Topological superfluid phases of attractive Fermi-Hubbard model in narrow-band cold-atom optical lattices

Tudor D. Stanescu Department of Physics and Astronomy, West Virginia University, Morgantown, WV 26506, USA    Sumanta Tewari Department of Physics and Astronomy, Clemson University, Clemson, SC 29634, USA    V. W. Scarola Department of Physics, Virginia Tech, Blacksburg, Virginia 24061, USA
Abstract

We investigate the effects of attractive Hubbard interaction on two-component fermionic atoms in narrow two-dimensional (2D) energy bands that exhibit Rashba spin-orbit coupling (SOC) in the presence of an applied Zeeman field. This narrow-band 2D spin-orbit coupled attractive Fermi-Hubbard model can potentially be realized in cold atom systems in optical lattices with artificially engineered Rashba SOC and Zeeman field. Employing a self-consistent mean field approximation for the pairing potential, we uncover a complex phase diagram featuring various topological superfluid (TS) phases, dependent on the chemical potential and the Zeeman field. We focus on the pairing potential and the corresponding quasiparticle gap characterizing the TS phases, which are notably small for a wide-band model with quadratic dispersion near the ΓΓ\Gammaroman_Γ-point, as found in earlier work, and we identify the parameter regimes that maximize the gap. We find that, while generally the value of the pairing potential increases with the reduction of the fermionic bandwidth, as expected for narrow- or flat-band systems, the magnitude of the topological gap characterizing the TS phases reaches a maximum of about 1012.5%10percent12.510-12.5\%10 - 12.5 % of the interaction strength at finite values of the hopping amplitude, Rashba coupling, and Zeeman field.

I Introduction:

Topologically non-trivial superconductivity [1, 2] or superfluidity has been predicted in two-dimensional (2D) spin-half fermionic systems with Rashba spin-orbit coupling (SOC), Zeeman field, and on-site attractive interaction [3, 4]. The on-site attractive interaction aims to produce s-wave superconductivity or superfluidity in the Rashba system, which becomes topological when the applied Zeeman field exceeds a critical value. Such a spin-orbit coupled Fermi-Hubbard model with an applied Zeeman field can naturally occur in non-centrosymmetric superconductors in a magnetic field [5, 6] and can potentially be realized in cold atom systems in optical lattices [3, 4, 7, 8] with artificially engineered SOC [9, 10, 11, 12] and Zeeman field. Such a setup would leverage the significant experimental progress in realizing artificial SOC with ultracold fermions [13, 14, 15, 16, 17, 18, 19, 20, 21, 22].

A self-consistent mean field calculation for the pairing potential in the spin-orbit coupled Fermi-Hubbard model under a Zeeman field, with fermions following a quadratic dispersion, has shown that although a topological superconducting phase can emerge when the Zeeman field exceeds a critical value, the magnitude of the self-consistent pair potential in the TS phase is extremely small [23]. This makes it challenging to experimentally realize the TS phase in non-centrosymmetric superconductors where the fermion density is low enough to follow a quadratic dispersion. An alternative approach for experimentally achieving a topological superfluid state in spin-orbit coupled systems is to use two-component cold fermionic atoms in optical lattices with artificially engineered SOC and Zeeman field [4, 7, 8, 10, 11]. In this setup, an attractive Hubbard-type onsite interaction can be induced by an s𝑠sitalic_s-wave Feshbach resonance. The fermion density can be sufficiently high so that the (relevant) fermion dispersion relation corresponds to the lattice dispersion, which can be narrowed by adjusting the hopping parameter and the spin-orbit coupling, nearly reaching the flat-band limit [24, 25, 26, 27].

In flat bands, the critical temperature for Cooper pair formation is predicted to be linearly proportional to the attractive interaction between the Cooper pair constituents [28, 29]. This contrasts sharply with the Bardeen–Cooper–Schrieffer (BCS) theory of superconductivity, valid for quadratic fermion dispersion, where Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT is proportional to the exponential of the inverse of the interaction strength. This indicates that the superconducting critical temperature, and thus the pair potential, is exponentially enhanced in flat-band systems, as compared to dispersive systems within the BCS formalism. The enhancement arises from the high density of states (DOS) and the dominance of interactions over kinetic energy. The band does not need to be perfectly flat to benefit from this; any band where the Hubbard interaction strength is larger than the bandwidth will suffice. In this spirit, we explore in this paper the possibility of realizing topological superfluid states in a narrow-band spin-orbit coupled Fermi-Hubbard system with a Zeeman field, where the relevant system parameters have values smaller than (but comparable to) the interaction strength, so that the pairing potential is enhanced by the narrowness of the band making the realization of robust TS phases more feasible experimentally.

We investigate the effects of attractive Hubbard interactions on two-component fermionic atoms in narrow two-dimensional energy bands with Rashba spin-orbit coupling (SOC) and applied Zeeman field. Using self-consistent mean field theory for evaluating the pairing potential, we uncover a complex phase diagram featuring various topological superfluid phases that depend on the chemical potential and the Zeeman field. We examine whether the self-consistent pairing potential characterizing the TS phases, which is notably small for a simple quadratic dispersion near the ΓΓ\Gammaroman_Γ-point, as found in earlier work [23], increases as the bandwidth narrows in the optical lattice, as predicted for systems near the flat-band limit [28, 29]. In addition, we determine the (topological) quasiparticle gap that protects different TS phases and identify the optimal system parameters that maximize this gap, ensuring the realization of a robust topological superfluid. We find that, in general, reducing the fermionic bandwidth enhances the pairing potential, consistent with the expected behavior for narrow- or flat-band systems. However, the gap characterizing the TS phases does not exceed maximum values evaluated within our mean-field approximation at about 12.5%percent12.512.5\%12.5 % of the interaction strength. As discussed latter in the work, these maximum gap values are realized within specific parameter regimes characterized by finite values of the hopping amplitude, Rashba coupling, and Zeeman field. A topological gap 1012%10percent1210-12\%10 - 12 % of the attractive interaction strength V0subscript𝑉0V_{0}italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, which can be controlled by a Feshbach resonance, is a significant improvement over the topological gaps potentially achievable in naturally occurring systems, such as noncentrosymmetric superconductors [23]. In recent work [30], the phase diagram of this model was discussed as a function of Rashba coupling and Zeeman field but only near the half-filled limit. Thus, the TS phase found in this work is the Chern number C=2𝐶2C=2italic_C = 2 phase near chemical potential equal to 4t4𝑡4t4 italic_t (which corresponds to half-filling in our model) in Fig. 2 below. By contrast, in the present work, we discuss a variety of TS phases and discuss the phase diagram as a function of the control parameters chemical potential and Zeeman field.

The manuscript is organized as follows. In Sec. II we define the attractive Hubbard model and briefly describe our mean-field approach. In Sec. III we present numerical solutions to the mean field equations, discussing the topological phase diagram and the dependence of the pairing potential and quasiparticle gap on relevant parameters and identifying the optimal regimes that maximize the topological gaps characterizing different TS phases. We summarize in Sec. IV.

II Theoretical model

We consider a two-dimensional (2D) interacting system with Rashba-type spin-orbit coupling (SOC) and perpendicular Zeeman field and we describe it using a tight-binding model defined on a square lattice. In the basis corresponding to the eigenstates |ϕ𝒌σketsubscriptitalic-ϕ𝒌𝜎|\phi_{{\bm{k}}\sigma}\rangle| italic_ϕ start_POSTSUBSCRIPT bold_italic_k italic_σ end_POSTSUBSCRIPT ⟩ of the (non-interacting) system with no SOC and no Zeeman field, the corresponding (second quantized) Hamiltonian has the form H=H0+Hint𝐻subscript𝐻0subscript𝐻𝑖𝑛𝑡H=H_{0}+H_{int}italic_H = italic_H start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_H start_POSTSUBSCRIPT italic_i italic_n italic_t end_POSTSUBSCRIPT, with the non-interacting component having the form

