. 6 (a). Thus, with periodic boundary condition, the system has a translation symmetry along the xx direction so that momentum kxk_{x} is a good quantum number. The 2D optical lattice can be regarded as layered 1D chain with transverse tunneling and SOC effects. The effective Hamiltonian is then written as

HBdG(kx)\displaystyle H_{BdG}(k_{x}) =\displaystyle= κ0h0(kx)+μdκcσ0τz+Δaκcσyτy\displaystyle\kappa_{0}h_{0}\left(k_{x}\right)+\mu_{d}\kappa_{c}\sigma_{0}\tau_{z}+\Delta_{a}\kappa_{c}\sigma_{y}\tau_{y} (14)
tyκxσ0τztsoyκyσyτz.\displaystyle-t_{y}\kappa_{x}\sigma_{0}\tau_{z}-t_{soy}\kappa_{y}\sigma_{y}\tau_{z}.

Here, the matrix κ\kappa acts on chain space, with κ0\kappa_{0} identity matrix, (κc)i,i=1\left(\kappa_{c}\right)_{i,i}=1 for i=nyci=n_{y_{c}} and 0 otherwise, (κx)i,j=1\left(\kappa_{x}\right)_{i,j}=1 for |ij|=1\left|i-j\right|=1 and (κy)i,i1=i\left(\kappa_{y}\right)_{i,i\mp 1}=\mp i and 0 otherwise. The term proportional to μd\mu_{d} (Δa\Delta_{a}) describes the dip potential (the SF-order) difference between the central chain and other individual chains. The term proportional to tyt_{y} (tsoyt_{soy}) describes the hopping (SOC) along the yy direction. The following Hamiltonian

h0(kx)\displaystyle h_{0}\left(k_{x}\right) =\displaystyle= (2txcoskxμ)σ0τz+2tsoxsinkxσyτz\displaystyle\left(-2t_{x}\cos k_{x}-\mu\right)\sigma_{0}\tau_{z}+2t_{sox}\sin k_{x}\sigma_{y}\tau_{z} (15)
+hzσzτz+Δsσyτy\displaystyle+h_{z}\sigma_{z}\tau_{z}+\Delta_{s}\sigma_{y}\tau_{y}

describes the original uniform individual 1D chain along the xx direction. σ\sigma and τ\tau are Pauli matrices acting on spin space and particle-hole space, respectively. The above BDG Hamiltonian HBdG(kx)H_{BdG}(k_{x}) has intrinsic particle-hole symmetry 𝒫\mathcal{P}: 𝒫HBdG(kx)𝒫1=HBdG(kx)\mathcal{P}H_{BdG}\left(k_{x}\right)\mathcal{P}^{-1}=-H_{BdG}\left(-k_{x}\right) with 𝒫=τ~x𝒦\mathcal{P}=\tilde{\tau}_{x}\mathcal{K}, τ~x=τxσ0η0\tilde{\tau}_{x}=\tau_{x}\sigma_{0}\eta_{0}, where σ0\sigma_{0} is 2×22\times 2 identity matrix, η0\eta_{0} is a Ns×NsN_{s}\times N_{s} identity matrix acting on the lattice site space, and 𝒦\mathcal{K} is the complex conjugation. If the superfluid order parameter is real (or has a constant phase that can be eliminated by gauge transformations), the Hamiltonian preserves a generalized time-reversal symmetry 𝒯\mathcal{T}: 𝒯HBdG(kx)𝒯1=HBdG(kx)\mathcal{T}H_{BdG}\left(k_{x}\right)\mathcal{T}^{-1}=H_{BdG}\left(-k_{x}\right) with 𝒯=𝒦\mathcal{T}=\mathcal{K}. The composite operation of 𝒫\mathcal{P} and 𝒯\mathcal{T} also leads to a chiral symmetry 𝒞\mathcal{C}: 𝒞HBdG(kx)𝒞1=HBdG(kx)\mathcal{C}H_{BdG}\left(k_{x}\right)\mathcal{C}^{-1}=-H_{BdG}\left(k_{x}\right), with 𝒞=𝒫𝒯=τ~x\mathcal{C}=\mathcal{P}\mathcal{T}=\tilde{\tau}_{x}. From above symmetry analyses, the Hamiltonian belongs to BDI\mathrm{BDI} class, characterized by a \mathbb{Z} topological invariant (winding number).

