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

Kramers Fulde-Ferrell state and superconducting spin diode effect

Yi Zhang [email protected] Department of Physics, Shanghai University, Shanghai 200444, China    Ziqiang Wang [email protected] Department of Physics, Boston College, Chestnut Hill, MA 02467, USA
Abstract

We study a novel equal-spin pairing state with opposite center of mass momentum for each spin polarization. This state, dubbed a Kramers Fulde-Ferrell (KFF) state, respects time-reversal symmetry and can be realized in a one-dimensional system with spin-orbit coupling and nearest neighbor attraction. We find that the KFF state supports nonreciprocal spin transport for both bulk superconductor and Josephson junctions. In addition to the spin Josephson diode effect, the charge transport is controlled by intriguing dynamics of bound states whose transitions can be manipulated by the length of the KFF superconductor. The KFF state is relevant for embedded quantum structures in monolayer Fe-based superconductors and dissipationless superconducting spintronics.

I Introduction

Recent experimental observations of the diode effect in superconductors [1, 2, 3, 4] and Josephson junctions (JJ) [5, 6, 7, 8, 9] have stimulated the research of nonreciprocal transport properties in superconducting (SC) systems. Following the proposal of SC diode effect [10], the so-called ϕ0\phi_{0} Josephson state has been extensively studied [11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23] as a possible mechanism to realize nonreciprocal transport in JJs. More recently, many theoretical proposals [24, 25, 26, 27, 28, 29, 30, 31] have been put forward for SC diode in bulk superconductors. In particular, the finite-momentum pairing Flude-Ferrell-Larkin-Ovchinnikov state [32, 33] is believed to provide a physical mechanism, since the order parameter of Fulde-Ferrell (FF) state Δ(r)=Δeiqr\Delta(\textbf{r})=\Delta e^{i\textbf{q}\cdot\textbf{r}} can directly generate a difference in the critical current along and against the direction of q, leading to SC diode effect [28, 27]. The FF order, also known as helical superconductivity, can be realized in noncentrosymmetric superconductors with spin-orbit coupling (SOC) and time reversal symmetry breaking fields [34, 35, 36, 37, 38, 39, 40, 41, 42, 43, 44, 45, 46, 47].

So far, the study of SC diode effect focused on the nonreciprocity of charge transport. One may wonder if there exists a similar nonreciprocal property in spin transport in certain SC systems. From the symmetry point of view, SC diode effect in charge transport, where critical currents in opposite directions have different magnitudes, requires the system to break both inversion and time-reversal (𝒯{\cal T}) symmetry, since the charge current operator changes sign under either inversion or 𝒯{\cal T}. While the spin current operator changes sign under inversion, it is invariant under 𝒯{\cal T}. Thus, the nonreciprocity in spin transport only requires breaking inversion symmetry and can be realized in 𝒯{\cal T} invariant superconductors.

In this article, we propose a novel SC state that can realize nonreciprocal spin transport. This state has equal-spin pairing and a FF type of order parameter Δσ(r)=ΔeiσQr\Delta_{\sigma}(\textbf{r})=\Delta e^{i\sigma\textbf{Q}\cdot\textbf{r}}, with opposite Cooper pair center of mass momentum for opposite spin polarizations as shown schematically in Fig. 1(a). We term this SC state as a Kramers FF (KFF) state since 𝒯{\cal T} symmetry is maintained. Such FF state has pairing field with only one Q vector in each pairing channel, which is translational invariant unlike the Larkin-Ovchinnikov state also known as the pair density wave state, where the pairing field has both Q and Q-\textbf{Q} vectors in each channel so that the pairing order parameters varies in space. We demonstrate that the KFF state can be realized in a meanfield theory of a concrete model describing a spin-orbit coupled chain with nearest neighbor attractions as illustrated in Fig. 1(a,b). The nonzero Q pairing across the Fermi points in Fig. 1(a) is enabled by the SOC split bands. We study the condition to realize the nonreciprocal spin transport where the critical spin current along positive and negative directions are unequal in magnitude for both bulk SC state and Josephson junction structures. Moreover, we find intriguing properties and rich phases in the charge transport across JJs of the KFF state, which can be realized by simply changing the length of the SC chain. Similar pairing state was also studied in the two dimensional honeycomb system where the valley degree of freedom plays the role of spin here [48].

This article is organized as follows. We start with the introduction of our model Hamiltonian and meanfield formulation for the KFF state in Sec. II. In Sec. III, we discuss the nonreciprocal spin transport for the bulk SC with KFF order, which is followed by the discussion of the hidden inversion symmetry that is related to the nonreciprocal spin transport in Sec. IV Then we study the transport properties of the Josephson junction structure constructed from the KFF state in Sec. V and discuss various phases realized in the charge transport across JJs of the KFF state in Sec. VI. We finalize the discussion in Sec. VII.

II Formulation

II.1 Model Hamiltonian

We first consider a one-dimensional (1D) spin-orbit coupled chain with nearest neighbor attraction described by the Hamiltonian

H^=i,j,σtijciσcjσ+αiiσciσci+1σ+h.c.Vinini+1\hat{H}=-\sum_{i,j,\sigma}t_{ij}c_{i\sigma}^{\dagger}c_{j\sigma}+\alpha\sum_{i}i\sigma c_{i\sigma}^{\dagger}c_{i+1\sigma}+h.c.-V\sum_{i}n_{i}n_{i+1} (1)

where tijt_{ij} are hopping parameters up to the 2nd neighbor (t1t_{1} and t2t_{2}) and α\alpha describes a nearest neighbor SOC. In 1D, SOC leaves a conserved spin quantum number which is taken to be the spin quantization axis along the chain direction. The nearest neighbor attraction VV responsible for SC order can be decomposed into equal-spin and opposite-spin pairing channels as

HI=Vinini+1=Vi,σσciσci+1σci+1σciσ=VNck,k,q,σσei(kk)ck+q2σck+q2σck+q2σck+q2σ=VNck,k,q,σsinksinkck+q2σck+q2σck+q2σck+q2σVNck,k,q,σei(kk)ck+q2σck+q2σ¯ck+q2σ¯ck+q2σ\begin{split}H_{I}&=-V\sum_{i}n_{i}n_{i+1}=-V\sum_{i,\sigma\sigma^{\prime}}c_{i\sigma}^{\dagger}c_{i+1\sigma^{\prime}}^{\dagger}c_{i+1\sigma^{\prime}}c_{i\sigma}\\ &=-\frac{V}{N_{c}}\sum_{k,k^{\prime},q,\sigma\sigma^{\prime}}e^{i(k-k^{\prime})}c_{k+\frac{q}{2}\sigma}^{\dagger}c_{-k+\frac{q}{2}\sigma^{\prime}}^{\dagger}c_{-k^{\prime}+\frac{q}{2}\sigma^{\prime}}c_{k^{\prime}+\frac{q}{2}\sigma}\\ &=-\frac{V}{N_{c}}\sum_{k,k^{\prime},q,\sigma}\sin k\sin k^{\prime}c_{k+\frac{q}{2}\sigma}^{\dagger}c_{-k+\frac{q}{2}\sigma}^{\dagger}c_{-k^{\prime}+\frac{q}{2}\sigma}c_{k^{\prime}+\frac{q}{2}\sigma}\\ &-\frac{V}{N_{c}}\sum_{k,k^{\prime},q,\sigma}e^{i(k-k^{\prime})}c_{k+\frac{q}{2}\sigma}^{\dagger}c_{-k+\frac{q}{2}\bar{\sigma}}^{\dagger}c_{-k^{\prime}+\frac{q}{2}\bar{\sigma}}c_{k^{\prime}+\frac{q}{2}\sigma}\end{split} (2)

where the first term corresponds to the attraction between the electrons with the same spin and the second term corresponds to the attraction between the electrons with opposite spins and NcN_{c} is the number of sites. If we further define the two pairing operators in equal-spin and opposite-spin channels as

{Δ^,q,σ=1Nckisinkck+q2σck+q2σΔ^,q,σ=1Nckeikck+q2σ¯ck+q2σ\begin{cases}\hat{\Delta}_{\parallel,q,\sigma}=\frac{1}{N_{c}}\sum_{k}i\sin kc_{-k+\frac{q}{2}\sigma}c_{k+\frac{q}{2}\sigma}\\ \hat{\Delta}_{\perp,q,\sigma}=\frac{1}{N_{c}}\sum_{k}e^{-ik}c_{-k+\frac{q}{2}\bar{\sigma}}c_{k+\frac{q}{2}\sigma}\end{cases} (3)

Eq. 2 can be written as

HI=NcV1qΔ^,q,σΔ^,q,σNcV2qΔ^,q,σΔ^,q,σH_{I}=-N_{c}V_{1}\sum_{q}\hat{\Delta}_{\parallel,q,\sigma}^{\dagger}\hat{\Delta}_{\parallel,q,\sigma}-N_{c}V_{2}\sum_{q}\hat{\Delta}_{\perp,q,\sigma}^{\dagger}\hat{\Delta}_{\perp,q,\sigma} (4)

where these two terms correspond to the pairing channels with equal and opposite spin respectively and here we denote the attraction in these two channels as V1V_{1} and V2V_{2}. While V1=V2=VV_{1}=V_{2}=V in the original model in Eq. (1), we consider here a more general model where the effective attraction V1V_{1} and V2V_{2} can be different. The equal-spin pairing can be induced in embedded quantum structures in high-Tc superconductors due to spatial symmetry breaking [49], such as along the atomic line defects in monolayer FeTeSe  [50]. Then the total Hamiltonian becomes

H^=k,σεk,σckσckσNcV1qΔ^,q,σΔ^,q,σNcV2qΔ^,q,σΔ^,q,σ\hat{H}=\sum_{k,\sigma}\varepsilon_{k,\sigma}c_{k\sigma}^{\dagger}c_{k\sigma}-N_{c}V_{1}\sum_{q}\hat{\Delta}_{\parallel,q,\sigma}^{\dagger}\hat{\Delta}_{\parallel,q,\sigma}-N_{c}V_{2}\sum_{q}\hat{\Delta}_{\perp,q,\sigma}^{\dagger}\hat{\Delta}_{\perp,q,\sigma} (5)

where

εkσ=2tαcos(kσθα)2t2cos(2k)\varepsilon_{k\sigma}=-2t_{\alpha}\cos(k-\sigma\theta_{\alpha})-2t_{2}\cos(2k) (6)

is the band dispersion with

{tα=t12+α2θα=arctan(α/t1)\begin{cases}t_{\alpha}=\sqrt{t_{1}^{2}+\alpha^{2}}\\ \theta_{\alpha}=\arctan(\alpha/t_{1})\end{cases} (7)

which determines the positions of the Fermi points.