H0=𝒌,σ(ξ𝒌+σΓ)c𝒌σc𝒌σ+𝒌(α𝒌c𝒌c𝒌+h.c.),H_{0}=\sum_{{\bm{k}},\sigma}(\xi_{{\bm{k}}}+\sigma\Gamma)\leavevmode\nobreak\ % \!c_{{\bm{k}}\sigma}^{\dagger}c_{{\bm{k}}\sigma}+\sum_{\bm{k}}\left(\alpha_{% \bm{k}}\leavevmode\nobreak\ \!c_{{\bm{k}}\uparrow}^{\dagger}c_{{\bm{k}}% \downarrow}+h.c.\right),italic_H start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT bold_italic_k , italic_σ end_POSTSUBSCRIPT ( italic_ξ start_POSTSUBSCRIPT bold_italic_k end_POSTSUBSCRIPT + italic_σ roman_Γ ) italic_c start_POSTSUBSCRIPT bold_italic_k italic_σ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_c start_POSTSUBSCRIPT bold_italic_k italic_σ end_POSTSUBSCRIPT + ∑ start_POSTSUBSCRIPT bold_italic_k end_POSTSUBSCRIPT ( italic_α start_POSTSUBSCRIPT bold_italic_k end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT bold_italic_k ↑ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_c start_POSTSUBSCRIPT bold_italic_k ↓ end_POSTSUBSCRIPT + italic_h . italic_c . ) , (1)

where the operator c𝒌σsuperscriptsubscript𝑐𝒌𝜎c_{{\bm{k}}\sigma}^{\dagger}italic_c start_POSTSUBSCRIPT bold_italic_k italic_σ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT creates a particle in the single-particle state |ϕ𝒌σketsubscriptitalic-ϕ𝒌𝜎|\phi_{{\bm{k}}\sigma}\rangle| italic_ϕ start_POSTSUBSCRIPT bold_italic_k italic_σ end_POSTSUBSCRIPT ⟩ with momentum 𝒌=(kx,ky)𝒌subscript𝑘𝑥subscript𝑘𝑦{\bm{k}}=(k_{x},k_{y})bold_italic_k = ( italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT , italic_k start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT ) and spin projection σ=±1\sigma=\pm 1\equiv\uparrow\downarrowitalic_σ = ± 1 ≡ ↑ ↓ along the (perpendicular) z𝑧zitalic_z-axis, ξ𝒌=2t(2coskxcosky)μsubscript𝜉𝒌2𝑡2subscript𝑘𝑥subscript𝑘𝑦𝜇\xi_{\bm{k}}=2t\leavevmode\nobreak\ \!(2-\cos k_{x}-\cos k_{y})-\muitalic_ξ start_POSTSUBSCRIPT bold_italic_k end_POSTSUBSCRIPT = 2 italic_t ( 2 - roman_cos italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT - roman_cos italic_k start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT ) - italic_μ is the energy spectrum relative to the chemical potential μ𝜇\muitalic_μ, assuming nearest-neighbor hopping with amplitude t𝑡titalic_t, ΓΓ\Gammaroman_Γ is the Zeeman field, and α𝒌=α(sinky+isinkx)subscript𝛼𝒌𝛼subscript𝑘𝑦𝑖subscript𝑘𝑥\alpha_{\bm{k}}=\alpha\leavevmode\nobreak\ \!(\sin k_{y}+i\sin k_{x})italic_α start_POSTSUBSCRIPT bold_italic_k end_POSTSUBSCRIPT = italic_α ( roman_sin italic_k start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT + italic_i roman_sin italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT ) is the SOC contribution. For convenience, we have chosen the lattice constant a𝑎aitalic_a as the unit for length, i.e., we have a=1𝑎1a=1italic_a = 1. Considering purely local attractive interactions, the second term of the Hamiltonian becomes

Hint=V0𝒌,𝒌,𝒒c𝒌+𝒒c𝒌c𝒌𝒒c𝒌subscript𝐻𝑖𝑛𝑡subscript𝑉0subscript𝒌superscript𝒌𝒒superscriptsubscript𝑐𝒌𝒒absentsubscript𝑐𝒌absentsuperscriptsubscript𝑐superscript𝒌𝒒absentsubscript𝑐superscript𝒌absentH_{int}=-V_{0}\sum_{{\bm{k}},{\bm{k}}^{\prime},{\bm{q}}}c_{{\bm{k}}+{\bm{q}}% \uparrow}^{\dagger}c_{{\bm{k}}\uparrow}\leavevmode\nobreak\ \!c_{{\bm{k}}^{% \prime}-{\bm{q}}\downarrow}^{\dagger}c_{{\bm{k}}^{\prime}\downarrow}italic_H start_POSTSUBSCRIPT italic_i italic_n italic_t end_POSTSUBSCRIPT = - italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT bold_italic_k , bold_italic_k start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , bold_italic_q end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT bold_italic_k + bold_italic_q ↑ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_c start_POSTSUBSCRIPT bold_italic_k ↑ end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT bold_italic_k start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT - bold_italic_q ↓ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_c start_POSTSUBSCRIPT bold_italic_k start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ↓ end_POSTSUBSCRIPT (2)

with V0>0subscript𝑉00V_{0}>0italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT > 0 being the interaction strength. Note that, using a real space basis, |ϕiσketsubscriptitalic-ϕ𝑖𝜎|\phi_{i\sigma}\rangle| italic_ϕ start_POSTSUBSCRIPT italic_i italic_σ end_POSTSUBSCRIPT ⟩, with i𝑖iitalic_i labeling lattice sites, the Hamiltonian in Eq. (2) takes the familiar Hubbard form, Hint=V0ininisubscript𝐻𝑖𝑛𝑡subscript𝑉0subscript𝑖subscript𝑛𝑖absentsubscript𝑛𝑖absentH_{int}=-V_{0}\sum_{i}n_{i\uparrow}n_{i\downarrow}italic_H start_POSTSUBSCRIPT italic_i italic_n italic_t end_POSTSUBSCRIPT = - italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_i ↑ end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_i ↓ end_POSTSUBSCRIPT, with niσ=ciσciσsubscript𝑛𝑖𝜎superscriptsubscript𝑐𝑖𝜎subscript𝑐𝑖𝜎n_{i\sigma}=c_{i\sigma}^{\dagger}c_{i\sigma}italic_n start_POSTSUBSCRIPT italic_i italic_σ end_POSTSUBSCRIPT = italic_c start_POSTSUBSCRIPT italic_i italic_σ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_c start_POSTSUBSCRIPT italic_i italic_σ end_POSTSUBSCRIPT.

In this study, we are interested in the narrow band regime characterized by system parameters (t𝑡titalic_t, α𝛼\alphaitalic_α, and V0subscript𝑉0V_{0}italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT) and control parameters (ΓΓ\Gammaroman_Γ and μ𝜇\muitalic_μ) having comparable values and, for convenience, we chose the interaction strength, V0subscript𝑉0V_{0}italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT as the energy unit. We focus on investigating the possible emergence of topological superfluid phases and finding the optimal regime(s) associated with these phases. To obtain a clear general understanding of the relevant phases and to efficiently explore the rather large parameter space, we use a mean-field approach. Note, however, that in the narrow band regime correlation effects could be significant, hence testing the validity of our results beyond mean-field remains an important future task.

Within a mean-field approach, the interaction term can be approximated by a sum of pairing contributions. The corresponding effective Hamiltonian is Heff=H0+HΔsubscript𝐻𝑒𝑓𝑓subscript𝐻0subscript𝐻ΔH_{eff}=H_{0}+H_{\Delta}italic_H start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT = italic_H start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_H start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT, with the pairing Hamiltonian having the form

HΔ=𝒌(Δc𝒌c𝒌+Δc𝒌c𝒌),subscript𝐻Δsubscript𝒌Δsuperscriptsubscript𝑐𝒌absentsuperscriptsubscript𝑐𝒌absentsuperscriptΔsubscript𝑐𝒌absentsubscript𝑐𝒌absentH_{\Delta}=\sum_{\bm{k}}\left(\Delta\leavevmode\nobreak\ \!c_{{\bm{k}}\uparrow% }^{\dagger}c_{{-\bm{k}}\downarrow}^{\dagger}+\Delta^{*}\leavevmode\nobreak\ \!% c_{{-\bm{k}}\downarrow}c_{{\bm{k}}\uparrow}\right),italic_H start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT bold_italic_k end_POSTSUBSCRIPT ( roman_Δ italic_c start_POSTSUBSCRIPT bold_italic_k ↑ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_c start_POSTSUBSCRIPT - bold_italic_k ↓ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT + roman_Δ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT italic_c start_POSTSUBSCRIPT - bold_italic_k ↓ end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT bold_italic_k ↑ end_POSTSUBSCRIPT ) , (3)

with the pairing potential ΔΔ\Deltaroman_Δ being determined self-consistently by solving a mean-field self-consistent gap equation. A derivation of the gap equation can be found, for example, in Ref. 23. At zero temperature, the pairing ΔΔ\Deltaroman_Δ is a solution of the gap equation θ(Δ)=0𝜃Δ0\theta(\Delta)=0italic_θ ( roman_Δ ) = 0, with