Refer to caption
Figure 6:  (a) The amplitude of SF order parameter in real space. (b) The density distribution of the MZM. (c,d) z(kx)z(k_{x}) in the complex plane for opposite winding number. In all panels, we set tx=ty=1t_{x}=t_{y}=1, hz=1.4h_{z}=1.4, μ=8.0\mu=-8.0, μd=4\mu_{d}=-4, Δs=0.3\Delta_{s}=0.3 and Δa=0.11\Delta_{a}=0.11. We choose tsox=tsoy=2t_{sox}=t_{soy}=2 in (a-c) and tsox=tsoy=2t_{sox}=t_{soy}=-2 in (d).

The winding number ww can characterize the topological properties of BdG Hamiltonian (14) Tewari2012 . Because the BdG Hamiltonian HBDGH_{BDG} has the chiral symmetry, it can be transformed into an off-diagonal form in particle-hole space under a unitary transformation U=eiπ4τyU=e^{-i\frac{\pi}{4}\tau_{y}},

UHBdG(kx)U1=(0(kx)T(kx)0).UH_{BdG}(k_{x})U^{-1}=\left(\begin{array}[]{cc}0&\mathcal{B}\left(k_{x}\right)\\ \mathcal{B}^{T}\left(-k_{x}\right)&0\end{array}\right). (16)

Here, (kx)=1(kx)i2\mathcal{B}\left(k_{x}\right)=\mathcal{B}_{1}\left(k_{x}\right)-i\mathcal{B}_{2}, where 1(kx)=(2txcoskxμ)κ0σ0+2tsoxsinkxκ0σy+hzκ0σz+μdκcσ0tyκxσ0tsoyκyσy\mathcal{B}_{1}\left(k_{x}\right)=\left(-2t_{x}\cos k_{x}-\mu\right)\kappa_{0}\sigma_{0}+2t_{sox}\sin k_{x}\kappa_{0}\sigma_{y}+h_{z}\kappa_{0}\sigma_{z}+\mu_{d}\kappa_{c}\sigma_{0}-t_{y}\kappa_{x}\sigma_{0}-t_{soy}\kappa_{y}\sigma_{y} and 2=(Δsκ0σy+Δaκcσy)\mathcal{B}_{2}=\left(\Delta_{s}\kappa_{0}\sigma_{y}+\Delta_{a}\kappa_{c}\sigma_{y}\right). The winding number is defined as Tewari2012

w=iπ0πdzz(kx),w=-\frac{i}{\pi}\int_{0}^{\pi}\frac{dz}{z\left(k_{x}\right)}, (17)

where z(kx)=det((kx))/|det((kx))|z\left(k_{x}\right)=\det\left(\mathcal{B}\left(k_{x}\right)\right)/\left|\det\left(\mathcal{B}\left(k_{x}\right)\right)\right|. As shown in Figs. 6 (c) and (d) with hz>(μ+2tx+2tyμd)2+Δs2h_{z}>\sqrt{\left(\mu+2t_{x}+2t_{y}-\mu_{d}\right)^{2}+\Delta_{s}^{2}}, the complex value of z(kx)z\left(k_{x}\right) varies when kxk_{x} changes from 0 to π\pi, indicating |w|=1\left|w\right|=1. By considering the trajectory of z(kx)z\left(k_{x}\right) in the complex plane as kxk_{x} changing from 0 to π\pi, z(kx)z\left(k_{x}\right) moves from a point on the negatively real axis to the positive axis while crossing the imaginary axis exactly once. It is clear that the winding number w=1w=-1 when tsox>0t_{sox}>0 in topological phase, as shown in Fig. 6 (c), and the winding number w=+1w=+1 when tsox<0t_{sox}<0 in topological phase, as shown in Fig. 6 (d). Namely, the sign of winding number for the defect chain is determined by the sign of SOC in the topological phase.

Appendix B Low-energy theory of topological superfluids on a defective rectangle

Remarkably, from Eq. (7) the term Δedge\Delta_{\mathrm{edge}} doesn’t change sign, but the coefficient α(s)\alpha\left(s\right) changes sign at two corners of defect-rectangle. This will give rise to a Majorana pair at the corner where α(s)\alpha\left(s\right) changes sign. Hereafter, we will give the analytic solutions of Majorana corner modes.