II.2 Meanfield decoupling

From the structure of Fermi points shown in Fig. 1(a), we can solve the model in Eq. (5) within a meanfield approximation assuming the following meanfield ansatz

{Δ^,Q,=Δ^,Q,=ΔΔ^,0,=ΔeiϕΔ^,0,=Δeiϕ\begin{cases}\left\langle\hat{\Delta}_{\parallel,Q,\uparrow}\right\rangle=\left\langle\hat{\Delta}_{\parallel,-Q,\downarrow}\right\rangle=\Delta_{\parallel}\\ \left\langle\hat{\Delta}_{\perp,0,\uparrow}\right\rangle=\Delta_{\perp}e^{i\phi_{\perp}}\\ \left\langle\hat{\Delta}_{\perp,0,\downarrow}\right\rangle=-\Delta_{\perp}e^{-i\phi_{\perp}}\end{cases} (8)

Here, in equal-spin pairing channel, electrons with up (down) spin pair into the FF state with a nonzero center of mass momentum Q(-Q). The resulting Kramers doublet ensures that 𝒯{\cal T} symmetry is preserved. In opposite-spin pairing channel, electrons with up and down spins form zero-momentum pairs, which is in general a mixture of ss-wave and pzp_{z}-wave pairing depending on the phase ϕ\phi_{\perp}. Specifically, ϕ=0\phi_{\perp}=0 corresponds to ss-wave pairing and ϕ=π2\phi_{\perp}=\frac{\pi}{2} corresponds to pzp_{z}-wave pairing, while other values give rise to a mixed parity state.

After the meanfield decoupling, the meanfield Hamiltonian can be written as

H^MFμN^=kσ(εkσμ)ckσckσV1Δkσ(isink)ck+σQ2σck+σQ2σ2V2Δkckckcos(k+ϕ)+h.c.+2Nc(V1Δ2+V2Δ2)\begin{split}\hat{H}_{MF}-\mu\hat{N}&=\sum_{k\sigma}\left(\varepsilon_{k\sigma}-\mu\right)c_{k\sigma}^{\dagger}c_{k\sigma}\\ &-V_{1}\Delta_{\parallel}\sum_{k\sigma}(-i\sin k)c_{k+\frac{\sigma Q}{2}\sigma}^{\dagger}c_{-k+\frac{\sigma Q}{2}\sigma}^{\dagger}\\ &-2V_{2}\Delta_{\perp}\sum_{k}c_{k\uparrow}^{\dagger}c_{-k\downarrow}^{\dagger}\cos\left(k+\phi_{\perp}\right)+h.c.\\ &+2N_{c}(V_{1}\Delta_{\parallel}^{2}+V_{2}\Delta_{\perp}^{2})\end{split} (9)

In the Nambu basis ψk=(ck+Q2,ckQ2,ck+Q2,ckQ2)\psi_{k}^{\dagger}=\left(c_{k+\frac{Q}{2}\uparrow}^{\dagger},c_{k-\frac{Q}{2}\downarrow}^{\dagger},c_{-k+\frac{Q}{2}\uparrow},c_{-k-\frac{Q}{2}\downarrow}\right), Eq. 9 can be written as

HMFμN=12kψkhkψk+2NcV1Δ2+2NcV2Δ2μNcH_{MF}-\mu N=\frac{1}{2}\sum_{k}\psi_{k}^{\dagger}h_{k}\psi_{k}+2N_{c}V_{1}\Delta_{\parallel}^{2}+2N_{c}V_{2}\Delta_{\perp}^{2}-\mu N_{c} (10)

with

hk=[εk+Q2,μ02iV1Δsink2V2Δcos(k+Q2+ϕ)0εkQ2,μ2V2Δcos(k+Q2+ϕ)2iV1Δsink2iV1Δsink2V2Δcos(k+Q2+ϕ)εk+Q2,+μ02V2Δcos(k+Q2+ϕ)2iV1Δsink0εkQ2,+μ]h_{k}=\left[\begin{array}[]{cccc}\varepsilon_{k+\frac{Q}{2},\uparrow}-\mu&0&2iV_{1}\Delta_{\parallel}\sin k&-2V_{2}\Delta_{\perp}\cos(k+\frac{Q}{2}+\phi_{\perp})\\ 0&\varepsilon_{k-\frac{Q}{2},\downarrow}-\mu&2V_{2}\Delta_{\perp}\cos(-k+\frac{Q}{2}+\phi_{\perp})&2iV_{1}\Delta_{\parallel}\sin k\\ -2iV_{1}\Delta_{\parallel}\sin k&2V_{2}\Delta_{\perp}\cos(-k+\frac{Q}{2}+\phi_{\perp})&-\varepsilon_{-k+\frac{Q}{2},\uparrow}+\mu&0\\ -2V_{2}\Delta_{\perp}\cos(k+\frac{Q}{2}+\phi_{\perp})&-2iV_{1}\Delta_{\parallel}\sin k&0&-\varepsilon_{k-\frac{Q}{2},\downarrow}+\mu\end{array}\right] (11)

This meanfield Hamiltonian can be solved self-consistently for a fixed chemical potential μ\mu and various values of QQ and ϕ\phi_{\perp} with the self-consistent equations

{Δ=12Nckσisinkck+σQ2,σck+σQ2,σΔ=1Nckcos(k+ϕ)ckck=1Nckcos(k+Q2+ϕ)ckQ2ck+Q2\begin{cases}\Delta_{\parallel}=\frac{1}{2N_{c}}\sum_{k\sigma}i\sin k\left\langle c_{-k+\frac{\sigma Q}{2},\sigma}c_{k+\frac{\sigma Q}{2},\sigma}\right\rangle\\ \begin{split}\Delta_{\perp}&=\frac{1}{N_{c}}\sum_{k}\cos(k+\phi_{\perp})\left\langle c_{-k\downarrow}c_{k\uparrow}\right\rangle\\ &=\frac{1}{N_{c}}\sum_{k}\cos(k+\frac{Q}{2}+\phi_{\perp})\left\langle c_{-k-\frac{Q}{2}\downarrow}c_{k+\frac{Q}{2}\uparrow}\right\rangle\end{split}\end{cases} (12)

and the ground state is determined by the states with the lowest free energy density Ω=1NcHMFμN\Omega=\frac{1}{N_{c}}\left\langle H_{MF}-\mu N\right\rangle which also determines the value of QQ and ϕ\phi_{\perp}.

We perform the calculation with a general set of parameter and the obtained meanfield phase diagram in Fig. 1(b) shows that the novel KFF state is a more stable ground state than the mixed parity state when V1V_{1} is larger than V2V_{2}. We thus focus on the KFF state driven by equal-spin pairing and investigate its many intriguing properties. A more detailed analysis of the meanfield phase diagram as well as the behavior of the order parameters are shown in Appendix. A. The detailed calculation for the mixed parity state is also shown in Appendix. E.

Refer to caption
Figure 1: (a) Schematic band structure of the spin-orbit coupled 1D chain, showing the two pairing channels across the Fermi points. (b) Phase diagram obtained from meanfield calculations with parameters t1=1t_{1}=1, t2=0.5t_{2}=-0.5, α=0.4\alpha=0.4, μ=0.53\mu=-0.53. (c) Schematic illustration of nonreciprocal spin transport due to different critical spin currents |js,c|>|js,c+||j_{s,c-}|>|j_{s,c+}| in “++”and “-” directions. A spin current |js,c+|<|js,±|<|js,c||j_{s,c+}|<|j_{s,\pm}|<|j_{s,c-}| flows as a dissipationless supercurrent in “-” direction (marked by |js||j_{s-}|), but can only be transported as a normal dissipative current in “++” direction (marked as |js+||j_{s+}|).

III Nonreciprocal spin transport and Spin diode effect in bulk KFF state

The meanfield Hamiltonian in KFF state becomes

H^MFμN^=kσ(εkσμ)ckσckσ+V1Δkσisinkck+σQ2,σck+σQ2,σ+h.c.+2NcV1Δ2\begin{split}\hat{H}_{MF}-\mu\hat{N}&=\sum_{k\sigma}\left(\varepsilon_{k\sigma}-\mu\right)c_{k\sigma}^{\dagger}c_{k\sigma}\\ &+V_{1}\Delta_{\parallel}\sum_{k\sigma}i\sin kc_{k+\frac{\sigma Q}{2},\sigma}^{\dagger}c_{-k+\frac{\sigma Q}{2},\sigma}^{\dagger}+h.c.+2N_{c}V_{1}\Delta_{\parallel}^{2}\end{split} (13)

which can be written in the Nambu basis ψkσ=(ck+σQ2,σ,ck+σQ2,σ)\psi_{k\sigma}^{\dagger}=\left(c_{k+\frac{\sigma Q}{2},\sigma}^{\dagger},c_{-k+\frac{\sigma Q}{2},\sigma}\right) as

H^MFμN^=12kσψkσhk,Q,σψkσ+2NcV1Δ2μNc\hat{H}_{MF}-\mu\hat{N}=\frac{1}{2}\sum_{k\sigma}\psi_{k\sigma}^{\dagger}h_{k,Q,\sigma}\psi_{k\sigma}+2N_{c}V_{1}\Delta_{\parallel}^{2}-\mu N_{c} (14)

where,

hk,Q,σ=[εk+σQ2,σμ2iV1Δsink2iV1Δsinkεk+σQ2,σ+μ]h_{k,Q,\sigma}=\left[\begin{array}[]{cc}\varepsilon_{k+\frac{\sigma Q}{2},\sigma}-\mu&2iV_{1}\Delta_{\parallel}\sin k\\ -2iV_{1}\Delta_{\parallel}\sin k&-\varepsilon_{-k+\frac{\sigma Q}{2},\sigma}+\mu\end{array}\right] (15)

is block diagonal in spin space, leading to the eigenenergy as

Ekσ,±=12(εk+σQ2,σεk+σQ2,σ)±[12(εk+σQ2,σ+εk+σQ2,σ)μ]2+4V12Δ2sin2k\begin{split}E_{k\sigma,\pm}&=\frac{1}{2}\left(\varepsilon_{k+\frac{\sigma Q}{2},\sigma}-\varepsilon_{-k+\frac{\sigma Q}{2},\sigma}\right)\\ &\pm\sqrt{\left[\frac{1}{2}(\varepsilon_{k+\frac{\sigma Q}{2},\sigma}+\varepsilon_{-k+\frac{\sigma Q}{2},\sigma})-\mu\right]^{2}+4V_{1}^{2}\Delta_{\parallel}^{2}\sin^{2}k}\end{split} (16)

and we have Ekσ,±=Ekσ¯,±E_{k\sigma,\pm}=E_{-k\bar{\sigma},\pm} due to the 𝒯{\cal T} symmetry. Then the free energy density at zero temperature Ω(Δ,Q)\Omega(\Delta_{\parallel},Q) can be calculated as

Ω(Δ,Q)=1NcH^MFμN^=12Nskσ,n=±Ekσ,nΘ(Ekσ,n)+2V1Δ2μ\begin{split}\Omega(\Delta_{\parallel},Q)&=\frac{1}{N_{c}}\left\langle\hat{H}_{MF}-\mu\hat{N}\right\rangle\\ &=\frac{1}{2N_{s}}\sum_{k\sigma,n=\pm}E_{k\sigma,n}\Theta(-E_{k\sigma,n})+2V_{1}\Delta_{\parallel}^{2}-\mu\end{split} (17)

where Θ(x)\Theta(x) the Heaviside step function. The order parameter Δ\Delta_{\parallel} for a given QQ can be determined self-consistently by minimizing Ω(Δ,Q)\Omega(\Delta_{\parallel},Q) with respect to Δ\Delta_{\parallel}, leading to the self-consistency equation