θ(Δ)𝜃Δ\displaystyle\theta(\Delta)italic_θ ( roman_Δ ) =\displaystyle== 1V04Nk𝒌[1E1+1E2\displaystyle-1-\frac{V_{0}}{4N_{k}}\sum_{\bm{k}}\left[\frac{1}{E_{1}}+\frac{1% }{E_{2}}\right.- 1 - divide start_ARG italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG 4 italic_N start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUBSCRIPT bold_italic_k end_POSTSUBSCRIPT [ divide start_ARG 1 end_ARG start_ARG italic_E start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG + divide start_ARG 1 end_ARG start_ARG italic_E start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG
\displaystyle-- Γ2ξ𝒌2|α𝒌|2+Γ2(ξ𝒌2+|Δ|2)(1E11E2)],\displaystyle\left.\frac{\Gamma^{2}}{\sqrt{\xi_{\bm{k}}^{2}|\alpha_{\bm{k}}|^{% 2}+\Gamma^{2}(\xi_{\bm{k}}^{2}+|\Delta|^{2})}}\left(\frac{1}{E_{1}}-\frac{1}{E% _{2}}\right)\right],divide start_ARG roman_Γ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG square-root start_ARG italic_ξ start_POSTSUBSCRIPT bold_italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT | italic_α start_POSTSUBSCRIPT bold_italic_k end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + roman_Γ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_ξ start_POSTSUBSCRIPT bold_italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + | roman_Δ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_ARG end_ARG ( divide start_ARG 1 end_ARG start_ARG italic_E start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG - divide start_ARG 1 end_ARG start_ARG italic_E start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG ) ] ,

where the summation is done over the 2D Brillouin zone containing Nksubscript𝑁𝑘N_{k}italic_N start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT momentum values. The energies E10subscript𝐸10E_{1}\geq 0italic_E start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≥ 0 and E2>0subscript𝐸20E_{2}>0italic_E start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT > 0, which depend on the system parameters and the pairing ΔΔ\Deltaroman_Δ, are the (positive) eigenvalues of the effective Hamiltonian, Heffsubscript𝐻𝑒𝑓𝑓H_{eff}italic_H start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT. Explicitly, we have

E1(2)2=ξ𝒌2+|α𝒌|2+Γ2+|Δ|22ξ𝒌2|α𝒌|2+Γ2(ξ𝒌2+|Δ|2).superscriptsubscript𝐸122minus-or-plussuperscriptsubscript𝜉𝒌2superscriptsubscript𝛼𝒌2superscriptΓ2superscriptΔ22superscriptsubscript𝜉𝒌2superscriptsubscript𝛼𝒌2superscriptΓ2superscriptsubscript𝜉𝒌2superscriptΔ2E_{1(2)}^{2}=\xi_{\bm{k}}^{2}+|\alpha_{\bm{k}}|^{2}+\Gamma^{2}+|\Delta|^{2}\mp 2% \sqrt{\xi_{\bm{k}}^{2}|\alpha_{\bm{k}}|^{2}\!+\!\Gamma^{2}(\xi_{\bm{k}}^{2}\!+% \!|\Delta|^{2})}.italic_E start_POSTSUBSCRIPT 1 ( 2 ) end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = italic_ξ start_POSTSUBSCRIPT bold_italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + | italic_α start_POSTSUBSCRIPT bold_italic_k end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + roman_Γ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + | roman_Δ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ∓ 2 square-root start_ARG italic_ξ start_POSTSUBSCRIPT bold_italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT | italic_α start_POSTSUBSCRIPT bold_italic_k end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + roman_Γ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_ξ start_POSTSUBSCRIPT bold_italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + | roman_Δ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_ARG . (5)

Note that the solutions of the pairing equation θ(Δ)=0𝜃Δ0\theta(\Delta)=0italic_θ ( roman_Δ ) = 0 corresponding to dθ/dΔ>0𝑑𝜃𝑑Δ0d\theta/d\Delta>0italic_d italic_θ / italic_d roman_Δ > 0 are unphysical [23]. For a given set of system and control parameters, (t𝑡titalic_t, α𝛼\alphaitalic_α, V0subscript𝑉0V_{0}italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, ΓΓ\Gammaroman_Γ, and μ𝜇\muitalic_μ), we determine the pairing ΔΔ\Deltaroman_Δ by solving numerically the equation θ(Δ)=0𝜃Δ0\theta(\Delta)=0italic_θ ( roman_Δ ) = 0 and we characterize the low-energy physics of the system (at the mean-field level) by solving the eigenvalue problem associated with the effective Hamiltonian Heffsubscript𝐻𝑒𝑓𝑓H_{eff}italic_H start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT. In particular, we calculate the quasiparticle gap Δqp(𝒌)ΔsubscriptΔ𝑞𝑝𝒌Δ\Delta_{qp}({\bm{k}})\leq\Deltaroman_Δ start_POSTSUBSCRIPT italic_q italic_p end_POSTSUBSCRIPT ( bold_italic_k ) ≤ roman_Δ characterising the system in different regimes. Note that ΔqpsubscriptΔ𝑞𝑝\Delta_{qp}roman_Δ start_POSTSUBSCRIPT italic_q italic_p end_POSTSUBSCRIPT vanishes at a topological quantum phase transition (TQPT), although the pairing ΔΔ\Deltaroman_Δ is finite, or in the absence of pairing, Δ=0Δ0\Delta=0roman_Δ = 0.

Refer to caption
Figure 1: Energy spectra (left panels) and the corresponding DOS (right panels) for a non-interacting (Δ=0Δ0\Delta=0roman_Δ = 0) narrow-band system with two sets of parameters: (a) t=0.3V0𝑡0.3subscript𝑉0t=0.3V_{0}italic_t = 0.3 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, α=0.4V0𝛼0.4subscript𝑉0\alpha=0.4V_{0}italic_α = 0.4 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, and Γ=0.23V0Γ0.23subscript𝑉0\Gamma=0.23V_{0}roman_Γ = 0.23 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT; (b) t=0.1V0𝑡0.1subscript𝑉0t=0.1V_{0}italic_t = 0.1 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, α=0.4V0𝛼0.4subscript𝑉0\alpha=0.4V_{0}italic_α = 0.4 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, and Γ=0.3V0Γ0.3subscript𝑉0\Gamma=0.3V_{0}roman_Γ = 0.3 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. Note that these parameters correspond to maximum values of the topological gap for the regimes discussed in Sec. III.2. The DOS is given in units of 1/2πV012𝜋subscript𝑉01/2\pi V_{0}1 / 2 italic_π italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, which is the DOS at the bottom of a (spin-degenerate) band for a square lattice model with nearest-neighbor hopping t=V0𝑡subscript𝑉0t=V_{0}italic_t = italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT.

The main objective of this study is to determine to what degree narrowing the bandwidth, which enhances the density of states (DOS) and, therefore, is expected to strengthen the pairing potential, results in an enhancement of the topological gap. Two examples of narrow-band non-interacting energy spectra and the corresponding DOS are shown in Fig. 1. Note that the density of states is measured in units corresponding to the DOS at the bottom of the band for a square lattice model with nearest-neighbor hopping t=V0𝑡subscript𝑉0t=V_{0}italic_t = italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, where V0subscript𝑉0V_{0}italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT can be viewed as an unspecified energy scale to be determined (in an interacting system) by the interaction strength. In the wide band limit (tV0much-greater-than𝑡subscript𝑉0t\gg V_{0}italic_t ≫ italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT), e.g., for a wide-band model with quadratic dispersion near the ΓΓ\Gammaroman_Γ-point, the corresponding DOS is 1/2πt1/2πV0much-less-than12𝜋𝑡12𝜋subscript𝑉01/2\pi t\ll 1/2\pi V_{0}1 / 2 italic_π italic_t ≪ 1 / 2 italic_π italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. The DOS shown in Fig. 1 has values that are orders of magnitude higher that those corresponding to the large bandwidth regime. Below we investigate the impact of this large DOS on the stability of topological superfluid phases that emerge in the presence of onsite attractive interaction.

III Results

In this section we present the results of our numerical analysis, starting with a general discussion of the topological phase diagram for an effective Hamiltonian Heffsubscript𝐻𝑒𝑓𝑓H_{eff}italic_H start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT with constant pairing (Sec. III.1). We continue in Sec. III.2 with self-consistent calculations of the topological phase diagram corresponding to two different parameter regimes. Finally, in Sec. III.3 we identify the optimal parameter regimes that maximize the topological gap, i.e., the minimum over the Brillouin zone of Δqp(𝒌)subscriptΔ𝑞𝑝𝒌\Delta_{qp}({\bm{k}})roman_Δ start_POSTSUBSCRIPT italic_q italic_p end_POSTSUBSCRIPT ( bold_italic_k ).