According to the particle-hole symmetry of HedgeH_{\mathrm{edge}}, i.e., {Hedge,σyτy}=0\left\{H_{\mathrm{edge}},\sigma_{y}\tau_{y}\right\}=0, we have

Hedgeσyτy=σyτyHedge.H_{\mathrm{edge}}\sigma_{y}\tau_{y}=-\sigma_{y}\tau_{y}H_{\mathrm{edge}}. (18)

It can be concluded that if there exist zero-energy states of HedgeH_{\mathrm{edge}}, these states are also eigenstates of σyτy\sigma_{y}\tau_{y}. Therefore, we assume the zero-energy wave functions in the “edge coordinate” ss have the following forms:

Ψ0,+\displaystyle\Psi_{0,+} =\displaystyle= f+(s)|y+σ|y+τ+g+(s)|yσ|yτ,\displaystyle f_{+}\left(s\right)\left|y_{+}\right\rangle_{\sigma}\left|y_{+}\right\rangle_{\tau}+g_{+}\left(s\right)\left|y_{-}\right\rangle_{\sigma}\left|y_{-}\right\rangle_{\tau}, (19)
Ψ0,\displaystyle\Psi_{0,-} =\displaystyle= f(s)|y+σ|yτ+g(s)|yσ|y+τ.\displaystyle f_{-}\left(s\right)\left|y_{+}\right\rangle_{\sigma}\left|y_{-}\right\rangle_{\tau}+g_{-}\left(s\right)\left|y_{-}\right\rangle_{\sigma}\left|y_{+}\right\rangle_{\tau}. (20)

Then we have σyτyΨ0,±=±Ψ0,±\sigma_{y}\tau_{y}\Psi_{0,\pm}=\pm\Psi_{0,\pm} and the Schrödinger equation at the corner between the edge I and II (corner s=0s=0 ) is HedgeΨ0,+=0.H_{\mathrm{edge}}\Psi_{0,+}=0. Using the eigenvector Φ=(f+(s),g+(s))T\Phi=\left(f_{+}\left(s\right),g_{+}\left(s\right)\right)^{T}, the above equation (18) can be rewritten as

(ts2+iαsσz+μedgeiΔedgeσz+hzσx)Φ=0.\left(t\partial_{s}^{2}+i\alpha\partial_{s}\sigma_{z}+\mu_{\mathrm{edge}}-i\Delta_{\mathrm{edge}}\sigma_{z}+h_{z}\sigma_{x}\right)\Phi=0. (21)

We assume f+(s)=A+eη+sf_{+}\left(s\right)=A_{+}e^{-\eta_{+}s}, g+(s)=B+eη+sg_{+}\left(s\right)=B_{+}e^{-\eta_{+}s} (s>0s>0) and write

(itη+2σz+αη+iμedgeσzΔedge+hzσy)Φ~=0,\left(-it\eta_{+}^{2}\sigma_{z}+\alpha\eta_{+}-i\mu_{\mathrm{edge}}\sigma_{z}-\Delta_{\mathrm{edge}}+h_{z}\sigma_{y}\right)\tilde{\Phi}=0, (22)

with Φ~=(A+,B+)T\tilde{\Phi}=\left(A_{+},B_{+}\right)^{T}. According to the vanishing determinant of the above matrix, we obtain

η+=122ϰ3+δ124ϰ3aδ+ζ+>0,\eta_{+}=\frac{1}{2}\sqrt{-\frac{2\varkappa}{3}+\delta}-\frac{1}{2}\sqrt{-\frac{4\varkappa}{3a}-\delta+\zeta_{+}}>0, (23)

where ϰ\varkappa, δ\delta, aa, and ζ+\zeta_{+} have been written explicitly in the main text. Then we have A+/B+=eiϕ+\ A_{+}/B_{+}=e^{i\phi_{+}} with

eiϕ+=hz[(tη+2+μedge)+i(αη+Δedge)](tη+2+μedge)2+(αη+Δedge)2.e^{i\phi_{+}}=\frac{h_{z}\left[-\left(t\eta_{+}^{2}+\mu_{\mathrm{edge}}\right)+i\left(\alpha\eta_{+}-\Delta_{\mathrm{edge}}\right)\right]}{\left(t\eta_{+}^{2}+\mu_{\mathrm{edge}}\right)^{2}+\left(\alpha\eta_{+}-\Delta_{\mathrm{edge}}\right)^{2}}. (24)

At last, we get the MZM Ψ0,+\Psi_{0,+} at the corner between the edge I and II. Following similar approach as in previous calculations, we can get another zero energy solution Ψ0,=Ceη|s|(eiϕ2|y+σ|yτ+eiϕ2|yσ|y+τ)\Psi_{0,-}=C_{-}e^{-\eta_{-}\left|s\right|}\left(e^{i\frac{\phi_{-}}{2}}\left|y_{+}\right\rangle_{\sigma}\otimes\left|y_{-}\right\rangle_{\tau}+e^{-i\frac{\phi_{-}}{2}}\left|y_{-}\right\rangle_{\sigma}\otimes\left|y_{+}\right\rangle_{\tau}\right) with CC_{-} the normalization constant, where