Δ=12Nckσisinkck+σQ2,σck+σQ2,σ\Delta_{\parallel}=\frac{1}{2N_{c}}\sum_{k\sigma}i\sin k\left\langle c_{-k+\frac{\sigma Q}{2},\sigma}c_{k+\frac{\sigma Q}{2},\sigma}\right\rangle (18)

The optimal QQ value can be further determined by minimizing Ω(Δ,Q)\Omega(\Delta_{\parallel},Q) with respect to QQ, i.e., QΩ(Δ,Q)=0\partial_{Q}\Omega(\Delta_{\parallel},Q)=0. Because the latter is directly related to the spin current carried by the KFF state (see Appendix. B) js(Q)=σσ2jσ=QΩ(Δ,Q,T)j_{s}(Q)=\sum_{\sigma}\frac{\sigma}{2}j_{\sigma}=\partial_{Q}\Omega(\Delta_{\parallel},Q,T), the ground state with optimized Q=Q0Q=Q_{0} does not carry any net spin current since QΩ(Δ,Q0,T)=0\partial_{Q}\Omega(\Delta_{\parallel},Q_{0},T)=0. When the KFF state is driven out of equilibrium into a state with QQ0Q\neq Q_{0}, a nonzero applied spin current js(QQ0)=js0j_{s}(Q\neq Q_{0})=j_{s}\neq 0 is realized. Throughout the remaining text, we define the charge and spin currents in units of e\frac{e}{\hbar} and unity, respectively. Since the spin current carrying state has equal but opposite center of mass momentum ±Q\pm Q for Cooper pairs in opposite spin channels, the charge current always vanishes due to 𝒯\cal T symmetry. The critical spin currents in ++ and - directions are determined by the maximum and minimum values of js(Q)j_{s}(Q) sustained by the SC state according to js,c+=maxQ[js(Q)]j_{s,c+}=\max_{Q}[j_{s}(Q)] and js,c=minQ[js(Q)]j_{s,c-}=\min_{Q}[j_{s}(Q)]. When |js,c+||js,c|\left|j_{s,c+}\right|\neq\left|j_{s,c-}\right|, the SC state enables nonreciprocal spin transport as shown in Fig. 1(c).

We performed meanfield calculations using two sets of parameters. In the first case, we set t2=0t_{2}=0 and obtain analytically that the ground state has Q0=2θα=σ(kfσ,++kfσ,)Q_{0}=2\theta_{\alpha}=\sigma(k_{f\sigma,+}+k_{f\sigma,-}), consistent with the SOC split bands where electrons pair across the Fermi points kfσ,±k_{f\sigma,\pm} in the same spin sector, giving rise to finite Cooper pair momenta σQ0\sigma Q_{0} with more details shown in Appendix. C. The zero temperature free energy in Eq. (17) and the current density from its momentum derivative are calculated numerically and plotted in Figs. 2(a, c) as a function of QQ. The critical spin currents js,c±j_{s,c\pm} are determined by the maximum and minimum values of the spin current density. Fig. 2(c) shows that the two critical momenta Q±Q_{\pm} at which the spin current reaches critical values coincide with the two momenta where the SC order parameter Δ(Q±)\Delta_{\parallel}(Q_{\pm}) vanishes. In this case with t2=0t_{2}=0, we find that the critical spin currents js,c+=js,cj_{s,c+}=-j_{s,c-}, as shown in Fig. 2(a), and the spin transport is reciprocal. The absence of nonreciprocal spin transport turns out to be due to a hidden inversion symmetry when t2=0t_{2}=0.

IV Hidden inversion symmetry

To demonstrate the hidden inversion symmetry, we can perform a local gauge transformation ciσei2σQxidiσc_{i\sigma}^{\dagger}\rightarrow e^{-\frac{i}{2}\sigma Qx_{i}}d_{i\sigma}^{\dagger}, corresponding to ckσdkσQ2,σc_{k\sigma}^{\dagger}\rightarrow d_{k-\frac{\sigma Q}{2},\sigma}^{\dagger} in momentum space. Then the meanfield Hamiltonian Eq. 13 becomes

H^MFμN^=kσ(εkσμ)dkσQ2,σdkσQ2,σ+V1Δkσisinkdk,σdk,σ+h.c.+2NcV1Δ2=kσ(εk+σQ2,σμ)dk,σdk,σ+V1Δkσisinkdk,σdk,σ+h.c.+2NcV1Δ2\begin{split}\hat{H}_{MF}-\mu\hat{N}&=\sum_{k\sigma}\left(\varepsilon_{k\sigma}-\mu\right)d_{k-\frac{\sigma Q}{2},\sigma}^{\dagger}d_{k-\frac{\sigma Q}{2},\sigma}\\ &+V_{1}\Delta_{\parallel}\sum_{k\sigma}i\sin kd_{k,\sigma}^{\dagger}d_{-k,\sigma}^{\dagger}+h.c.+2N_{c}V_{1}\Delta_{\parallel}^{2}\\ &=\sum_{k\sigma}\left(\varepsilon_{k+\frac{\sigma Q}{2},\sigma}-\mu\right)d_{k,\sigma}^{\dagger}d_{k,\sigma}\\ &+V_{1}\Delta_{\parallel}\sum_{k\sigma}i\sin kd_{k,\sigma}^{\dagger}d_{-k,\sigma}^{\dagger}+h.c.+2N_{c}V_{1}\Delta_{\parallel}^{2}\end{split} (19)

Since in the absence of t2t_{2}, εk+σQ2,σ=2tαcos(k+σQ2σθα)\varepsilon_{k+\frac{\sigma Q}{2},\sigma}=-2t_{\alpha}\cos(k+\frac{\sigma Q}{2}-\sigma\theta_{\alpha}), we can see that in the new basis, the only inversion breaking term θα\theta_{\alpha} owing to the spin-orbit coupling α\alpha, is cancelled if Q=Q0=2θαQ=Q_{0}=2\theta_{\alpha}. In other words, when Q=Q0=2θαQ=Q_{0}=2\theta_{\alpha}, the inversion symmetry can be recovered in the new basis, where the meanfield Hamiltonian becomes

H^MFμN^=kσ(2tαcoskμ)dk,σdk,σ+V1Δkσisinkdk,σdk,σ+h.c.+2NcV1Δ2\begin{split}\hat{H}_{MF}-\mu\hat{N}&=\sum_{k\sigma}\left(-2t_{\alpha}\cos k-\mu\right)d_{k,\sigma}^{\dagger}d_{k,\sigma}\\ &+V_{1}\Delta_{\parallel}\sum_{k\sigma}i\sin kd_{k,\sigma}^{\dagger}d_{-k,\sigma}^{\dagger}+h.c.+2N_{c}V_{1}\Delta_{\parallel}^{2}\end{split} (20)

The transformed Hamiltonian describes two spin-degenerate pp-wave Kitaev chains [51] with inversion symmetry in each spin sector. It is precisely this hidden inversion symmetry that forbids the nonreciprocal property of spin current, since this hidden inversion symmetry changes the sign of the current operator j^σ\hat{j}_{\sigma}, while keeping the total Hamiltonian invariant, which guarantees a one to one correspondence between the positive and negative current, such that js,c+j_{s,c+} and js,cj_{s,c-} have to have the same magnitude.

In the case with finite t2t_{2}, the dispersion of the noninteracting Hamiltonian reads εkσ=2tαcos(kσθα)2t2cos(2k)\varepsilon_{k\sigma}=-2t_{\alpha}\cos(k-\sigma\theta_{\alpha})-2t_{2}\cos(2k), and then the Fermi momentum kfσ,±k_{f\sigma,\pm} no longer have a closed form and the sum of the two Fermi momentum belonging to the same spin polarization is incommensurate in general. Here, we can immediately see that the gauge transformation above can not recover the inversion symmetry as above, since now εk+σQ2,σ=2tαcos(k+σQ2σθα)2t2cos(2k+σQ)\varepsilon_{k+\frac{\sigma Q}{2},\sigma}=-2t_{\alpha}\cos(k+\frac{\sigma Q}{2}-\sigma\theta_{\alpha})-2t_{2}\cos(2k+\sigma Q), and the inversion breaking phase of the two cosine function σQ2σθα\frac{\sigma Q}{2}-\sigma\theta_{\alpha} and σQ\sigma Q cannot be cancelled by QQ simultaneously, so that there is no hidden inversion symmetry that forbids the presence of the nonreciprocal spin transport. Therefore, the transformed model describes the two p-wave Kitaev chains with complex hoppings, which are time-reversal counterparts of each other but not identical and can be written in the real space as

H^MFμN^=iσ(tαeiσ(Q2θα)di,σdi+1,σt2eiσQdi,σdi+2,σ+V1Δdi,σdi+1,σ+h.c.μdi,σdi,σ)+2NcV1Δ2\begin{split}\hat{H}_{MF}-\mu\hat{N}&=\sum_{i\sigma}\Bigl{(}-t_{\alpha}e^{i\sigma(\frac{Q}{2}-\theta_{\alpha})}d_{i,\sigma}^{\dagger}d_{i+1,\sigma}-t_{2}e^{i\sigma Q}d_{i,\sigma}^{\dagger}d_{i+2,\sigma}\\ &+V_{1}\Delta_{\parallel}d_{i,\sigma}^{\dagger}d_{i+1,\sigma}^{\dagger}+h.c.-\mu d_{i,\sigma}^{\dagger}d_{i,\sigma}\Bigl{)}+2N_{c}V_{1}\Delta_{\parallel}^{2}\end{split} (21)

Indeed, the results shown in Fig. 2(b) confirms that critical spin currents are nonreciprocal with js,c+=0.11j_{s,c+}=0.11 and js,c=0.20j_{s,c-}=-0.20 along ++ and - directions, respectively. As a result, a spin current jsj_{s} satisfying 0.11<|js|<0.200.11<|j_{s}|<0.20 flows as dissipationless supercurrent in the negative direction since |js|<|js,c||j_{s}|<|j_{s,c-}|, but can only be transported as a dissipative normal current in the positive direction since |js|>|js,c+||j_{s}|>|j_{s,c+}|. This SC spin diode effect is shown schematically in Fig. 1(c).

Refer to caption
Figure 2: Spin current density (a, b), free energy and SC order parameter (c, d) as a function of momentum QQ. Parameters for (a) and (c): t2=0t_{2}=0 and t1=1t_{1}=1, α=tan(π20)\alpha=\tan(\frac{\pi}{20}), μ=0\mu=0, V1=2V_{1}=2. Parameters for (b) and (d): t2=0.5t_{2}=-0.5 and t1=1t_{1}=1, α=0.4\alpha=0.4, μ=0.53\mu=-0.53, V1=2V_{1}=2. The spin current is defined as js=12(jj)j_{s}=\frac{1}{2}(j_{\uparrow}-j_{\downarrow}).