III.1 Non-interacting topological phase diagram

To clearly identify the possible topological phases hosted by the system and to understand the basic structure of the topological phase diagram, we first consider the effective Hamiltonian Heff=H0+HΔsubscript𝐻𝑒𝑓𝑓subscript𝐻0subscript𝐻ΔH_{eff}=H_{0}+H_{\Delta}italic_H start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT = italic_H start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_H start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT, with H0subscript𝐻0H_{0}italic_H start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT given by Eq. (1) and a paring term HΔsubscript𝐻ΔH_{\Delta}italic_H start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT given by Eq. (3) having a constant (i.e., parameter-independent) pairing potential ΔΔ\Deltaroman_Δ. Note that Heffsubscript𝐻𝑒𝑓𝑓H_{eff}italic_H start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT, which has particle-hole symmetry, but no time-reversal symmetry, belongs to symmetry class D, hence it supports topological phases with a \mathbb{Z}blackboard_Z classification in two dimensions [31]. These topological phases are characterized by a Chern number invariant that can be expressed in terms of the Green’s function [32, 33] G(𝒌,iω)=(iωHeff)1𝐺𝒌𝑖𝜔superscript𝑖𝜔subscript𝐻𝑒𝑓𝑓1G({\bm{k}},i\omega)=(i\omega-H_{eff})^{-1}italic_G ( bold_italic_k , italic_i italic_ω ) = ( italic_i italic_ω - italic_H start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT as

C=12d2k(2π)2𝐶12superscript𝑑2𝑘superscript2𝜋2\displaystyle C=\frac{1}{2}\int\!\!\frac{d^{2}k}{(2\pi)^{2}}italic_C = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ∫ divide start_ARG italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_k end_ARG start_ARG ( 2 italic_π ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG dωTr[GkxG1GkyG1GωG1\displaystyle\int\!\!d\omega\leavevmode\nobreak\ {\rm Tr}\left[G\partial_{k_{x% }}G^{-1}\leavevmode\nobreak\ \!G\partial_{k_{y}}G^{-1}\leavevmode\nobreak\ \!G% \partial_{\omega}G^{-1}\right.∫ italic_d italic_ω roman_Tr [ italic_G ∂ start_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_G start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT italic_G ∂ start_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_G start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT italic_G ∂ start_POSTSUBSCRIPT italic_ω end_POSTSUBSCRIPT italic_G start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT (6)
GkyG1GkxG1GωG1].\displaystyle-\left.G\partial_{k_{y}}G^{-1}\leavevmode\nobreak\ \!G\partial_{k% _{x}}G^{-1}\leavevmode\nobreak\ \!G\partial_{\omega}G^{-1}\right].- italic_G ∂ start_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_G start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT italic_G ∂ start_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_G start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT italic_G ∂ start_POSTSUBSCRIPT italic_ω end_POSTSUBSCRIPT italic_G start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ] .

We note that the topological phase boundaries separating phases that correspond to different C𝐶Citalic_C values are associated with a vanishing energy gap at high-symmetry points in the Brillouin zone, 𝑲=(0,0),(0,π),(π,0),(π,π)𝑲000𝜋𝜋0𝜋𝜋{\bm{K}}=(0,0),\leavevmode\nobreak\ (0,\pi),\leavevmode\nobreak\ (\pi,0),% \leavevmode\nobreak\ (\pi,\pi)bold_italic_K = ( 0 , 0 ) , ( 0 , italic_π ) , ( italic_π , 0 ) , ( italic_π , italic_π ). For the model considered here, this condition can be expressed analytically using the energy eigenvalues E1,E2subscript𝐸1subscript𝐸2E_{1},E_{2}italic_E start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_E start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT in Eq. (5) , with the corresponding phase boundary equations having the form

Γ2=(μϵi)2+Δ2,superscriptΓ2superscript𝜇subscriptitalic-ϵ𝑖2superscriptΔ2\displaystyle\Gamma^{2}=(\mu-\epsilon_{i})^{2}+\Delta^{2},roman_Γ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = ( italic_μ - italic_ϵ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + roman_Δ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (7)

where ϵ1=0subscriptitalic-ϵ10\epsilon_{1}=0italic_ϵ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 0, ϵ2=4tsubscriptitalic-ϵ24𝑡\epsilon_{2}=4titalic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 4 italic_t, and ϵ3=8tsubscriptitalic-ϵ38𝑡\epsilon_{3}=8titalic_ϵ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = 8 italic_t. Note that the topological phase boundaries do not have an explicit dependence on the SOC strength, α𝛼\alphaitalic_α, but may implicitly depend on this parameter through the pairing potential ΔΔ\Deltaroman_Δ (when calculated self-consistently; see Sec. III.2 below).

Refer to caption
Figure 2: Topological phase diagram for a system described by the effective Hamiltonian Heff=H)+HΔsubscript𝐻𝑒𝑓𝑓subscript𝐻)subscript𝐻ΔH_{eff}=H_{)}+H_{\Delta}italic_H start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT = italic_H start_POSTSUBSCRIPT ) end_POSTSUBSCRIPT + italic_H start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT with a constant pairing potential, ΔΔ\Deltaroman_Δ. The white areas are topologically-trivial, while the colored regions represent three distinct topological superfluid phases. The phase boundaries (black lines) are given by Eq. (7), while the values of the Chern topological invariant C𝐶Citalic_C are calculated using Eq. (6). The energy spectra for a finite-width ribbon with control parameter values corresponding to the points a,b,,e𝑎𝑏𝑒a,b,\dots,eitalic_a , italic_b , … , italic_e are shown in Fig. 3.

The topological phase diagram in the Zeeman field–chemical potential plane for a two-dimensional system described by Heffsubscript𝐻𝑒𝑓𝑓H_{eff}italic_H start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT with constant pairing potential ΔΔ\Deltaroman_Δ is shown in Fig. 2. Calculating the topological invariant given by Eq. 6 reveals the presence of three topological superfluid phases corresponding to C=1𝐶1C=-1italic_C = - 1 (yellow shading in Fig. 2), C=+1𝐶1C=+1italic_C = + 1 (green), and C=+2𝐶2C=+2italic_C = + 2 (cyan). The minimum Zeeman field associated with the emergence of these phases, Γmin=|Δ|subscriptΓ𝑚𝑖𝑛Δ\Gamma_{min}=|\Delta|roman_Γ start_POSTSUBSCRIPT italic_m italic_i italic_n end_POSTSUBSCRIPT = | roman_Δ |, is controlled by the pairing potential, while the chemical potential range associated with the topological phases (for a given Zeeman field ΓΓ\Gammaroman_Γ) is controlled by the hopping t𝑡titalic_t. Note that upon reducing the hopping, t0𝑡0t\rightarrow 0italic_t → 0, the topological regions shrink and eventually collapse.

Refer to caption
Figure 3: Energy spectrum as a function of momentum for α=4t𝛼4𝑡\alpha=4titalic_α = 4 italic_t for an infinitely long ribbon with system parameters corresponding to the points marked a,b,,e𝑎𝑏𝑒a,b,\dots,eitalic_a , italic_b , … , italic_e in Fig. 2. The dark blue shading indicates bulk states, while the in-gap blue and orange lines correspond to chiral edge modes propagating on the opposite boundaries of the ribbon. Note the opposite chirality of the edge modes supported by the C=+1𝐶1C=+1italic_C = + 1 (green) and C=1𝐶1C=-1italic_C = - 1 (yellow) phases and the presence of two chiral modes on each edge for C=+2𝐶2C=+2italic_C = + 2 (cyan phase).

To shed light on the distinction between phases characterized by different (nonzero) values of the Chern invariant, we consider an infinitely-long, finite width ribbon with control parameter values corresponding to the points marked a,b,,e𝑎𝑏𝑒a,b,\dots,eitalic_a , italic_b , … , italic_e in Fig. 2 and we calculate the energy spectrum as a function of momentum along the ribbon, kxsubscript𝑘𝑥k_{x}italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT. The results are shown in Fig. 3. The topological phases are characterized by a finite bulk gap and by the presence of gapless chiral Majorana edge modes. The number and the chirality of these modes are correlated with the value of the Chern invariant characterizing each topological phase, illustrating the bulk-boundary correspondence. In addition, we note the symmetry characterizing the energy spectra in Eq. (5), which are invariant under the transformation μ8tμ𝜇8𝑡𝜇\mu\rightarrow 8t-\muitalic_μ → 8 italic_t - italic_μ, kxπkxsubscript𝑘𝑥𝜋subscript𝑘𝑥k_{x}\rightarrow\pi-k_{x}italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT → italic_π - italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT. For the two-dimensional system, the corresponding symmetry transformation, μ8tμ𝜇8𝑡𝜇\mu\rightarrow 8t-\muitalic_μ → 8 italic_t - italic_μ, 𝒌(π,π)𝒌𝒌𝜋𝜋𝒌{\bm{k}}\rightarrow(\pi,\pi)-{\bm{k}}bold_italic_k → ( italic_π , italic_π ) - bold_italic_k, results in ξ𝒌ξ𝒌subscript𝜉𝒌subscript𝜉𝒌\xi_{\bm{k}}\rightarrow-\xi_{\bm{k}}italic_ξ start_POSTSUBSCRIPT bold_italic_k end_POSTSUBSCRIPT → - italic_ξ start_POSTSUBSCRIPT bold_italic_k end_POSTSUBSCRIPT and α𝒌α𝒌subscript𝛼𝒌subscript𝛼𝒌\alpha_{\bm{k}}\rightarrow\alpha_{\bm{k}}italic_α start_POSTSUBSCRIPT bold_italic_k end_POSTSUBSCRIPT → italic_α start_POSTSUBSCRIPT bold_italic_k end_POSTSUBSCRIPT, so that the energies in Eq. (5) and the function θ(Δ)𝜃Δ\theta(\Delta)italic_θ ( roman_Δ ) given by Eq. (II) remain invariant. Below, we exploit this property and focus on the regime μ4t𝜇4𝑡\mu\leq 4titalic_μ ≤ 4 italic_t.