η\displaystyle\eta_{-} =\displaystyle= 122ϰ3+δ+124ϰ3aδ+ζ>0,\displaystyle\frac{1}{2}\sqrt{-\frac{2\varkappa}{3}+\delta}+\frac{1}{2}\sqrt{-\frac{4\varkappa}{3a}-\delta+\zeta_{-}}>0, (25)
eiϕ\displaystyle e^{i\phi_{-}} =\displaystyle= hz[(tη2+μedge)+i(αη+Δedge)](tη2+μedge)2+(αη+Δedge)2,\displaystyle\frac{h_{z}\left[-\left(t\eta_{-}^{2}+\mu_{\mathrm{edge}}\right)+i\left(\alpha\eta_{-}+\Delta_{\mathrm{edge}}\right)\right]}{\left(t\eta_{-}^{2}+\mu_{\mathrm{edge}}\right)^{2}+\left(\alpha\eta_{-}+\Delta_{\mathrm{edge}}\right)^{2}}, (26)

and the coefficients ϰ\varkappa, δ\delta, aa and ζ\zeta_{-} are listed in the main text. In summary, there are two Majorana modes (a Majorana corner pair) localized around the corner with analytic solution given in Eq. (8). Regarding the corner between III and IV, there exists another SOC domain wall and we similarly have a Majorana pair there.

Appendix C Low-energy theory of topological superfluids on a defective ring

Refer to caption
Figure 7: The  real and imaginary parts of coefficients (am,bm,cm,dm)\left(a_{m},b_{m},c_{m},d_{m}\right) for the wave functions of four Majorana modes, respectively. Parameters α~0=0.04\tilde{\alpha}_{0}=0.04, η=0.005,μ=0.2,hz=1.4,Δ=0.8\eta=0.005,\mu^{\prime}=0.2,h_{z}=-1.4,\Delta=0.8 are used for all panels.

The general Hamiltonian for a SOC fermi gas with Cooper pairing is given by

Hr=12m0(kk0σ)2μτz+hzσz+Δsτx,H_{\mathrm{r}}=\frac{1}{2m_{0}}\left(\vec{k}-k_{0}\vec{\sigma}\right)^{2}-\mu\tau_{z}+h_{z}\sigma_{z}+\Delta_{s}\tau_{x}, (27)

where k=i\vec{k}=-i\vec{\nabla}. Assuming spin σ\vec{\sigma} is along xx and neglecting the constant energy shift k02/(2m0)k_{0}^{2}/(2m_{0}), we get the Hamiltonian in Eq. (10). In the polar coordinate,

(ij)=(cosϕsinϕsinϕcosϕ)(eρeφ).\left(\begin{array}[]{c}\vec{i}\\ \vec{j}\end{array}\right)=\left(\begin{array}[]{cc}\cos\phi&-\sin\phi\\ \sin\phi&\cos\phi\end{array}\right)\left(\begin{array}[]{c}\vec{e}_{\rho}\\ \vec{e}_{\varphi}\end{array}\right). (28)

The Laplace operator in Descartes and polar coordinates is written as

=xi+yj=ρeρ+1ρϕeϕ.\vec{\bigtriangledown}=\partial_{x}\vec{i}+\partial_{y}\vec{j}=\partial_{\rho}\vec{e}_{\rho}+\frac{1}{\rho}\partial_{\phi}\vec{e}_{\phi}. (29)

Then we have k2==(ρ2+1ρ2ϕ2)\vec{k}^{2}=-\vec{\bigtriangledown}\cdot\vec{\bigtriangledown}=-\left(\partial_{\rho}^{2}+\frac{1}{\rho^{2}}\partial_{\phi}^{2}\right) and

k(σxi)=i[cosϕρ+1ρϕ(sinϕ)]σx.\vec{k}\cdot\left(\sigma_{x}\vec{i}\right)=-i\left[\cos\phi\partial_{\rho}+\frac{1}{\rho}\partial_{\phi}\left(-\sin\phi\right)\right]\sigma_{x}. (30)

Finally, Eq. (10) can be rewritten as the following form:

\displaystyle\mathcal{H} =\displaystyle= 12m0[(ρ2+1ρ2ϕ2)ik0(cosϕρ1ρsinϕϕ)\displaystyle\frac{1}{2m_{0}}\left[-\left(\partial_{\rho}^{2}+\frac{1}{\rho^{2}}\partial_{\phi}^{2}\right)-ik_{0}\left(\cos\phi\partial_{\rho}-\frac{1}{\rho}\sin\phi\partial_{\phi}\right)\right. (31)
×σxik0(cosϕρ1ρsinϕϕ)σx]τzμτz\displaystyle\left.\times\sigma_{x}-ik_{0}\left(\cos\phi\partial_{\rho}-\frac{1}{\rho}\sin\phi\partial_{\phi}\right)\sigma_{x}\right]\tau_{z}-\mu\tau_{z}
+hzσz+Δsτx.\displaystyle+h_{z}\sigma_{z}+\Delta_{s}\tau_{x}.