V Josephson junctions

We next study the transport properties of Josephson chains consisting of a normal metal sandwiched between two KFF superconductors depicted in Fig. 3(a) with more detailed setups shown in Appendix. D. We consider the general case with nonzero t2t_{2} in the KFF state.

V.1 Spin Diode effect with spin-independent phase bias

We first study the transport properties of JJ with spin-independent phase bias ϕ\phi. The Josephson currents can be calculated by the formula

I(ϕ)=2eΩJc(ϕ)ϕI(\phi)=\frac{2e}{\hbar}\frac{\partial\Omega_{Jc}(\phi)}{\partial\phi} (22)

where ΩJc(ϕ)\Omega_{Jc}(\phi) is the free energy of the system [52]. The results for both spin components and total charge and spin currents are shown in Fig. 3(b). Due to 𝒯{\cal T} symmetry, the spin-dependent critical currents satisfy Iσ,c+=Iσ¯,cI_{\sigma,c+}=-I_{\bar{\sigma},c-} such that charge current is reciprocal |Ie,c+|=|Ie,c||I_{e,c+}|=|I_{e,c-}|. In contrast, the critical current for each spin is asymmetric, i.e. |Iσ,c+||Iσ,c||I_{\sigma,c+}|\neq|I_{\sigma,c-}|, giving rise to nonreciprocal Josephson spin currents |Is,c+||Is,c||I_{s,c+}|\neq|I_{s,c-}| and the SC spin diode effect.

Refer to caption
Figure 3: (a) Schematics of an S-N-S Josephson junction. The lengths of the KFF-SC (S) and normal metal (N) regions are NsN_{s} and NnN_{n}. (b) Current phase relation for the Josephson chain with a typical KFF-SC order determined self-consistently for parameters: t1=1t_{1}=1, t2=0.5t_{2}=-0.5, α=0.5\alpha=0.5, μ=0.7\mu=-0.7, V1=1V_{1}=1, leading to Δ=0.063\Delta_{\parallel}=0.063 and Q=0.372πQ=0.372\pi. Upper and lower horizontal dashed lines indicate critical current I,c+I_{\uparrow,c+} and I,cI_{\uparrow,c-} along “±\pm” directions. (c) Energy-phase spectrum of the Josephson chain in (b). Two vertical dashed lines in (b) and (c) correspond to phase bias ϕ0\phi_{0\uparrow} and ϕ0\phi_{0\downarrow} where the bound states cross zero, leading to jumps in the Josephson current shown in (b). Ns=319N_{s}=319 and Nn=3N_{n}=3. The “N” region has nearest neighbor hopping tN=1t_{N}=1. The couplings between “N” and “S” regions are described by tL=tR=1t_{L}=t_{R}=1 as defined in Appendix. D.

V.2 Spin diode effect in the Josephson junction with spin phase

The spin diode effect with nonreciprocal spin current by applying a spin-dependent phase bias known as the spin phase [53] is obvious. The spin phase was first introduced in Ref. 53. In the current case, we consider the spin phase in the zz direction, which corresponds to the spin-dependent phase bias ϕσ=σϕ\phi_{\sigma}=\sigma\phi applied to the Josephson junction. We consider the same parameter set as that shown in Fig. 3 and the resulting current phase relation is shown in Fig. 4. Here, since the spin phase ϕσ=σϕ\phi_{\sigma}=\sigma\phi still respects the 𝒯\cal{T} symmetry, the current for the opposite spin polarization are always equal in magnitude and opposite in the direction, i.e., I(ϕ)=I(ϕ)I_{\uparrow}(\phi)=-I_{\downarrow}(\phi), so that the total charge current Ie(ϕ)I_{e}(\phi) always vanishes and the total spin current Is(ϕ)=12(I(ϕ)I(ϕ))=I(ϕ)I_{s}(\phi)=\frac{1}{2}(I_{\uparrow}(\phi)-I_{\downarrow}(\phi))=I_{\uparrow}(\phi), which is nonreciprocal as long as the hidden inversion symmetry is broken by finite t2t_{2}. The nonreciprocal spin current of the Josephson junction with spin phase is inherited from the nonreciprocal spin transport of the bulk KFF state.

Refer to caption
Figure 4: (a) Schematics of a S-N-S Josephson junction with a spin phase in zz direction. The lengths of the KFF-SC (S) and normal metal (N) regions are NsN_{s} and NnN_{n}. (b) Current phase relation for the Josephson chain with a typical KFF-SC order determined self-consistently for parameters: t1=1t_{1}=1, t2=0.5t_{2}=-0.5, α=0.5\alpha=0.5, μ=0.7\mu=-0.7, V1=1V_{1}=1, leading to Δ=0.063\Delta_{\parallel}=0.063 and Q=0.372πQ=0.372\pi. Upper and lower horizontal dashed lines indicate critical current I,c+I_{\uparrow,c+} and I,cI_{\uparrow,c-} along “±\pm” directions. Ns=319N_{s}=319 and Nn=3N_{n}=3 are used in the calculations, and the “N” region has nearest neighbor hopping tN=1t_{N}=1. The coupling between the “N” and “S” regions are described by tL=tR=1t_{L}=t_{R}=1 as defined in Appendix. D.

VI Length controlled phase transitions in charge transport

An intriguing feature in Fig. 3(b) is the phase difference between Josephson currents in two spin sectors, which is clearly revealed in the energy spectrum plotted in Fig. 3(c). Apart from the continuum states outside the SC gap, there are eight in-gap states. Among them, four are at exactly zero energy, corresponding to two pairs of Majorana zero modes (one for each spin) located at the two ends of the chain due to pp-wave nature of the KFF state [51], which do not contribute to Josephson current. The other four in-gap states are Andreev bound states of the S-N-S junction. For junctions made of conventional superconductors, the energies of the bound states cross at ϕ=π\phi=\pi [54], corresponding to zero modes trapped by π\pi-junction [55]. In the current KFF JJs, the energy spectrum of the bound states for two spin species shifts in opposite directions as shown in Fig. 3(c). As a result, the phase bias where the bound states cross zero shifts from ±π\pm\pi to ±ϕ0σ\pm\phi_{0\sigma}, where the Josephson current jumps due to branch switching as shown in Fig. 3(b).

Such phase differences have a great impact on charge transport. The charge current Ie=I+II_{e}=I_{\uparrow}+I_{\downarrow} crosses zero at both ϕ=0\phi=0 and ϕ=π\phi=\pi with positive slopes, indicating the free energy ΩJc(ϕ)\Omega_{Jc}(\phi) of the junction reaches a local minimum at both ϕ=0\phi=0 and π\pi. This state is called a 𝟎\mathbf{0}^{\prime} or 𝝅\pmb{\pi}^{\prime} state depending on the momentum of the global minimum. It was studied previously in JJs where two superconductors are coupled through an Anderson impurity [56] or a magnetic quantum dot [57]. Here, these remarkable states are realized in 𝒯{\cal T}-invariant systems due to the novel KFF SC order. Remarkably, we find that distinct Josephson junction states (𝟎\mathbf{0}, 𝝅\pmb{\pi}, 𝟎\mathbf{0}^{\prime} and 𝝅\pmb{\pi}^{\prime}) can all be realized by tuning the phase shift, which can be easily achieve by changing the length of the SC region. The definition of these states are listed in Table. 1.

The dependence of the phase shift on the length NsN_{s} of the SC region can be understood by performing a local gauge transformation that maps the KFF JJ onto a Kitaev JJ consisting of two spin-degenerate pp-wave Kitaev chains subject to spin-dependent phase bias ϕσ=ϕ+σQ(Ns1)\phi_{\sigma}=\phi+\sigma Q(N_{s}-1) as shown in Appendix. D. If we further assume the Josephson current for the Josephson chain consisting of two spin degenerate pp-wave Kitaev chains with phase bias ϕ\phi as I0(ϕ)I_{0}(\phi) which is identical for the two spin species due to the spin degeneracy, we can then immediately get the Josephson current for each spin species as Iσ(ϕ)=I0(ϕ+σQ(Ns1))I_{\sigma}(\phi)=I_{0}(\phi+\sigma Q(N_{s}-1)), i.e., the Josephson current Iσ(ϕ)I_{\sigma}(\phi) is shifted from the current of the transformed junction I0(ϕ)I_{0}(\phi) by a phase σQ(Ns1)\sigma Q(N_{s}-1) (mod 2π2\pi) so that the relative phase difference of the current between the two spin species is δϕ=2σQ(Ns1)\delta\phi=2\sigma Q(N_{s}-1) (mod 2π2\pi). Various Josephson junction states can be realized by tuning δϕ\delta\phi from 0 to 2π2\pi through the length NsN_{s}. This relation is verified numerically in Figs. 5(a-e) for Q=π10Q=\frac{\pi}{10} and Ns[321,332]N_{s}\in[321,332]. Any combination of QQ and NsN_{s} can produce similar results as long as δϕ\delta\phi covers the range [0,2π][0,2\pi]. We also calculate the total free energy of the system for different NsN_{s} and indeed observe transitions between all these states controlled by NsN_{s} as shown in Fig. 5(f). Specifically, for NsN_{s}=321 (331), there is only one global minimum located at ϕ=0(π)\phi=0\ (\pi) and the system is in 𝟎(𝝅)\mathbf{0}\ (\pmb{\pi}) state. For NsN_{s}=323 (329), the free energy has a global minimum at ϕ=0(π)\phi=0\ (\pi) and a local minimum at ϕ=π(0)\phi=\pi\ (0) so that the system is in 𝟎\mathbf{0}^{\prime} (𝝅\pmb{\pi}^{\prime}) state. The transition between the two states is reached at Ns=326N_{s}=326, where δϕ=π\delta\phi=\pi and ΩJc(0)=ΩJc(π)\Omega_{Jc}(0)=\Omega_{Jc}(\pi).

Interestingly, in this case with only nearest neighbor hopping t1t_{1}, the charge current Ie(ϕ)I_{e}(\phi) acquires a period of π\pi in phase ϕ\phi instead of 2π2\pi as in conventional Josephson current at the critical point (Ns=326N_{s}=326), which can also be understood in the gauge transformed basis. In this case with Ns=326N_{s}=326, the phase bias becomes ϕσ=ϕ+σπ2\phi_{\sigma}=\phi+\frac{\sigma\pi}{2} and we thus have Iσ(ϕ)=I0(ϕ+σπ2)I_{\sigma}(\phi)=I_{0}(\phi+\frac{\sigma\pi}{2}), from which we can get I(ϕ+π)=I0(ϕ+3π2)=I0(ϕπ2)=I(ϕ)I_{\uparrow}(\phi+\pi)=I_{0}(\phi+\frac{3\pi}{2})=I_{0}(\phi-\frac{\pi}{2})=I_{\downarrow}(\phi). Then we can derive a new relation for the total charge current IeI_{e} as

Ie(ϕ+π)=I(ϕ+π)+I(ϕ+π)=I(ϕ)+I(ϕ)=Ie(ϕ)I_{e}(\phi+\pi)=I_{\uparrow}(\phi+\pi)+I_{\downarrow}(\phi+\pi)=I_{\downarrow}(\phi)+I_{\uparrow}(\phi)=I_{e}({\phi})

so that the charge current Ie(ϕ)I_{e}(\phi) acquires a period of π\pi instead of 2π2\pi. The spin current Is(ϕ)I_{s}(\phi) then acquires a minus sign when progressing π\pi phase, so that its period is still 2π2\pi.