Refer to caption
Figure 4: Mean-field topological phase diagram for a narrow-band system with t=0.3V0𝑡0.3subscript𝑉0t=0.3V_{0}italic_t = 0.3 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and α=0.4V0𝛼0.4subscript𝑉0\alpha=0.4V_{0}italic_α = 0.4 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. The light gray area corresponds to an effectively unpaired (Δ=0Δ0\Delta=0roman_Δ = 0) phase, while the white, yellow, and cyan regions indicate the same type of gapped phases (trivial or topological) as in Fig. 2. Note that the green phase (C=+1𝐶1C=+1italic_C = + 1), as seen in Fig. (2) is not accessible because the hopping parameter t𝑡titalic_t is relatively large.

III.2 Mean-field topological phase diagram

Refer to caption
Figure 5: Mean-field topological phase diagram for a narrow-band system with t=0.1V0𝑡0.1subscript𝑉0t=0.1V_{0}italic_t = 0.1 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and α=0.4V0𝛼0.4subscript𝑉0\alpha=0.4V_{0}italic_α = 0.4 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. The color code is the same as in Fig. 2 and Fig. 4.

We now consider an interacting system and calculate the topological phase diagram (at the mean-field level) by explicitly determining the pairing potential ΔΔ\Deltaroman_Δ as the solution of the equation θ(Δ)=0𝜃Δ0\theta(\Delta)=0italic_θ ( roman_Δ ) = 0 for each set of control parameters, (Γ,μ)Γ𝜇(\Gamma,\mu)( roman_Γ , italic_μ ) for the system parameter values (t,α)=(0.3,0.4)V0𝑡𝛼0.30.4subscript𝑉0(t,\alpha)=(0.3,0.4)V_{0}( italic_t , italic_α ) = ( 0.3 , 0.4 ) italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and (t,α)=(0.1,0.4)V0𝑡𝛼0.10.4subscript𝑉0(t,\alpha)=(0.1,0.4)V_{0}( italic_t , italic_α ) = ( 0.1 , 0.4 ) italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. The corresponding phase diagrams are shown in Fig. 4 and Fig. 5, respectively. First, we point out that for certain values of the control parameters, the pairing ΔΔ\Deltaroman_Δ vanishes. This is in contrast with the wide-band regime with a quadratic dispersion ([23]), where the equation θ(Δ)=0𝜃Δ0\theta(\Delta)=0italic_θ ( roman_Δ ) = 0 has a solution for a nonzero ΔΔ\Deltaroman_Δ for all relevant control parameter values and arbitrarily weak interaction strength. This is the result of E1subscript𝐸1E_{1}italic_E start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT generating a divergent contribution to the right-hand side of Eq. (II) in the limit |Δ|0Δ0|\Delta|\rightarrow 0| roman_Δ | → 0 (from k𝑘kitalic_k points on the Fermi line [Bardee1957, 23]). However, in a narrow-band system the condition E1(𝒌)|Δ=0=0evaluated-atsubscript𝐸1𝒌Δ00E_{1}({\bm{k}})|_{\Delta=0}=0italic_E start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( bold_italic_k ) | start_POSTSUBSCRIPT roman_Δ = 0 end_POSTSUBSCRIPT = 0, which implies |ξ𝒌|=|α𝒌|2+Γ2subscript𝜉𝒌superscriptsubscript𝛼𝒌2superscriptΓ2|\xi_{\bm{k}}|=\sqrt{|\alpha_{\bm{k}}|^{2}+\Gamma^{2}}| italic_ξ start_POSTSUBSCRIPT bold_italic_k end_POSTSUBSCRIPT | = square-root start_ARG | italic_α start_POSTSUBSCRIPT bold_italic_k end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + roman_Γ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG, cannot be satisfied when the chemical potential is below (or above) the narrow band, or when it lies in a gap between the spin subbands (for large-enough values of ΓΓ\Gammaroman_Γ). In these parameter regimes there is no Fermi line and the equation θ(Δ)=0𝜃Δ0\theta(\Delta)=0italic_θ ( roman_Δ ) = 0 does not have a finite solution for arbitrarily weak (attractive) interaction. In our numerical calculations, we consider the system as being effectively unpaired (i.e., either a gapless phase or a trivial insulator) if Δ<104V0Δsuperscript104subscript𝑉0\Delta<10^{-4}V_{0}roman_Δ < 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT.

Refer to caption
Figure 6: Zeeman field dependence of the pairing potential (blue lines) and quasiparticle gap (red lines) for a system with parameters as in Fig. 4 and different chemical potential values. The vanishing of the quasiparticle gap signal a TQPT. The maximum topological gap (0.1V0absent0.1subscript𝑉0\approx 0.1V_{0}≈ 0.1 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT) corresponds to half filling (μ=4t𝜇4𝑡\mu=4titalic_μ = 4 italic_t, cyan phase with C=+2𝐶2C=+2italic_C = + 2).

Next, we focus on the topological superfluid phases that emerge at the mean-field level. The phase diagrams in Fig. 4 and Fig. 5 show that these phases can occur within significant control parameter regions, characterized by energy scales of order V0subscript𝑉0V_{0}italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. This is in sharp contrast with a wide-band system with quadratic dispersion [23], where satisfying the topological condition was found to be quite challenging and the TS phase occurred in a small region of the parameter space. This is one of the central results of the present work: in the narrow-band regime with a lattice dispersion, the topological superfluid phase is much more easily accessible in the space of the control parameters than in the wide-band case with a quadratic dispersion. However, the extent of a phase in the parameter space, i.e., the area in the phase diagram, provides direct information only about its accessibility for different values of the control parameters, but not about its robustness. The property that is most relevant to the robustness or stability of the topological phase is the size of the topological gap which depends on the pair potential ΔΔ\Deltaroman_Δ. It has been suggested that the magnitude of ΔΔ\Deltaroman_Δ increases with decreasing bandwidth, eventually becoming linearly proportional to the interaction strength V0subscript𝑉0V_{0}italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT in the flat-band limit [28, 29]. This is in contrast to the BCS theory, which applies to the wide-band regime with a quadratic dispersion. According to the BCS theory the pair potential ΔΔ\Deltaroman_Δ depends on the interaction strength V0subscript𝑉0V_{0}italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT as Δe1/g0V0proportional-toΔsuperscript𝑒1subscript𝑔0subscript𝑉0\Delta\propto e^{-1/g_{0}V_{0}}roman_Δ ∝ italic_e start_POSTSUPERSCRIPT - 1 / italic_g start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT, where g0subscript𝑔0g_{0}italic_g start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the density of states at the Fermi energy. Therefore, one can conjecture that going to the narrow- or flat-band regime the magnitude of ΔΔ\Deltaroman_Δ should increase exponentially and the topological phase should become more stable. To address this problem, we calculate the quasiparticle gap along five representative μ=const.𝜇const\mu={\rm const.}italic_μ = roman_const . cuts through the phase diagrams. Note that, by symmetry, the energy gaps at chemical potential values μ𝜇\muitalic_μ and 8tμ8𝑡𝜇8t-\mu8 italic_t - italic_μ are identical.

Refer to caption
Figure 7: Zeeman field dependence of the pairing potential (blue lines) and quasiparticle gap (red lines) for a system with parameters as in Fig. 5 and different chemical potential values. The maximum topological gap values (0.1V0absent0.1subscript𝑉0\approx 0.1V_{0}≈ 0.1 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT) correspond to half filling (μ=4t𝜇4𝑡\mu=4titalic_μ = 4 italic_t, cyan phase with C=+2𝐶2C=+2italic_C = + 2) and to the yellow (C=1𝐶1C=-1italic_C = - 1) phase (μ=0𝜇0\mu=0italic_μ = 0 and μ=2t𝜇2𝑡\mu=-2titalic_μ = - 2 italic_t). Note the non-monotonic dependence of ΔΔ\Deltaroman_Δ on the Zeeman field for μ=4t𝜇4𝑡\mu=-4titalic_μ = - 4 italic_t.