Consider ultracold atoms trapped in a ring-shaped trapping potential, where the radii ρ\rho is fixed. Terms with respect to ρ\partial_{\rho} disappear. After substituting η=12m0ρ2,\eta=\frac{1}{2m_{0}\rho^{2}}, α~0=k0m0ρ=α0ρ\tilde{\alpha}_{0}=\frac{k_{0}}{m_{0}\rho}=\frac{\alpha_{0}}{\rho}, and μ=μk022m0\mu^{\prime}=\mu-\frac{k_{0}^{2}}{2m_{0}}, the Hamiltonian \mathcal{H} becomes Eq. (11).

Because (ϕ)\mathcal{H}\left(\phi\right) is invariant under a 2nπ2n\pi rotation with nn being an integer, we assume the wave functions take following form as Φ(ϕ)=(ua(ϕ),ub(ϕ),uc(ϕ),ud(ϕ))T\Phi\left(\phi\right)=\left(u_{a}\left(\phi\right),u_{b}\left(\phi\right),u_{c}\left(\phi\right),u_{d}\left(\phi\right)\right)^{T}, where uν=a,b,c,d(ϕ)=m=NNνmeimϕu_{\nu=a,b,c,d}\left(\phi\right)=\sum_{m=-N}^{N}\nu_{m}e^{im\phi}. Plugging (ϕ)\mathcal{H}\left(\phi\right) and Φ(ϕ)\Phi\left(\phi\right) into the Schrödinger equation Φ(ϕ)=EΦ(ϕ)\mathcal{H}\Phi\left(\phi\right)=E\Phi\left(\phi\right), and matching coefficients for eimϕe^{im\phi}, one can obtain a series of coupled equations as

Eam\displaystyle Ea_{m} =\displaystyle= (η2m2μ+hz)am+iα~02(m1)bm1\displaystyle\left(\eta^{2}m^{2}-\mu^{\prime}+h_{z}\right)a_{m}+i\frac{\tilde{\alpha}_{0}}{2}\left(m-1\right)b_{m-1} (32)
iα~02(m+1)bm+1+Δscm,\displaystyle-i\frac{\tilde{\alpha}_{0}}{2}\left(m+1\right)b_{m+1}+\Delta_{s}c_{m},
Ebm\displaystyle Eb_{m} =\displaystyle= iα~02(m1)am1iα~02[m+1]am+1\displaystyle\frac{i\tilde{\alpha}_{0}}{2}\left(m-1\right)a_{m-1}-i\frac{\tilde{\alpha}_{0}}{2}\left[m+1\right]a_{m+1} (33)
+(ηm2μhz)bm+Δsdm,\displaystyle+\left(\eta m^{2}-\mu^{\prime}-h_{z}\right)b_{m}+\Delta_{s}d_{m},
Ecm\displaystyle Ec_{m} =\displaystyle= Δsam+[ηm2+μ+hz]cmiα~02(m1)dm1\displaystyle\Delta_{s}a_{m}+\left[-\eta m^{2}+\mu^{\prime}+h_{z}\right]c_{m}-i\frac{\tilde{\alpha}_{0}}{2}\left(m-1\right)d_{m-1} (34)
+iα~02(m+1)dm+1,\displaystyle+i\frac{\tilde{\alpha}_{0}}{2}\left(m+1\right)d_{m+1},
Edm\displaystyle Ed_{m} =\displaystyle= Δsbmiα~02(m1)cm1+iα~02(m+1)cm+1\displaystyle\Delta_{s}b_{m}-i\frac{\tilde{\alpha}_{0}}{2}\left(m-1\right)c_{m-1}+i\frac{\tilde{\alpha}_{0}}{2}\left(m+1\right)c_{m+1} (35)
+[ηm2+μhz]dm.\displaystyle+\left[-\eta m^{2}+\mu^{\prime}-h_{z}\right]d_{m}.