Is(ϕ+π)=I(ϕ+π)I(ϕ+π)=I(ϕ)I(ϕ)=Is(ϕ)I_{s}(\phi+\pi)=I_{\uparrow}(\phi+\pi)-I_{\downarrow}(\phi+\pi)=I_{\downarrow}(\phi)-I_{\uparrow}(\phi)=-I_{s}({\phi})

The realization of diverse Josephson junction states by simply controlling the length of the superconductor is an intriguing property. It originates from the opposite nonzero momentum of Cooper pairs in each spin sector in the novel KFF state. The quantum interference of relative phase shifted pairing functions with opposite spin polarization leads to rich phases and physical phenomena.

State label Distribution of minimums of the free energy
𝟎\mathbf{0} global minimum at ϕ=0\phi=0
𝟎\mathbf{0^{\prime}} global minimum at ϕ=0\phi=0 and local minimum at ϕ=π\phi=\pi
𝟎𝝅\mathbf{0^{\prime}}-\pmb{\pi}^{\prime} global minimum at both ϕ=0\phi=0 and ϕ=π\phi=\pi
𝝅\pmb{\pi}^{\prime} global minimum at ϕ=π\phi=\pi and local minimum at ϕ=0\phi=0
𝝅\pmb{\pi} global minimum at ϕ=π\phi=\pi
Table 1: Definitions of the various Josephson junction states via the distributions of the minimums of the free energy.
Refer to caption
Figure 5: (a-e) Josephson current for different values of NsN_{s}. The parameters are t1=1t_{1}=1, t2=0t_{2}=0, α=tan(π20)\alpha=\tan(\frac{\pi}{20}), μ=0\mu=0, V1=2V_{1}=2 in the SC region. The KFF state has Q=π10Q=\frac{\pi}{10} and Δ=0.169\Delta_{\parallel}=0.169. The normal region has tN=tL=tR=1t_{N}=t_{L}=t_{R}=1 as defined in Appendix. D. (f) The total free energy at the corresponding values of NsN_{s} in (a-e), showing various junction states labelled by 𝟎\mathbf{0}, 𝟎\mathbf{0^{\prime}}, 𝟎𝝅\mathbf{0^{\prime}}-\pmb{\pi}^{\prime}, 𝝅\pmb{\pi}^{\prime} and 𝝅\pmb{\pi} determined by the distribution of the (local) minima at ϕ=0\phi=0 and π\pi. The charge current Ie(ϕ)I_{e}(\phi) in (a-e) vanishes at both ϕ=0\phi=0 and π\pi, the minima of the free energy in (f).

VII Discussion

We reported the theoretical discovery of a novel time-reversal invariant, finite momentum pairing Fulde-Ferrell state – the KFF state. The concrete effective 1D model we used to realize the KFF state and its many unprecedented properties is intimately connected to the novel physics observed at the atomic line defect (ALD) in monolayer iron-based superconductor Fe(Te,Se), where zero-energy bound states emerge at both ends of the ALD with no signatures of 𝒯{\cal T} symmetry breaking [50]. The missing atoms cause inversion symmetry breaking and induces Rashba SOC. It was shown that significant equal-spin triplet pairing can be induced by coherent quantum mechanical processes along such a Rashba ALD embedded in 2D unconventional superconductors  [49]. This makes it plausible for materializing the effective 1D model with significant equal-spin triplet pairing to generate the KFF state. More recently, evidence for finite momentum pair density wave order has been observed in monolayer Fe(Te,Se) along one-dimensional domain walls [58]. The experimental evidence suggests time-reversal symmetry is preserved, which makes the KFF state a plausible candidate in addition to the Larkin-Ovchinnikov state.

The most remarkable of the KFF state is that, in the presence of broken inversion symmetry, it supports nonreciprocal spin supercurrent in both bulk superconductor and JJ. In contrast to nonreciprocal charge transport in SC systems which requires breaking both inversion and 𝒯{\cal T} symmetry, nonreciprocal SC spin transport only requires breaking inversion symmetry. This is because the spin current operator is invariant under time-reversal, such that systems with positive and negative spin current are unrelated by the 𝒯{\cal T} operation. This is true regardless of whether the system respects the 𝒯{\cal T} symmetry or not, making it free of the constraint by the Onsager relation. We thus propose a novel SC spin diode effect as a potential new frontier for using spins to make dissipationless electronic devices in SC spintronics. While future work is clearly needed which is outside the scope of the current paper, we point out that the unique properties of the KFF state make it plausible for possible realizations in JJs consisting of a ferromagnetic barrier. The exchange field of the barrier favors spin-triplet pairing and has very little effect on the critical current of equal-spin triplet pairing such as in the KFF state, resulting in the slow decay of the critical current with increasing barrier length. Both effects have been demonstrated experimentally [59, 60]. In turn, detecting the nonreciprocal spin transport together with the slow decay of the critical current with the barrier length can serve as the smoking gun evidence for the KFF state.

Acknowledgements.
We thank Kun Jiang for helpful discussions. YZ is supported in part by National Natural Science Foundation of China (NSFC) Grants No. 12004383, No. 12074276 and No 12274279. ZW is supported by the U.S. Department of Energy, Basic Energy Sciences, Grant No. DE FG02-99ER45747.

Appendix A More detailed results from the meanfield calculation

We perform the calculation with two sets of parameters, which are shown in Fig. 6. As shown in Fig. 6, the ground states with finite pairing are either the KFF states with order parameters solely condensed in the equal-spin pairing channel or the mixture of ss and pzp_{z} wave pairing states whose order parameters are solely condensed in the opposite-spin pairing channel and no mixed states with the coexistence of the order in both channels are found as the ground states except along the phase boundary of the two states where these two states are degenerate. This means that we can consider the two pairing channels separately, which further simplifies the meanfield Hamiltonian and is helpful for us in studying the properties of each state.

Refer to caption
Figure 6: (a,b) Phase diagram determined from the meanfield calculations with two sets of parameters. (c,d) The superconducting order parameter for both channels as the function of the attraction V showing the Cooper instability for the two sets of paramters. The parameters used are t1=1t_{1}=1, t2=0t_{2}=0, α=tan(π20)\alpha=\tan(\frac{\pi}{20}), μ=0\mu=0 for (a,c) and t1=1t_{1}=1, t2=0.5t_{2}=-0.5, α=0.4\alpha=0.4, μ=0.53\mu=-0.53 for (b,d).

Appendix B Derivation of the spin current for the KFF state

As shown in the main text, the meanfield Hamiltonian purely in the equal-spin pairing channel can be written as

H^MFμN^=kσ(εkσμ)ckσckσ+V1Δkσisinkck+σQ2,σck+σQ2,σ+h.c.+2NcV1Δ2\begin{split}\hat{H}_{MF}-\mu\hat{N}&=\sum_{k\sigma}\left(\varepsilon_{k\sigma}-\mu\right)c_{k\sigma}^{\dagger}c_{k\sigma}\\ &+V_{1}\Delta_{\parallel}\sum_{k\sigma}i\sin kc_{k+\frac{\sigma Q}{2},\sigma}^{\dagger}c_{-k+\frac{\sigma Q}{2},\sigma}^{\dagger}+h.c.+2N_{c}V_{1}\Delta_{\parallel}^{2}\end{split} (23)

which can be further simplified in the Nambu basis ψkσ=(ck+σQ2,σ,ck+σQ2,σ)\psi_{k\sigma}^{\dagger}=\left(c_{k+\frac{\sigma Q}{2},\sigma}^{\dagger},c_{-k+\frac{\sigma Q}{2},\sigma}\right) as

H^MFμN^=12kσψkσhk,Q,σψkσ+2NcV1Δ2μNc\hat{H}_{MF}-\mu\hat{N}=\frac{1}{2}\sum_{k\sigma}\psi_{k\sigma}^{\dagger}h_{k,Q,\sigma}\psi_{k\sigma}+2N_{c}V_{1}\Delta_{\parallel}^{2}-\mu N_{c} (24)

with

hk,Q,σ=[εk+σQ2,σμ2iV1Δsink2iV1Δsinkεk+σQ2,σ+μ]h_{k,Q,\sigma}=\left[\begin{array}[]{cc}\varepsilon_{k+\frac{\sigma Q}{2},\sigma}-\mu&2iV_{1}\Delta_{\parallel}\sin k\\ -2iV_{1}\Delta_{\parallel}\sin k&-\varepsilon_{-k+\frac{\sigma Q}{2},\sigma}+\mu\end{array}\right] (25)

In order to derive the expression for the spin current, let’s consider the free energy density of the KFF state at finite temperature T, which is given as

Ω(Δ,Q,T)=TNclnTr[eH^MFμN^T]=T2Nckσtr[ln(1+ehk,Q,σT)]+2V1Δ2μ=σΩσ(Δ,Q,T)μ\begin{split}\Omega(\Delta_{\parallel},Q,T)&=-\frac{T}{N_{c}}\ln\text{Tr}\left[e^{-\frac{\hat{H}_{MF}-\mu\hat{N}}{T}}\right]\\ &=-\frac{T}{2N_{c}}\sum_{k\sigma}\text{tr}\left[\ln(1+e^{-\frac{h_{k,Q,\sigma}}{T}})\right]+2V_{1}\Delta_{\parallel}^{2}-\mu\\ &=\sum_{\sigma}\Omega_{\sigma}(\Delta_{\parallel},Q,T)-\mu\end{split} (26)

with

Ωσ(Δ,Q,T)=T2Nctr[ln(1+ehk,Q,σT)]+V1Δ2\Omega_{\sigma}(\Delta_{\parallel},Q,T)=-\frac{T}{2N_{c}}\text{tr}\left[\ln(1+e^{-\frac{h_{k,Q,\sigma}}{T}})\right]+V_{1}\Delta_{\parallel}^{2} (27)

The current operator for each spin species in the system studied is defined as

j^σ=1Nckkεkσckσckσ\hat{j}_{\sigma}=\frac{1}{N_{c}\hbar}\sum_{k}\partial_{k}\varepsilon_{k\sigma}c_{k\sigma}^{\dagger}c_{k\sigma} (28)