The results corresponding to the parameter regimes illustrated in Fig. 4 and Fig. 5 are presented in Fig. 6 and Fig. 7, respectively. The blue lines represent the dependence of the pairing potential Δ(Γ,μ)ΔΓ𝜇\Delta(\Gamma,\mu)roman_Δ ( roman_Γ , italic_μ ) on the applied Zeeman field, while the red lines represent the minimum (over the Brillouin zone) of the quasiparticle gap Δqp(𝒌)=E1(𝒌)subscriptΔ𝑞𝑝𝒌subscript𝐸1𝒌\Delta_{qp}({\bm{k}})=E_{1}({\bm{k}})roman_Δ start_POSTSUBSCRIPT italic_q italic_p end_POSTSUBSCRIPT ( bold_italic_k ) = italic_E start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( bold_italic_k ), where E1subscript𝐸1E_{1}italic_E start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT is given by Eq. (5) with Δ=Δ(Γ,μ)ΔΔΓ𝜇\Delta=\Delta(\Gamma,\mu)roman_Δ = roman_Δ ( roman_Γ , italic_μ ). We note that the phase boundary crossings are signaled by the vanishing of the quasiparticle gap. Increasing the Zeeman field typically results in the reduction and eventual collapse of the pairing potential, with the non-monotonic behavior in the lower panel of Fig.7 (μ=4t𝜇4𝑡\mu=-4titalic_μ = - 4 italic_t) illustrating the (less likely) scenario associated with small filling values (or, from symmetry, nearly filled systems). The key property that determines the stability of the topological phase is the size of the topological gap, i.e., quasiparticle gap in the topological regime. In the examples shown in Fig. 6 and Fig. 7, the maximum topological gap values are of the order 0.1V00.1subscript𝑉00.1V_{0}0.1 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and are realized in the C=+2𝐶2C=+2italic_C = + 2 (cyan) phase for μ=4t𝜇4𝑡\mu=4titalic_μ = 4 italic_t (i.e., half filling) — see top panels of Fig. 6 and Fig. 7 — or in the C=1𝐶1C=-1italic_C = - 1 (yellow) phase — panels μ=0𝜇0\mu=0italic_μ = 0 and μ=2t𝜇2𝑡\mu=-2titalic_μ = - 2 italic_t in Fig. 7. By contrast, the values of the topological gap characterizing the C=+1𝐶1C=+1italic_C = + 1 (green) phase from Fig. 5 are significantly smaller, while this phase is completely inaccessible in the regime corresponding to Fig. 4. Finally, we note that the maximum values of the topological gap are much smaller than the maximum values of the pairing potential, which are realized at Γ=0Γ0\Gamma=0roman_Γ = 0, i.e., in the topologically trivial phase. In Fig. 7, for example, Δ(Γ=0)ΔΓ0\Delta(\Gamma=0)roman_Δ ( roman_Γ = 0 ) can exceed 0.3V00.3subscript𝑉00.3V_{0}0.3 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, which is about three times the maximum value of the topological gap.

Refer to caption
Figure 8: Pairing potential (left panels) and topological gap (right panels) as functions of hopping (t𝑡titalic_t) and Rashba coupling (α𝛼\alphaitalic_α) for a half-filled system (μ=4t𝜇4𝑡\mu=4titalic_μ = 4 italic_t, cyan phase). In the right panels the contours are spaced by 0.0125V00.0125subscript𝑉00.0125\leavevmode\nobreak\ \!V_{0}0.0125 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (top and middle) and 0.01V00.01subscript𝑉00.01\leavevmode\nobreak\ \!V_{0}0.01 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (bottom), while in the right panels the spacing between contours is 0.01V00.01subscript𝑉00.01\leavevmode\nobreak\ \!V_{0}0.01 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. The green region is topologically trivial. The maximum value of the topological gap is about 0.115V00.115subscript𝑉00.115\leavevmode\nobreak\ \!V_{0}0.115 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (for Γ=0.25V0Γ0.25subscript𝑉0\Gamma=0.25V_{0}roman_Γ = 0.25 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT).

III.3 Optimization of the system parameters

Our analysis has shown that a narrow band system described by an attractive Hubbard-type model with spin-orbit coupling and Zeeman field can host topological superfluid phases within significant ranges of control parameters, with maximum values of the topological gap of about 10%percent1010\%10 % of the interaction strength. We can now address the main objective of this work: identifying the optimal parameter regimes that maximize the topological gap and, implicitly, the robustness of the corresponding topological phase.

Refer to caption
Figure 9: Pairing potential (left panels) and topological gap (right panels) as functions of hopping (t𝑡titalic_t) and Rashba coupling (α𝛼\alphaitalic_α) for a system with μ=0𝜇0\mu=0italic_μ = 0 (yellow phase). The spacing between contours is 0.0125V00.0125subscript𝑉00.0125\leavevmode\nobreak\ \!V_{0}0.0125 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (left panels) and 0.01V00.01subscript𝑉00.01\leavevmode\nobreak\ \!V_{0}0.01 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (right panels). The maximum topological gap is about 0.125V00.125subscript𝑉00.125\leavevmode\nobreak\ \!V_{0}0.125 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (for Γ=0.3V0Γ0.3subscript𝑉0\Gamma=0.3\leavevmode\nobreak\ \!V_{0}roman_Γ = 0.3 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT).

We emphasize that maximizing the topological gap is not the same as maximizing the pairing potential. As shown by results in Figs. 6 and 7, the largest values of ΔΔ\Deltaroman_Δ are obtained at low Zeeman fields. However, to access a topological phase ΓΓ\Gammaroman_Γ has to exceed the boundary value given by Eq. (7), in particular one has to satisfy the condition Γ>ΔΓΔ\Gamma>\Deltaroman_Γ > roman_Δ. Hence, the optimal regime cannot be realized at low Zeeman fields. On the other hand, increasing ΓΓ\Gammaroman_Γ eventually results in the collapse of the pairing potential. We conclude that, for a given set of system parameters, the maximum topological gap corresponds to a finite Zeeman field that remains to be determined.

Refer to caption
Figure 10: Same as in Fig. 9 for a system with chemical potential μ=t𝜇𝑡\mu=-titalic_μ = - italic_t (yellow phase). Topological gap values above 0.011V00.011subscript𝑉00.011V_{0}0.011 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT are obtained for Zeeman fields in the range 0.30.35V00.30.35subscript𝑉00.3-0.35\leavevmode\nobreak\ \!V_{0}0.3 - 0.35 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. Note that the pairing potential in the trivial region corresponding to Γ=0.25V0Γ0.25subscript𝑉0\Gamma=0.25\leavevmode\nobreak\ \!V_{0}roman_Γ = 0.25 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is up to 0.4V00.4subscript𝑉00.4\leavevmode\nobreak\ \!V_{0}0.4 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (top left panel).

Based on our discussion in the previous section, the largest topological gaps are likely to be realized in the C=+2𝐶2C=+2italic_C = + 2 (cyan) or C=1𝐶1C=-1italic_C = - 1 (yellow) topological phases. For the cyan phase one can easily determine that the gap is maximized at half filling, i.e., for μ=4t𝜇4𝑡\mu=4titalic_μ = 4 italic_t. For the yellow phase we expect the optimum regime to be away from the phase boundaries, in the range 3tμtless-than-or-similar-to3𝑡𝜇less-than-or-similar-to𝑡-3t\lesssim\mu\lesssim t- 3 italic_t ≲ italic_μ ≲ italic_t or 7tμ11tless-than-or-similar-to7𝑡𝜇less-than-or-similar-to11𝑡7t\lesssim\mu\lesssim 11t7 italic_t ≲ italic_μ ≲ 11 italic_t. We focus on two specific cases, μ=0𝜇0\mu=0italic_μ = 0 and μ=t𝜇𝑡\mu=-titalic_μ = - italic_t. By symmetry, the same physics will characterize a system with chemical potential values μ=8t𝜇8𝑡\mu=8titalic_μ = 8 italic_t and μ=9t𝜇9𝑡\mu=9titalic_μ = 9 italic_t, respectively. Our strategy for identifying the optimal regime is to fix the chemical potential to one of the values mentioned above, consider several Zeeman field strengths, and scan the tα𝑡𝛼t-\alphaitalic_t - italic_α parameter space. For each set of parameters we first calculate the pairing potential by solving the equation θ(Δ)=0𝜃Δ0\theta(\Delta)=0italic_θ ( roman_Δ ) = 0, then we determine the topological gap by minimizing Δqp(𝒌)subscriptΔ𝑞𝑝𝒌\Delta_{qp}({\bm{k}})roman_Δ start_POSTSUBSCRIPT italic_q italic_p end_POSTSUBSCRIPT ( bold_italic_k ) over the Brillouin zone.