By solving above coupled equations (32)-(35) with the truncation bounds for mm up to 5050, the eigenenergies and corresponding eigenfunctions Φ(ϕ)\Phi\left(\phi\right) could be obtained. Fig. 5 (c) in the main text presents the eigenspectrum, indicating that there are four Majorana zero modes. The coefficients νm\nu_{m} of wavefunctions of four Majorana modes are plotted in Fig. 7.

References

  • (1) C. Nayak, S. H. Simon, A. Stern, M. Freedman, and S. Das Sarma, Non-Abelian anyons and topological quantum computation, Rev. Mod. Phys. 80, 1083 (2008).
  • (2) A. Y. Kitaev, Fault-tolerant quantum computation by anyons, Ann. Phys. 303, 2 (2003).
  • (3) S. R. Elliott and M. Franz, Majorana fermions in nuclear, particle, and solid-state physics, Rev. Mod. Phys. 87, 137 (2015).
  • (4) N. Read and D. Green, Paired states of fermions in two dimensions with breaking of parity and time-reversal symmetries and the fractional quantum Hall effect, Phys. Rev. B 61, 10267 (2000).
  • (5) V. T. Phong, N. R. Walet, and F. Guinea, Majorana zero modes in a two-dimensional pp-wave superconductor, Phys. Rev. B 96, 060505(R) (2017).
  • (6) X.-L. Qi and S.-C. Zhang, Topological insulators and superconductors, Rev. Mod. Phys. 83, 1057 (2011).
  • (7) M. Z. Hasan and C. L. Kane, Colloquium: Topological insulators, Rev. Mod. Phys. 82, 3045 (2010).  
  • (8) M. Sato, Y. Takahashi, and S. Fujimoto, Non-Abelian topological order in ss-wave superuids of ultracold fermionic atoms, Phys. Rev. Lett. 103, 020401 (2009).
  • (9) C. Zhang, S. Tewari, R. M. Lutchyn, and S. Das Sarma, px+ipyp_{x}+ip_{y} superuid from ss-wave interactions of fermionic cold atoms, Phys.Rev. Lett. 101, 160401 (2008).
  • (10) L. Jiang, T. Kitagawa, J. Alicea, A. R. Akhmerov, D. Pekker, G. Refael, J. I. Cirac, E. Demler, M. D. Lukin, and P. Zoller, Majorana fermions in equilibrium and in driven cold-atom quantum chains, Phys. Rev. Lett. 106, 220402 (2011).
  • (11) J. Alicea, Y. Oreg, G. Refael, F. von Oppen, and M. P. A. Fisher, Non-Abelian statistics and topological quantum information processing in 1D chain networks, Nat. Phys. 7, 412 (2011).
  • (12) J. Alicea, New directions in the pursuit of Majorana fermions in solid state systems, Rep. Prog. Phys. 75, 076501 (2012).
  • (13) Y. Oreg, G. Refael, and F. von Oppen, Helical Liquids and Majorana Bound States in Quantum Wires, Phys. Rev. Lett. 105, 177002 (2010).
  • (14) R. M. Lutchyn, J. D. Sau, and S. Das Sarma, Majorana Fermions and a Topological Phase Transition in Semiconductor-Superconductor Heterostructures, Phys. Rev. Lett. 105, 077001 (2010).
  • (15) L. Jiang, C. Qu, and C. Zhang, One-dimensional topological chains with Majorana fermions in two-dimensional nontopological optical lattices, Phys. Rev. A 93, 063614 (2016).
  • (16) J. D. Sau, R. M. Lutchyn, S. Tewari, and S. Das Sarma, Generic New Platform for Topological Quantum Computation Using Semiconductor Heterostructures, Phys. Rev. Lett. 104, 040502 (2010).
  • (17) S. Tewari and J. D. Sau, Topological Invariants for Spin-Orbit Coupled Superconductor Nanowires, Phys. Rev. Lett. 109, 150408 (2012).
  • (18) T. Ojanen, Topological π\pi Josephson junction in superconducting Rashba wires, Phys. Rev. B 87, 100506 (2013).
  • (19) V. Mourik, K. Zuo, S. M Frolov, S. Plissard, E. Bakkers, and L. Kouwenhoven, Signatures of majorana fermions in hybrid superconductor-semiconductor nanowire devices, Science 336, 1003 (2012).