Next, from Eq. 25 we have

Qhk,Q,σ=[σ2kεk+σQ2,σ00σ2kεk+σQ2,σ]\partial_{Q}h_{k,Q,\sigma}=\left[\begin{array}[]{cc}\frac{\sigma}{2}\partial_{k}\varepsilon_{k+\frac{\sigma Q}{2},\sigma}&0\\ 0&-\frac{\sigma}{2}\partial_{k}\varepsilon_{-k+\frac{\sigma Q}{2},\sigma}\end{array}\right] (29)

and then we further have

1NckψkσQhk,Q,σψkσ=1Nck[σ2kεk+σQ2,σck+σQ2,σck+σQ2,σσ2kεk+σQ2,σ(1ck+σQ2,σck+σQ2,σ)]=1Nckσkεkσckσckσ=σj^σ\begin{split}&\frac{1}{N_{c}\hbar}\sum_{k}\psi_{k\sigma}^{\dagger}\partial_{Q}h_{k,Q,\sigma}\psi_{k\sigma}=\frac{1}{N_{c}\hbar}\sum_{k}\Bigl{[}\frac{\sigma}{2}\partial_{k}\varepsilon_{k+\frac{\sigma Q}{2},\sigma}c_{k+\frac{\sigma Q}{2},\sigma}^{\dagger}c_{k+\frac{\sigma Q}{2},\sigma}\\ &-\frac{\sigma}{2}\partial_{k}\varepsilon_{-k+\frac{\sigma Q}{2},\sigma}(1-c_{-k+\frac{\sigma Q}{2},\sigma}^{\dagger}c_{-k+\frac{\sigma Q}{2},\sigma})\Bigl{]}\\ &=\frac{1}{N_{c}\hbar}\sum_{k}\sigma\partial_{k}\varepsilon_{k\sigma}c_{k\sigma}^{\dagger}c_{k\sigma}=\sigma\hat{j}_{\sigma}\end{split} (30)

which means

j^σ=σNckψkσQhk,Q,σψkσ\hat{j}_{\sigma}=\frac{\sigma}{N_{c}\hbar}\sum_{k}\psi_{k\sigma}^{\dagger}\partial_{Q}h_{k,Q,\sigma}\psi_{k\sigma} (31)

Then the current for each spin species can be calculated as

jσ(Δ,Q,T)=Tr[j^σeH^MFμN^T]Tr[eH^MFμN^T]=σNctr[Qhk,Q,σf(hk,Q,σ)]=2σQΩσ(Δ,Q,T)\begin{split}j_{\sigma}(\Delta_{\parallel},Q,T)&=\frac{\text{Tr}\left[\hat{j}_{\sigma}e^{-\frac{\hat{H}_{MF}-\mu\hat{N}}{T}}\right]}{\text{Tr}\left[e^{-\frac{\hat{H}_{MF}-\mu\hat{N}}{T}}\right]}=\frac{\sigma}{N_{c}\hbar}\text{tr}\left[\partial_{Q}h_{k,Q,\sigma}f(h_{k,Q,\sigma})\right]\\ &=\frac{2\sigma}{\hbar}\partial_{Q}\Omega_{\sigma}(\Delta_{\parallel},Q,T)\end{split} (32)

where f(x)=(1+ex/T)1f(x)=(1+e^{x/T})^{-1} is the Fermi distribution function. Here, the operator j^σ\hat{j}_{\sigma} is the density current, from which we can define the charge current operator as j^e=e(j^+j^)\hat{j}_{e}=e(\hat{j}_{\uparrow}+\hat{j}_{\downarrow}) and the operator for the spin current carrying spin polarization in zz direction as j^s=2(j^j^)\hat{j}_{s}=\frac{\hbar}{2}(\hat{j}_{\uparrow}-\hat{j}_{\downarrow}). Therefore, we can conveniently define the unit for the charge and spin current as e\frac{e}{\hbar} and 1, so that both currents can be written in a similar format. We finally arrive at the expression for the spin current with spin up polarization as

js(Q)=σσ2jσ=QΩ(Δ,Q,T)j_{s}(Q)=\sum_{\sigma}\frac{\sigma\hbar}{2}j_{\sigma}=\partial_{Q}\Omega(\Delta_{\parallel},Q,T) (33)

This means the ground state with the optimized value of Q=Q0Q=Q_{0} does not carry any net spin current as expected, since QΩ(Δ,Q0,T)=0\partial_{Q}\Omega(\Delta_{\parallel},Q_{0},T)=0, and when a spin current jsj_{s} is applied to the KFF state, the state with a different value of QQ, satisfying js(Q)=jsj_{s}(Q)=j_{s} is realized. Then the critical spin current for the positive and negative directions are determined by the maximum and minimum values of js(Q)j_{s}(Q) sustained by the superconducting state, which can be defined as js,c+=maxQ[js(Q)]j_{s,c+}=\max_{Q}[j_{s}(Q)] and js,c=minQ[js(Q)]j_{s,c-}=min_{Q}[j_{s}(Q)]. Moreover, due to the presence of 𝒯{\cal T} symmetry, Ω(Δ,Q,T)=Ω(Δ,Q,T)\Omega_{\uparrow}(\Delta_{\parallel},Q,T)=\Omega_{\downarrow}(\Delta_{\parallel},Q,T), which means j(Δ,Q,T)=j(Δ,Q,T)j_{\uparrow}(\Delta_{\parallel},Q,T)=-j_{\downarrow}(\Delta_{\parallel},Q,T), such that the charge current jc(Q)=σjσj_{c}(Q)=\sum_{\sigma}j_{\sigma} always vanishes in the KFF state as expected and js(Q)=σσ2jσ=jj_{s}(Q)=\sum_{\sigma}\frac{\sigma}{2}j_{\sigma}=j_{\uparrow}.

Appendix C Determining the optimized Q for KFF state

We first consider a simpler case with only nearest neighbor hopping, i.e. t2=0t_{2}=0. In this case, the dispersion of the noninteracting Hamiltonian becomes εkσ=2t1cosk2σαsink=2tαcos(kσθα)\varepsilon_{k\sigma}=-2t_{1}\cos k-2\sigma\alpha\sin k=-2t_{\alpha}\cos(k-\sigma\theta_{\alpha}) with tα=t12+α2t_{\alpha}=\sqrt{t_{1}^{2}+\alpha^{2}} and θα=arctan(αt1)\theta_{\alpha}=\arctan(\frac{\alpha}{t_{1}}), which determines the Fermi momentum as

kfσ,±=σθα±arccos(μ2tα)k_{f\sigma,\pm}=\sigma\theta_{\alpha}\pm\arccos(-\frac{\mu}{2t_{\alpha}}) (34)

Then the quasiparticle energy Eq. 16 becomes

Ekσ,±=2σtαsin(θαQ2)sink±[2tαcos(θαQ2)coskμ]2+4V12Δ2sin2k\begin{split}E_{k\sigma,\pm}&=-2\sigma t_{\alpha}\sin(\theta_{\alpha}-\frac{Q}{2})\sin k\\ &\pm\sqrt{\left[-2t_{\alpha}\cos(\theta_{\alpha}-\frac{Q}{2})\cos k-\mu\right]^{2}+4V_{1}^{2}\Delta_{\parallel}^{2}\sin^{2}k}\end{split} (35)

and from Eq. 17 and the 𝒯{\cal T} symmetry, we have the free energy density at zero temperature as

Ω(Δ,Q)=12Nckσ,n=±Ekσ,nΘ(Ekσ,n)+2V1Δ2μ=1Nck,n=±Ek,nΘ(Ek,n)+2V1Δ2μ\Omega(\Delta_{\parallel},Q)=\frac{1}{2N_{c}}\sum_{k\sigma,n=\pm}E_{k\sigma,n}\Theta(-E_{k\sigma,n})+2V_{1}\Delta_{\parallel}^{2}-\mu=\frac{1}{N_{c}}\sum_{k,n=\pm}E_{k\uparrow,n}\Theta(-E_{k\uparrow,n})+2V_{1}\Delta_{\parallel}^{2}-\mu (36)

which leads to the expression for the spin current as

js(Q)=QΩ(Δ,Q)=1Nck,n=±Θ(Ek,n){tαcos(θαQ2)sink+[tαsin(θαQ2)cosk][2tαcos(θαQ2)cosk]Ek,n+2tαsin(θαQ2)sink}j_{s}(Q)=\partial_{Q}\Omega(\Delta_{\parallel},Q)=\frac{1}{N_{c}}\sum_{k,n=\pm}\Theta(-E_{k\uparrow,n})\left\{t_{\alpha}\cos(\theta_{\alpha}-\frac{Q}{2})\sin k+\frac{\left[t_{\alpha}\sin(\theta_{\alpha}-\frac{Q}{2})\cos k\right]\left[2t_{\alpha}\cos(\theta_{\alpha}-\frac{Q}{2})\cos k\right]}{E_{k\uparrow,n}+2t_{\alpha}\sin(\theta_{\alpha}-\frac{Q}{2})\sin k}\right\} (37)

We can easily see that when Q=Q0=2θαQ=Q_{0}=2\theta_{\alpha}, Ekσ,±=±[2tαcoskμ]2+4V12Δ2sin2k=±EkE_{k\sigma,\pm}=\pm\sqrt{\left[-2t_{\alpha}\cos k-\mu\right]^{2}+4V_{1}^{2}\Delta_{\parallel}^{2}\sin^{2}k}=\pm E_{k}, then we have js(Q0)=1Nck,n=±Θ(nEk)tαsink=2tαNcksink=0j_{s}(Q_{0})=\frac{1}{N_{c}}\sum_{k,n=\pm}\Theta(-nE_{k})t_{\alpha}\sin k=\frac{2t_{\alpha}}{N_{c}}\sum_{k}\sin k=0. Therefore, the ground state is characterized by the momentum Q0=2θα=σ(kfσ,++kfσ,)Q_{0}=2\theta_{\alpha}=\sigma(k_{f\sigma,+}+k_{f\sigma,-}), which is also consistent with the Fermi surface of the noninteracting Hamiltonian, where the electrons around the two Fermi points kfσ,±k_{f\sigma,\pm} within the same spin species pair together, giving rise to a finite Cooper pair momentum Q0Q_{0}. We numerically calculate the zero temperature free energy as well as the current density as shown in Fig. 2(a, c) of the main text, which indeed confirms that the free energy density reaches the minimum at Q=Q0=2θαQ=Q_{0}=2\theta_{\alpha} and the critical spin currents js,c±j_{s,c\pm} are determined by the maximum and minimum values of the spin current density.