The results of our numerical analysis are shown in Figs. 810. For the half-filled cyan phase (see Fig. 8), the topological gap is maximized around Γ=0.25V0Γ0.25subscript𝑉0\Gamma=0.25\leavevmode\nobreak\ \!V_{0}roman_Γ = 0.25 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT for hoping parameter values in the range 0.150.3V00.150.3subscript𝑉00.15-0.3\leavevmode\nobreak\ \!V_{0}0.15 - 0.3 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and Rashba SOC strengths of about 0.40.45V00.40.45subscript𝑉00.4-0.45\leavevmode\nobreak\ \!V_{0}0.4 - 0.45 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. For Γ0.25V0less-than-or-similar-toΓ0.25subscript𝑉0\Gamma\lesssim 0.25\leavevmode\nobreak\ \!V_{0}roman_Γ ≲ 0.25 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, the maximum topological gap decreases with decreasing Zeeman field because of the limitations imposed by the topological condition, reaching values of about Γ/2)\Gamma/2)roman_Γ / 2 ), while for Γ0.25V0greater-than-or-equivalent-toΓ0.25subscript𝑉0\Gamma\gtrsim 0.25\leavevmode\nobreak\ \!V_{0}roman_Γ ≳ 0.25 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT the maximum topological gap decreases with increasing Zeeman field, because the pairing potential gets suppressed. This general trend also holds for the yellow topological phase, as illustrated in Fig. 9 and Fig. 10. In this case, the optimal Zeeman field values are slightly higher (in the range 0.30.35V00.30.35subscript𝑉00.3-0.35\leavevmode\nobreak\ \!V_{0}0.3 - 0.35 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT), the optimal hoping is t0.12V0similar-to𝑡0.12subscript𝑉0t\sim 0.12\leavevmode\nobreak\ \!V_{0}italic_t ∼ 0.12 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, and the optimal Rashba coupling is in the range 0.5 0.7V00.50.7subscript𝑉00.5\leavevmode\nobreak\ 0.7\!V_{0}0.5 0.7 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, while the maximum value of the topological gap is about V0/8subscript𝑉08V_{0}/8italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / 8. Note that this value is similar to the hopping amplitude.

Our mean-field analysis of the two-dimensional attractive Hubbard-type model with Rashba SOC and Zeeman splitting in the narrow band limit predicts the emergence of topological superfluid phases with gaps of the order 10%12.5%percent10percent12.510\%-12.5\%10 % - 12.5 % of the interaction strength over a significant range of parameters. The optimal hopping amplitude is t0.120.25V0𝑡0.120.25subscript𝑉0t\approx 0.12-0.25\leavevmode\nobreak\ \!V_{0}italic_t ≈ 0.12 - 0.25 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. Further increasing the hopping, i.e., going toward the wide band regime, reduces the DOS, which results in smaller values of the pairing potential. On the other hand, reducing the hopping results in the collapse of the topological phases, as discussed in Sec. III.1. The optimal Zeeman field is in the range 0.250.35V00.250.35subscript𝑉00.25-0.35\leavevmode\nobreak\ \!V_{0}0.25 - 0.35 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. Further increasing the field suppresses the pairing, while reducing it imposes stronger constraints on the values of ΔΔ\Deltaroman_Δ consistent with the topological condition. The optimal Rashba coupling takes vales in the range 0.40.7V00.40.7subscript𝑉00.4-0.7\leavevmode\nobreak\ \!V_{0}0.4 - 0.7 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. Lowering the SOC strength results in the rapid collapse of the pairing potential at finite Zeeman field, while increasing the SOC strength reduces the maximum pairing (at zero field). Finally, the chemical potential for realizing the optimal C=+2𝐶2C=+2italic_C = + 2 (cyan) phase is μ=4t𝜇4𝑡\mu=4titalic_μ = 4 italic_t (half filling), while the optimal C=1𝐶1C=-1italic_C = - 1 (yellow) superfluid phase can be obtained within chemical potential windows of order 0.25V00.25subscript𝑉00.25\leavevmode\nobreak\ \!V_{0}0.25 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, more specifically 2tμ0less-than-or-similar-to2𝑡𝜇less-than-or-similar-to0-2t\lesssim\mu\lesssim 0- 2 italic_t ≲ italic_μ ≲ 0 and 8tμ10tless-than-or-similar-to8𝑡𝜇less-than-or-similar-to10𝑡8t\lesssim\mu\lesssim 10t8 italic_t ≲ italic_μ ≲ 10 italic_t.

IV Conclusion

We study a model of fermions with attractive Hubbard interaction in a square 2D optical lattice in the presence of SOC and a time-reversal symmetry breaking Zeeman splitting. We are interested in the large Zeeman energy regime, known to drive band flattening [24, 25, 26, 27] while simultaneously enabling topological superfluid phases. It has been suggested before that the superconducting/superfluid pairing potential ΔΔ\Deltaroman_Δ increases with decreasing bandwidth and in the limit of vanishing bandwidth (flat-band limit) ΔΔ\Deltaroman_Δ becomes linearly dependent on the attractive interaction V0subscript𝑉0V_{0}italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. This is in contrast to the case of the wide-band regime with a quadratic dispersion (near the ΓΓ\Gammaroman_Γ-point), where the BCS theory applies. According to the BCS theory, the superconducting pair potential ΔΔ\Deltaroman_Δ is proportional to the exponential of the inverse of V0subscript𝑉0V_{0}italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, and thus the pair potential should increase exponentially as the bandwidth narrows. This enhancement of the pair potential with narrowing of the bandwidth is ultimately connected to the enhancement of the density of states at the Fermi surface [28, 29]. In this paper, we explore whether the enhancement of the narrow/flat band density of states could enhance the pairing potential in the TS phase(s) and strengthen the topological superfluid phase by enhancing the corresponding topological gap.

Using the zeros of the energy eigenvalues (at 𝒌=0𝒌0{\bm{k}}=0bold_italic_k = 0), as given by Eq. (5), and the Chern number in Eq. (6), we first show that in the narrow band regime a model with constant pairing potential supports three distinct TS phases that occur in a parameter region significantly expanded as compared to the topological region associated with a wide-band system with quadratic dispersion [23]. Since the former applies to optical lattices and the latter applies to naturally-occurring noncentrosymmetric superconductors [23], we predict that the TS phases should be more easily accessible in optical lattice systems. However, greater accessibility, i.e., an expanded range of control parameters that support TS phases, does not automatically imply more robust topological phases. In general, the robustness of a TS phase is controlled by the topological (quasiparticle) gap that protects it. Using a self-consistent mean field theory to determine the pairing potential, we calculate the dependence of the topological gap on the relevant system parameters, including the hopping amplitude, Rashba coupling, Zeeman field, and chemical potential. As expected, we find that the pairing potential increases with reducing the bandwidth and, generally, decreases with the applied Zeeman field. However, accessing the topological regime requires a large-enough Zeeman field. In particular, ΓΓ\Gammaroman_Γ should exceed the value of the (mean-field) pairing potential, ΔΔ\Deltaroman_Δ. As a result, in the narrow lattice dispersion regime the pairing potential is significantly enhanced as compared to the wide-band, quadratic dispersion regime, but, ultimately, its values in the topological phase are constrained by the Zeeman field ΓΓ\Gammaroman_Γ satisfying the topological condition. In addition, the pairing potential represents an upper bound for the topological gap. To determine the maximum possible values of this gap (within our mean-field approach), we systematically investigate its dependence on the system parameters and identify the optimal regimes. Our best-case scenarios correspond to topological superfluid phases with gaps on the order of 1012.5%10percent12.510-12.5\%10 - 12.5 % of the attractive interaction strength, V0subscript𝑉0V_{0}italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. This is a significant improvement over the topological gaps achievable in naturally occurring systems, such as noncentrosymmetric superconductors.

Our work sets the stage to address additional important open questions regarding correlation effects and the thermal stability of topological superfluidity with broken time-reversal symmetry. The effect of correlations on the pairing potential can be investigated using non-perturbative approaches, such as, for example, the dynamical cluster approximation [34]. On the other hand, the Berenski-Kosterlitz-Thouless transition temperature, TBKTsubscript𝑇BKTT_{\text{BKT}}italic_T start_POSTSUBSCRIPT BKT end_POSTSUBSCRIPT, establishes the temperature below which we can expect to observe topological superfluidity. TBKTsubscript𝑇BKTT_{\text{BKT}}italic_T start_POSTSUBSCRIPT BKT end_POSTSUBSCRIPT is proportional to the superfluid stiffness [8, 35]. In contrast to strictly flat-band systems, the superfluid stiffness is not zero in the present case where the band is narrow but not completely flat. Nonetheless, there should be a contribution to the superfluid stiffness from quantum geometry. Future work will examine the geometric contribution [36, 37, 38, 39] to superfluid stiffness in modeling the topological system considered in this paper with broken time-reversal symmetry.

V Acknowledgement

ST and TDS acknowledge support from ONR-N000142312061. VS acknowledges support from AFOSR (FA2386-21-1-4081, FA9550-23-1-0034, FA9550-19-1-0272). ST and VS acknowledge support from ARO W911NF2210247.