  • (20) A. D. K. Finck, D. J. Van Harlingen, P. K. Mohseni, K. Jung, and X. Li, Anomalous modulation of a zerobias peak in a hybrid nanowire-superconductor device, Phys. Rev. Lett. 110, 126406 (2013).
  • (21) S. Nadj-Perge, I. K Drozdov, J. Li, H. Chen, S. Jeon, J. Seo, A. H. MacDonald, B. A. Bernevig, and A. Yazdani, Observation of majorana fermions in ferromagnetic atomic chains on a superconductor, Science 346, 602 (2014).
  • (22) J.-P. Xu, M.-X. Wang, Z. L. Liu, J.-F. Ge, X. Yang, C. Liu, Z. A. Xu, D. Guan, C. L. Gao, D. Qian, Y. Liu, Q.-H. Wang, F.-C. Zhang, Q.-K. Xue, and J.-F. Jia, Experimental detection of a majorana mode in the core of a magnetic vortex inside a topological insulator-superconductor Bi2Te3/NbSe2 heterostructure, Phys. Rev. Lett. 114, 017001 (2015).
  • (23) Z. F. Wang, H. Zhang, D. Liu, C. Liu, C. Tang, C. Song, Y. Zhong, J. Peng, F. Li, C. Nie, L. Wang, X. J. Zhou, X. Ma, Q. K. Xue, and F. Liu, Topological edge states in a high-temperature superconductor FeSe/SrTiO3 (001) film, Nat. Materials 15, 968 (2016).
  • (24) Q. L. He, L. Pan, A. L. Stern, E. Burks, X. Che, G. Yin, J. Wang, B. Lian, Q. Zhou, E. S. Choi, K. Murata, X. Kou, T. Nie, Q. Shao, Y. Fan, S.-C. Zhang, K. Liu, J. Xia, and K. L. Wang, Chiral Majorana edge state in a quantum anomalous Hall insulator-superconductor structure, Science 357, 294 (2017).
  • (25) Y.-J. Lin, K. Jiménez-García, and I.B. Spielman, Spin-orbit-coupled Bose-Einstein condensates, Nature 471, 83(2011).
  • (26) 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, Collective dipole oscillations of a spin-orbit coupled Bose-Einstein condensate, Phys. Rev. Lett. 109, 115301 (2012).
  • (27) P. Wang, Z.-Q. Yu, Z. Fu, J. Miao, L. Huang, S. Chai, H. Zhai, and J. Zhang, Spin-orbit coupled degenerate Fermi gases, Phys. Rev. Lett. 109, 095301 (2012).
  • (28) R. A. Williams, M. C. Beeler, L. J. LeBlanc, and I. B. Spielman, Raman-induced interactions in a single-component Fermi gas near an ss-wave Feshbach resonance, Phys. Rev. Lett. 111, 095301 (2013).
  • (29) L. Huang, Z. Meng, P. Wang, P. Peng, S.-L. Zhang, L. Chen, D. Li, Q. Zhou, and J. Zhang, Experimental realization of two-dimensional synthetic spin-orbit coupling in ultracoldFermi gases, Nat. Phys. 12, 540 (2016).
  • (30) Z. Meng, L. Huang, P. Peng, D. Li, L. Chen, Y. Xu, C. Zhang, P. Wang, and J. Zhang, Experimental observation of a topological band gap opening in ultracold Fermi gases with two-dimensional spin-orbit coupling, Phys. Rev. Lett. 117, 235304 (2016).
  • (31) Z. Wu, L. Zhang, W. Sun, X.-T. Xu, B.-Z. Wang, S.-C. Ji, Y. Deng, S. Chen, X.-J. Liu, J.-W. Pan, Realization of two-dimensional spin-orbitcoupling for Bose-Einstein condensates, Science 354, 83 (2016).
  • (32) A. Bühler, N. Lang, C. V. Kraus, G. Moller, S. D. Huber, H. P. Büchler, Nat. Commun. 5, 4504 (2014).
  • (33) D.-L. Deng, S.-T. Wang, K. Sun, and L.-M. Duan, Proposal for observing non-Abelian statistics of Majorana-Shockley fermions in an optical lattice, Phys. Rev. B 91, 094513 (2015).
  • (34) J. Langbehn, Y. Peng, L. Trifunovic, F. von Oppen, and P. W. Brouwer, Reflection-Symmetric Second-Order Topological Insulators and Superconductors, Phys. Rev. Lett. 119, 246401 (2017).
  • (35) W. A. Benalcazar, B. A. Bernevig, and T. L. Hughes, Quantized electric multipole insulators, Science 357, 61 (2017).
  • (36) Z. Song, Z. Fang, and C. Fang, (d2)\left(d-2\right)-Dimensional Edge States of Rotation Symmetry Protected Topological States, Phys. Rev. Lett. 119, 246402(2017).
  • (37) C. Fang and L. Fu, Rotation Anomaly and Topological Crystalline Insulators, arXiv:1709.01929.