Appendix D The relation between the phase difference and NsN_{s}

To study the transport properties of the Josephson chain, we consider the system consisting of two superconductors with KFF order sandwiching a normal metal in between as shown in Fig. 3(a) of the main text. By setting the lattice constant to 1, this Josephson chain can be described by the tight-binding Hamiltonian HJc(ϕ)=HSL+HN+HSR+HLN+HRNH_{Jc}(\phi)=H_{SL}+H_{N}+H_{SR}+H_{LN}+H_{RN}, where

HSL=i,j[1,Ns]σ(tijδijμ)ciσcjσ+i=1Ns1σ(iασciσci+1σ+ΔeiσQxi+iϕciσci+1σ)+h.c.H_{SL}=-\sum_{i,j\in[1,N_{s}]}\sum_{\sigma}(t_{ij}-\delta_{ij}\mu)c_{i\sigma}^{\dagger}c_{j\sigma}+\sum_{i=1}^{N_{s}-1}\sum_{\sigma}\left(i\alpha\sigma c_{i\sigma}^{\dagger}c_{i+1\sigma}+\Delta_{\parallel}e^{i\sigma Qx_{i}+i\phi}c_{i\sigma}^{\dagger}c_{i+1\sigma}^{\dagger}\right)+h.c. (38)
HSR=i,j[Ns+Nn+1,2Ns+Nn]σ(tijδijμ)ciσcjσ+i=Ns+Nn+12Ns+Nn1σ(iασciσci+1σ+ΔeiσQ(xiNsNn)ciσci+1σ)+h.c.H_{SR}=-\sum_{i,j\in[N_{s}+N_{n}+1,2N_{s}+N_{n}]}\sum_{\sigma}(t_{ij}-\delta_{ij}\mu)c_{i\sigma}^{\dagger}c_{j\sigma}+\sum_{i=N_{s}+N_{n}+1}^{2N_{s}+N_{n}-1}\sum_{\sigma}\left(i\alpha\sigma c_{i\sigma}^{\dagger}c_{i+1\sigma}+\Delta_{\parallel}e^{i\sigma Q(x_{i}-N_{s}-N_{n})}c_{i\sigma}^{\dagger}c_{i+1\sigma}^{\dagger}\right)+h.c. (39)

describe the two SC regions on the left and right sides,

HN=i,j[Ns+1,Ns+Nn]σ(tN,ijδijμ)ciσcjσH_{N}=-\sum_{i,j\in[N_{s}+1,N_{s}+N_{n}]}\sum_{\sigma}(t_{N,ij}-\delta_{ij}\mu)c_{i\sigma}^{\dagger}c_{j\sigma} (40)

describes the normal metal region in the middle and

HSNL=tLσcNsσcNs+1σ+h.c.H_{SNL}=-t_{L}\sum_{\sigma}c_{N_{s}\sigma}^{\dagger}c_{N_{s}+1\sigma}+h.c. (41)
HSNR=tRσcNs+NnσcNs+Nn+1σ+h.c.H_{SNR}=-t_{R}\sum_{\sigma}c_{N_{s}+N_{n}\sigma}^{\dagger}c_{N_{s}+N_{n}+1\sigma}+h.c. (42)

correspond to the coupling between the normal metal region and the left and right SC region, with ϕ\phi the phase bias between the two SCs. Then the Josephson currents can be calculated by the formula

I(ϕ)=2eϕnf(ϵn)ϵn(ϕ)I(\phi)=\frac{2e}{\hbar}\partial_{\phi}\sum_{n}f(\epsilon_{n})\epsilon_{n}(\phi) (43)

with ϵn\epsilon_{n} the n-th eigenvalue for HJcH_{Jc} at the phase bias ϕ\phi and f(ϵ)f(\epsilon) the Fermi distribution function.

Next, we demonstrate the dependence of the relative phase difference on the length of the superconducting region NsN_{s}. Let us consider a simpler case with only nearest neighbor hopping on the superconducting region. We first perform a local gauge transformation

{ciσeiσ2Q(xi12)diσfor i[1,Ns]ciσeiσ2Q(Ns12)diσfor i[Ns+1,Ns+Nn]ciσeiσ2Q(xiNn32)diσfor i[Ns+Nn+1,2Ns+Nn]\begin{cases}c_{i\sigma}^{\dagger}\rightarrow e^{-\frac{i\sigma}{2}Q(x_{i}-\frac{1}{2})}d_{i\sigma}^{\dagger}&\text{\text{for\ i$\in[1,N_{s}]$}}\\ c_{i\sigma}^{\dagger}\rightarrow e^{-\frac{i\sigma}{2}Q(N_{s}-\frac{1}{2})}d_{i\sigma}^{\dagger}&\text{for\ i$\in[N_{s}+1,N_{s}+N_{n}]$}\\ c_{i\sigma}^{\dagger}\rightarrow e^{-\frac{i\sigma}{2}Q(x_{i}-N_{n}-\frac{3}{2})}d_{i\sigma}^{\dagger}&\text{for\ i$\in[N_{s}+N_{n}+1,2N_{s}+N_{n}]$}\end{cases} (44)

then each term in Hamiltonian HJc(ϕ)H_{Jc}(\phi) becomes

HSL=i=1Ns1σ[tαeiσ(Q2θα)diσdi+1σ+Δeiϕdiσdi+1σ]+h.c.μi=1NsσdiσdiσH_{SL}=\sum_{i=1}^{N_{s}-1}\sum_{\sigma}\left[-t_{\alpha}e^{i\sigma(\frac{Q}{2}-\theta_{\alpha})}d_{i\sigma}^{\dagger}d_{i+1\sigma}+\Delta_{\parallel}e^{i\phi}d_{i\sigma}^{\dagger}d_{i+1\sigma}^{\dagger}\right]+h.c.-\mu\sum_{i=1}^{N_{s}}\sum_{\sigma}d_{i\sigma}^{\dagger}d_{i\sigma} (45)
HSR=i=Ns+Nn+12Ns+Nn1σ[tαeiσ(Q2θα)diσdi+1σ+ΔeiσQ(Ns1)diσdi+1σ]+h.c.μi=Ns+Nn+12Ns+NnσdiσdiσH_{SR}=\sum_{i=N_{s}+N_{n}+1}^{2N_{s}+N_{n}-1}\sum_{\sigma}\left[-t_{\alpha}e^{i\sigma(\frac{Q}{2}-\theta_{\alpha})}d_{i\sigma}^{\dagger}d_{i+1\sigma}+\Delta_{\parallel}e^{-i\sigma Q(N_{s}-1)}d_{i\sigma}^{\dagger}d_{i+1\sigma}^{\dagger}\right]+h.c.-\mu\sum_{i=N_{s}+N_{n}+1}^{2N_{s}+N_{n}}\sum_{\sigma}d_{i\sigma}^{\dagger}d_{i\sigma} (46)
HN=i,j[Ns+1,Ns+Nn]σ(tN,ijδijμ)diσdjσH_{N}=-\sum_{i,j\in[N_{s}+1,N_{s}+N_{n}]}\sum_{\sigma}(t_{N,ij}-\delta_{ij}\mu)d_{i\sigma}^{\dagger}d_{j\sigma} (47)
HSNL=tLσdNsσdNs+1σ+h.c.H_{SNL}=-t_{L}\sum_{\sigma}d_{N_{s}\sigma}^{\dagger}d_{N_{s}+1\sigma}+h.c. (48)
HSNR=tRσdNs+NnσdNs+Nn+1σ+h.c.H_{SNR}=-t_{R}\sum_{\sigma}d_{N_{s}+N_{n}\sigma}^{\dagger}d_{N_{s}+N_{n}+1\sigma}+h.c. (49)

We can see that HNH_{N}, HSNLH_{SNL} and HSNRH_{SNR} are unchanged in the new basis, and if we further use the relation for the KFF state Q=2θαQ=2\theta_{\alpha}, the phase factors eiσ(Q2θα)e^{i\sigma(\frac{Q}{2}-\theta_{\alpha})} of the hopping tαt_{\alpha} in Eq. 45, 46 disappear, and HSLH_{SL} and HSRH_{SR} then describe the spin degenerate p-wave Kitaev chains with superconducting phase ϕ\phi and σQ(Ns1)-\sigma Q(N_{s}-1), which means HJc(ϕ)H_{Jc}(\phi) describes the Josephson chain consisting of two spin degenerate pp-wave Kitaev chains with phase bias ϕσ=ϕ+σQ(Ns1)\phi_{\sigma}=\phi+\sigma Q(N_{s}-1) for spin species σ\sigma. If we further assume the Josephson current for the Josephson chain consisting of two spin degenerate pp-wave Kitaev chains with phase bias ϕ\phi as I0(ϕ)I_{0}(\phi) which is identical for the two spin species due to the spin degeneracy, we can then immediately get the Josephson current for each spin species as Iσ(ϕ)=I0(ϕ+σQ(Ns1))I_{\sigma}(\phi)=I_{0}(\phi+\sigma Q(N_{s}-1)), i.e., the Josephson current Iσ(ϕ)I_{\sigma}(\phi) is shifted from the current of the transformed junction I0(ϕ)I_{0}(\phi) by a phase σQ(Ns1)\sigma Q(N_{s}-1) (mod 2π2\pi) so that the relative phase difference of the current between the two spin species is δϕ=2σQ(Ns1)\delta\phi=2\sigma Q(N_{s}-1) (mod 2π2\pi). We verify this relation numerically in Fig. 5 of the main text where we take the parameters as t1=1t_{1}=1, t2=0t_{2}=0, α=tan(π20)\alpha=\tan(\frac{\pi}{20}), μ=0\mu=0, V1=2V_{1}=2, which leads to Q=π10Q=\frac{\pi}{10} and Δ=0.169\Delta_{\parallel}=0.169. For NsN_{s} =321, 323, 326, 329 and 331 which leads to the phase difference δϕ\delta\phi varying from 0 to 2π2\pi, IσI_{\sigma} is shifted by 0, ±π5\pm\frac{\pi}{5}, ±π2\pm\frac{\pi}{2}, ±4π5\pm\frac{4\pi}{5} and ±π\pm\pi, which is consistent with the results shown in Fig. 5(a-e) of the main text. Various Josephson junction states including 𝟎\mathbf{0}, 𝟎\mathbf{0^{\prime}}, 𝟎𝝅\mathbf{0^{\prime}}-\pmb{\pi^{\prime}}, 𝝅\pmb{\pi^{\prime}} and 𝝅\pmb{\pi} junction states can be realized by tuning δϕ\delta\phi from 0 to 2π2\pi. The definition of these states is listed in Table. 1 of the main text.

Moreover, if the second neighbor hopping t2t_{2} is finite, after the gauge transformation of Eq. 44, Eq. 45,46 acquire extra term t2eiσQdiσdi+2σ-t_{2}e^{i\sigma Q}d_{i\sigma}^{\dagger}d_{i+2\sigma} as the 2nd neighbor hopping. Now, since the relation Q=2θαQ=2\theta_{\alpha} no longer holds, neither this phase eiσQe^{i\sigma Q} nor the phase eiσ(Q2θα)e^{i\sigma(\frac{Q}{2}-\theta_{\alpha})} of tαt_{\alpha} in Eq. 45,46 can be gauged away, the transformed model no longer describes the Josephson chain consisting of two spin degenerate p-wave Kitaev chains, but rather two spin dependent p-wave Kitaev chains with complex hopping parameters that are time-reversal counterparts of each other, so that the time-reversal symmetry is not broken.