References

  • Read and Green [2000] 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).
  • Qi and Zhang [2011] X.-L. Qi and S.-C. Zhang, Topological insulators and superconductors, Rev. Mod. Phys. 83, 1057 (2011).
  • Zhang et al. [2008] C. Zhang, S. Tewari, R. M. Lutchyn, and S. Das Sarma, px+ipysubscript𝑝𝑥𝑖subscript𝑝𝑦{p}_{x}+i{p}_{y}italic_p start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT + italic_i italic_p start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT superfluid from s𝑠sitalic_s-wave interactions of fermionic cold atoms, Phys. Rev. Lett. 101, 160401 (2008).
  • Sato et al. [2009] M. Sato, Y. Takahashi, and S. Fujimoto, Non-Abelian topological order in s-wave superfluids of ultracold fermionic atoms, Phys. Rev. Lett. 103, 020401 (2009).
  • Bauer and Sigrist [2012] E. Bauer and M. Sigrist, eds., Non-Centrosymmetric Superconductors, Lecture Notes in Physics, Vol. 847 (Springer Berlin Heidelberg, Berlin, Heidelberg, 2012).
  • Smidman et al. [2017] M. Smidman, M. B. Salamon, H. Q. Yuan, and D. F. Agterberg, Superconductivity and spin–orbit coupling in non-centrosymmetric materials: a review, Repts. on Prog. in Phys. 80, 036501 (2017).
  • Zhu et al. [2011] S.-L. Zhu, L.-B. Shao, Z. D. Wang, and L.-M. Duan, Probing non-abelian statistics of majorana fermions in ultracold atomic superfluid, Phys. Rev. Lett. 106, 100404 (2011).
  • Liu et al. [2014] X.-J. Liu, K. T. Law, and T. K. Ng, Realization of 2D spin-orbit interaction and exotic topological orders in cold atoms, Phys. Rev. Lett. 112, 086401 (2014).
  • Galitski and Spielman [2013] V. Galitski and I. B. Spielman, Spin-orbit coupling in quantum gases, Nature 494, 49 (2013).
  • Goldman et al. [2016] N. Goldman, J. C. Budich, and P. Zoller, Topological quantum matter with ultracold gases in optical lattices, Nat. Phys. 12, 639 (2016).
  • Zhang et al. [2018] D.-W. Zhang, Y.-Q. Zhu, Y. X. Zhao, H. Yan, and S.-L. Zhu, Topological quantum matter with cold atoms, Adv. in Phys. 67, 253 (2018).
  • Valdés-Curiel et al. [2021] A. Valdés-Curiel, D. Trypogeorgos, Q.-Y. Liang, R. P. Anderson, and I. B. Spielman, Topological features without a lattice in Rashba spin-orbit coupled atoms, Nat. Comm. 12, 593 (2021).
  • Cheuk et al. [2012] L. W. Cheuk, A. T. Sommer, Z. Hadzibabic, T. Yefsah, W. S. Bakr, and M. W. Zwierlein, Spin-injection spectroscopy of a spin-orbit coupled Fermi gas, Phys. Rev. Lett. 109, 095302 (2012).
  • Song et al. [2016] B. Song, C. He, S. Zhang, E. Hajiyev, W. Huang, X.-J. Liu, and G.-B. Jo, Spin-orbit-coupled two-electron Fermi gases of ytterbium atoms, Phys. Rev. A 94, 061604 (2016).
  • Livi et al. [2016] L. F. Livi, G. Cappellini, M. Diem, L. Franchi, C. Clivati, M. Frittelli, F. Levi, D. Calonico, J. Catani, M. Inguscio, and L. Fallani, Synthetic dimensions and spin-orbit coupling with an optical clock transition, Phys. Rev. Lett. 117, 220401 (2016).
  • Huang et al. [2016] 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 ultracold Fermi gases, Nature 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, Experimental observation of a topological band gap opening in ultracold fermi gases with two-dimensional spin-orbit coupling, Phys. Rev. Lett. 117, 235304 (2016).
  • Kolkowitz et al. [2017] S. Kolkowitz, S. L. Bromley, T. Bothwell, M. L. Wall, G. E. Marti, A. P. Koller, X. Zhang, A. M. Rey, and J. Ye, Spin–orbit-coupled fermions in an optical lattice clock, Nature 542, 66 (2017).
  • Song et al. [2019] B. Song, C. He, S. Niu, L. Zhang, Z. Ren, X.-J. Liu, and G.-B. Jo, Observation of nodal-line semimetal with ultracold fermions in an optical lattice, Nat. Phys. 15, 911 (2019).
  • Aeppli et al. [2022] A. Aeppli, A. Chu, T. Bothwell, C. J. Kennedy, D. Kedar, P. He, A. M. Rey, and J. Ye, Hamiltonian engineering of spin-orbit–coupled fermions in a Wannier-Stark optical lattice clock, Sci. Adv. 8, eadc9242 (2022).
  • Lauria et al. [2022] P. Lauria, W.-T. Kuo, N. R. Cooper, and J. T. Barreiro, Experimental realization of a fermionic spin-momentum lattice, Phys. Rev. Lett. 128, 245301 (2022).
  • Liang et al. [2023] M.-C. Liang, Y.-D. Wei, L. Zhang, X.-J. Wang, H. Zhang, W.-W. Wang, W. Qi, X.-J. Liu, and X. Zhang, Realization of Qi-Wu-Zhang model in spin-orbit-coupled ultracold fermions, Phys. Rev. Research 5, L012006 (2023).
  • Tewari et al. [2011] S. Tewari, T. D. Stanescu, J. D. Sau, and S. Das Sarma, Topologically non-trivial superconductivity in spin-orbit coupled systems: bulk phases and quantum phase transitions, New J. Phys. 13, 065004 (2011).
  • Zhang and Zhang [2013] Y. Zhang and C. Zhang, Bose-einstein condensates in spin-orbit-coupled optical lattices: Flat bands and superfluidity, Phys. Rev. A 87, 023611 (2013).
  • Lin et al. [2014] F. Lin, C. Zhang, and V. W. Scarola, Emergent kinetics and fractionalized charge in 1d spin-orbit coupled flatband optical lattices, Phys. Rev. Lett. 112, 110404 (2014).
  • Chen and Scarola [2016] M. Chen and V. W. Scarola, Stability of emergent kinetics in optical lattices with artificial spin-orbit coupling, Phys. Rev. A 94, 043601 (2016).
  • Hui et al. [2017] H.-Y. Hui, Y. Zhang, C. Zhang, and V. W. Scarola, Superfluidity in the absence of kinetics in spin-orbit-coupled optical lattices, Phys. Rev. A 95, 033603 (2017).
  • Kopnin et al. [2011] N. B. Kopnin, T. T. Heikkilä, and G. E. Volovik, High-temperature surface superconductivity in topological flat-band systems, Phys. Rev. B 83, 220503 (2011).
  • Heikkilä et al. [2011] T. T. Heikkilä, N. B. Kopnin, and G. E. Volovik, Flat bands in topological media, Jetp Lett. 94, 233 (2011).
  • Han et al. [2023] R. Han, F. Yuan, and H. Zhao, Phase diagram, band structure and density of states in two-dimensional attractive fermi-hubbard model with rashba spin-orbit coupling, New Journal of Physics 25, 023011 (2023).
  • Schnyder et al. [2008] 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).
  • Volovik and Yakovenko [1989] G. E. Volovik and V. M. Yakovenko, Fractional charge, spin and statistics of solitons in superfluid 3he film, Journal of Physics: Condensed Matter 1, 5263 (1989).
  • Ghosh et al. [2010] P. Ghosh, J. D. Sau, S. Tewari, and S. Das Sarma, Non-abelian topological order in noncentrosymmetric superconductors with broken time-reversal symmetry, Phys. Rev. B 82, 184525 (2010).
  • Doak et al. [2023] P. Doak, G. Balduzzi, P. Laurell, E. Dagotto, and T. A. Maier, Spin-singlet topological superconductivity in the attractive rashba-hubbard model, Phys. Rev. B 107, 224501 (2023).
  • Xu and Zhang [2015] Y. Xu and C. Zhang, Berezinskii-Kosterlitz-Thouless phase transition in 2d spin-orbit-coupled Fulde-Ferrell superfluids, Phys. Rev. Lett. 114, 110401 (2015).
  • Peotta and Törmä [2015] S. Peotta and P. Törmä, Superfluidity in topologically nontrivial flat bands, Nat. Commun. 6, 8944 (2015).
  • Julku et al. [2016] A. Julku, S. Peotta, T. I. Vanhala, D.-H. Kim, and P. Törmä, Geometric origin of superfluidity in the lieb-lattice flat band, Phys. Rev. Lett. 117, 045303 (2016).
  • Liang et al. [2017] L. Liang, T. I. Vanhala, S. Peotta, T. Siro, A. Harju, and P. Törmä, Band geometry, Berry curvature, and superfluid weight, Phys. Rev. B 95, 024515 (2017).
  • Hu et al. [2019] X. Hu, T. Hyart, D. I. Pikulin, and E. Rossi, Geometric and conventional contribution to the superfluid weight in twisted bilayer graphene, Phys. Rev. Lett. 123, 237002 (2019).