  • (38) F. Schindler, A. M. Cook, M. G. Vergniory, Z. Wang, S. S. P. Parkin, B. A. Bernevig, T. Neupert, Higher-order topological insulators, Sci. Adv. 4, eaat0346 (2018).
  • (39) E. Khalaf, H. C. Po, A. Vishwanath, and H. Watanabe, Symmetry indicators and anomalous surface states of topological crystalline insulators, Phys. Rev. X 8, 031070 (2018).
  • (40) X. Zhu, Tunable Majorana corner states in a two-dimensional second-order topologicalsuperconductor induced by magnetic fields, Phys. Rev. B 97.205134 (2018).
  • (41) E. Khalaf, Higher-order topological insulators and superconductors protected by inversion symmetry, Phys. Rev. B. 97, 205136 (2018).
  • (42) Z. Yan, F. Song, and Z. Wang, Majorana Corner Modes in a High-Temperature Platform, Phys. Rev. Lett. 121, 096803 (2018).
  • (43) Q. Wang, C.-C. Liu, Y.-M. Lu, and F. Zhang, High-Temperature Majorana Corner States,  Phys. Rev. Lett. 121, 186801 (2018).
  • (44) Y. Wang, M. Lin, and T.L. Hughes, Weak-Pairing Higher Order Topological Superconductors Phys. Rev. B 98, 165144  (2018).
  • (45) T. Liu, J. J. He, and F. Nori, Majorana corner states in a two-dimensional magnetic topological insulator on a high-temperature superconductor, Phys. Rev. B 98, 245413 (2018).
  • (46) C.-H. Hsu, P. Stano, J. Klinovaja, and D. Loss, Majorana Kramers pairs in higher-order topological insulators, Phys. Rev. Lett. 121, 196801 (2018).
  • (47) Y. Xu, L. Mao, B. Wu, and C. Zhang, Dark Solitons with Majorana Fermions in Spin-Orbit-Coupled Fermi Gases, Phys. Rev. Lett. 113, 130404 (2014).
  • (48) X.-J. Liu, Soliton-induced Majorana fermions in a one-dimensional atomic topological superfluid, Phys. Rev. A 91, 023610 (2015).
  • (49) P. Würtz, T. Langen, T. Gericke, A. Koglbauer, and H. Ott, Experimental demonstration of single-site addressability in a two-dimensional optical lattice. Phys. Rev. Lett. 103, 080404 (2009).
  • (50) W. S. Bakr, J. I. Gillen, A. Peng, S. Folling, and M. Greiner, A quantum gas microscope for detecting single atoms in a Hubbard-regime optical lattice, Nature 462, 74 (2009).
  • (51) W. S. Bakr, A. Peng, M. E. Tai, R. Ma, J. Simon, J. I. Gillen, S. Foelling, L. Pollet, and M. Greiner, Probing the Superfluid–to–Mott Insulator Transition at the Single-Atom Level, Science 329, 547 (2010).
  • (52) J. F. Sherson, C. Weitenberg, M. Endres, M. Cheneau, I. Bloch, and S. Kuhr, Single-atom-resolved fluorescence imaging of an atomic Mott insulator, Nature 467, 68 (2010).
  • (53) C. Weitenberg, M. Endres, J. F. Sherson, M. Cheneau, P. Schauß, T. Fukuhara, I. Bloch, and S. Kuhr, Single-spin addressing in an atomic Mott insulator, Nature 471, 319 (2011).
  • (54) T. Köhler, K. Góral, and P. S. Julienne, Production of cold molecules via magnetically tunable Feshbach resonances, Rev. Mod. Phys. 78, 1311 (2006).
  • (55) C. Chin, R. Grimm, P. Julienne, and E. Tiesinga, Feshbach resonances in ultracold gases, Rev. Mod. Phys. 82, 1225 (2010).
  • (56) A. Altland and M. R. Zirnbauer, Nonstandard symmetry classes in mesoscopic normal-superconducting hybrid structures, Phys. Rev. B 55, 1142 (1997).
  • (57) A. P. Schnyder, S. Ryu, A. Furusaki, and A. W. W. Ludwig, Classification of topological insulators and superconductors in three spatial dimensions, Phys. Rev. B 78, 195125 (2008).
  • (58) A. Ramanathan, K. C. Wright, S. R. Muniz, M. Zelan, W. T. Hill III, C. J. Lobb, K. Helmerson, Superflow in a Toroidal Bose-Einstein Condensate: An Atom Circuit with a Tunable Weak Link, Phys. Rev. Lett. 106, 130401 (2011).