Appendix E Δ\Delta_{\perp} channel (mixture of ss and pzp_{z} wave pairing state)

If we consider the meanfield Hamiltonian purely in the opposite-spin pairing channel, then the meanfield Hamiltonian becomes

H^MFμN^=kσ(εkσμ)ckσckσ2V2Δkckckcos(k+ϕ)+h.c.+2NcV2Δ2\begin{split}\hat{H}_{MF}-\mu\hat{N}&=\sum_{k\sigma}\left(\varepsilon_{k\sigma}-\mu\right)c_{k\sigma}^{\dagger}c_{k\sigma}\\ &-2V_{2}\Delta_{\perp}\sum_{k}c_{k\uparrow}^{\dagger}c_{-k\downarrow}^{\dagger}\cos\left(k+\phi_{\perp}\right)\\ &+h.c.+2N_{c}V_{2}\Delta_{\perp}^{2}\end{split} (50)

which can be further simplified in the Nambu basis ψk=(ck,ck)\psi_{k}^{\dagger}=\left(c_{k\uparrow}^{\dagger},c_{-k\downarrow}\right) as

HMFμN=kψkhkψk+2NcV2Δ2μNcH_{MF}-\mu N=\sum_{k}\psi_{k}^{\dagger}h_{k}\psi_{k}+2N_{c}V_{2}\Delta_{\perp}^{2}-\mu N_{c} (51)

with

hk=[εkμ2V2Δcos(k+ϕ)2V2Δcos(k+ϕ)εk+μ]h_{k}=\left[\begin{array}[]{cc}\varepsilon_{k\uparrow}-\mu&-2V_{2}\Delta_{\perp}\cos\left(k+\phi_{\perp}\right)\\ -2V_{2}\Delta_{\perp}\cos\left(k+\phi_{\perp}\right)&-\varepsilon_{-k\downarrow}+\mu\end{array}\right] (52)

Diagonalizing hkh_{k} , and considering the relation εk=εk=εk=2tαcos(kθα)2t2cos(2k)\varepsilon_{k\uparrow}=\varepsilon_{-k\downarrow}=\varepsilon_{k}=-2t_{\alpha}\cos(k-\theta_{\alpha})-2t_{2}\cos(2k) owing to the 𝒯{\cal T} symmetry, we can get

Ek,±=±(εkμ)2+4V22Δ2cos2(k+ϕ)=±EkE_{k,\pm}=\pm\sqrt{(\varepsilon_{k}-\mu)^{2}+4V_{2}^{2}\Delta_{\perp}^{2}\cos^{2}(k+\phi_{\perp})}=\pm E_{k} (53)

Then the free energy density at zero temperature Ω(Δ,ϕ)\Omega(\Delta_{\perp},\phi_{\perp}) can be calculated as

Ω(Δ,ϕ)=1NcH^MFμN^=1Nck,n=±Ek,nΘ(Ek,n)+2V2Δ2μ\Omega(\Delta_{\perp},\phi_{\perp})=\frac{1}{N_{c}}\left\langle\hat{H}_{MF}-\mu\hat{N}\right\rangle=\frac{1}{N_{c}}\sum_{k,n=\pm}E_{k,n}\Theta(-E_{k,n})+2V_{2}\Delta_{\perp}^{2}-\mu (54)

with Θ(x)\Theta(x) the Heaviside step function. Therefore, for a given value of ϕ\phi_{\perp}, the order parameter Δ\Delta_{\perp} can be determined self-consistently by minimizing Ω(Δ\Omega(\Delta_{\perp}, ϕ)\phi_{\perp}) with respect to Δ\Delta_{\perp}, leading to the self-consistent equation

Δ=1Nckcos(k+ϕ)ckck\Delta_{\perp}=\frac{1}{N_{c}}\sum_{k}\cos\left(k+\phi_{\perp}\right)\left\langle c_{-k\downarrow}c_{k\uparrow}\right\rangle (55)

The value of ϕ\phi_{\perp} can be further determined by minimizing Ω(Δ,ϕ)\Omega(\Delta_{\perp},\phi_{\perp}) with respect to ϕ\phi_{\perp}, which is equivalent to have ϕΩ(Δ,ϕ)=0\partial_{\phi_{\perp}}\Omega(\Delta_{\perp},\phi_{\perp})=0. From Eq. 53 and Eq. 54, we have

ϕΩ(Δ,ϕ)=1NckϕEk=V22Δ22πππsin(2k+2ϕ)Ek𝑑k\partial_{\phi_{\perp}}\Omega(\Delta_{\perp},\phi_{\perp})=-\frac{1}{N_{c}}\sum_{k}\partial_{\phi_{\perp}}E_{k}=-\frac{V_{2}^{2}\Delta_{\perp}^{2}}{2\pi}\int_{-\pi}^{\pi}\frac{\sin(2k+2\phi_{\perp})}{E_{k}}dk (56)

If we further consider the case with t2=0t_{2}=0, then we have

ϕΩ(Δ,ϕ)=V22Δ22πππsin(2k)[2tαcos(kϕθα)μ]2+4V22Δ2cos2k𝑑k=V22Δ22πππsin(2k)[2tαcos(k+π2ϕθα)μ]2+4V22Δ2sin2k𝑑k\begin{split}\partial_{\phi_{\perp}}\Omega(\Delta_{\perp},\phi_{\perp})&=-\frac{V_{2}^{2}\Delta_{\perp}^{2}}{2\pi}\int_{-\pi}^{\pi}\frac{\sin(2k)}{\sqrt{\left[-2t_{\alpha}\cos(k-\phi_{\perp}-\theta_{\alpha})-\mu\right]^{2}+4V_{2}^{2}\Delta_{\perp}^{2}\cos^{2}k}}dk\\ &=\frac{V_{2}^{2}\Delta_{\perp}^{2}}{2\pi}\int_{-\pi}^{\pi}\frac{\sin(2k)}{\sqrt{\left[-2t_{\alpha}\cos(k+\frac{\pi}{2}-\phi_{\perp}-\theta_{\alpha})-\mu\right]^{2}+4V_{2}^{2}\Delta_{\perp}^{2}\sin^{2}k}}dk\end{split} (57)

Apparently, when ϕ=nπ2θα\phi_{\perp}=\frac{n\pi}{2}-\theta_{\alpha} with integer n, the denominator of the integral is even in k while the numerator sin(2k)\sin(2k) is odd in k, so that this integral vanishes, which means Ω(Δ,ϕ)\Omega(\Delta_{\perp},\phi_{\perp}) reaches extremum when ϕ=nπ2θα\phi_{\perp}=\frac{n\pi}{2}-\theta_{\alpha} and which one (odd n or even n) is the minimum depends on the details of the parameters. We note that when the free energy reaches a minimum at ϕ=π2θα\phi_{\perp}=\frac{\pi}{2}-\theta_{\alpha}, the energy spectrum becomes Ek(ϕπ2),±=±(2tαcoskμ)2+4V22Δ2sin2kE_{k-(\phi_{\perp}-\frac{\pi}{2}),\pm}=\pm\sqrt{(-2t_{\alpha}\cos k-\mu)^{2}+4V_{2}^{2}\Delta_{\perp}^{2}\sin^{2}k} which is identical to the spectrum of the KFF state if V1=V2V_{1}=V_{2}.

References

  • Ando et al. [2020] Fuyuki Ando, Yuta Miyasaka, Tian Li, Jun Ishizuka, Tomonori Arakawa, Yoichi Shiota, Takahiro Moriyama, Youichi Yanase,  and Teruo Ono, “Observation of superconducting diode effect,” Nature 584, 373–376 (2020).
  • Bauriedl et al. [2021] Lorenz Bauriedl, Christian Bäuml, Lorenz Fuchs, Christian Baumgartner, Nicolas Paulik, Jonas M. Bauer, Kai-Qiang Lin, John M. Lupton, Takashi Taniguchi, Kenji Watanabe, Christoph Strunk,  and Nicola Paradiso, “Supercurrent diode effect and magnetochiral anisotropy in few-layer NbSe2,” arXiv e-prints , arXiv:2110.15752 (2021), arXiv:2110.15752 [cond-mat.supr-con] .
  • Shin et al. [2021] Jeacheol Shin, Suhan Son, Jonginn Yun, Giung Park, Kaixuan Zhang, Young Jae Shin, Je-Geun Park,  and Dohun Kim, “Magnetic Proximity-Induced Superconducting Diode Effect and Infinite Magnetoresistance in van der Waals Heterostructure,” arXiv e-prints , arXiv:2111.05627 (2021), arXiv:2111.05627 [cond-mat.supr-con] .
  • Lin et al. [2022] Jiang-Xiazi Lin, Phum Siriviboon, Harley D. Scammell, Song Liu, Daniel Rhodes, K. Watanabe, T. Taniguchi, James Hone, Mathias S. Scheurer,  and J. I. A. Li, “Zero-field superconducting diode effect in small-twist-angle trilayer graphene,” Nature Physics 18, 1221–1227 (2022).
  • Wu et al. [2022] Heng Wu, Yaojia Wang, Yuanfeng Xu, Pranava K. Sivakumar, Chris Pasco, Ulderico Filippozzi, Stuart S. P. Parkin, Yu-Jia Zeng, Tyrel McQueen,  and Mazhar N. Ali, “The field-free josephson diode in a van der waals heterostructure,” Nature 604, 653–656 (2022).
  • Baumgartner et al. [2022a] Christian Baumgartner, Lorenz Fuchs, Andreas Costa, Simon Reinhardt, Sergei Gronin, Geoffrey C. Gardner, Tyler Lindemann, Michael J. Manfra, Paulo E. Faria Junior, Denis Kochan, Jaroslav Fabian, Nicola Paradiso,  and Christoph Strunk, “Supercurrent rectification and magnetochiral effects in symmetric josephson junctions,” Nature Nanotechnology 17, 39–44 (2022a).
  • Baumgartner et al. [2022b] C Baumgartner, L Fuchs, A Costa, Jordi Picó-Cortés, S Reinhardt, S Gronin, G C Gardner, T Lindemann, M J Manfra, P E Faria Junior, D Kochan, J Fabian, N Paradiso,  and C Strunk, “Effect of rashba and dresselhaus spin–orbit coupling on supercurrent rectification and magnetochiral anisotropy of ballistic josephson junctions,” Journal of Physics: Condensed Matter 34, 154005 (2022b).
  • Pal et al. [2022] Banabir Pal, Anirban Chakraborty, Pranava K. Sivakumar, Margarita Davydova, Ajesh K. Gopi, Avanindra K. Pandeya, Jonas A. Krieger, Yang Zhang, Mihir Date, Sailong Ju, Noah Yuan, Niels B. M. Schröter, Liang Fu,  and Stuart S. P. Parkin, “Josephson diode effect from cooper pair momentum in a topological semimetal,” Nature Physics 18, 1228–1233 (2022).
  • Díez-Mérida et al. [2023] J. Díez-Mérida, A. Díez-Carlón, S. Y. Yang, Y. M. Xie, X. J. Gao, J. Senior, K. Watanabe, T. Taniguchi, X. Lu, A. P. Higginbotham, K. T. Law,  and Dmitri K. Efetov, “Symmetry-broken josephson junctions and superconducting diodes in magic-angle twisted bilayer graphene,” Nature Communications 14, 2396 (2023).
  • Hu et al. [2007] Jiangping Hu, Congjun Wu,  and Xi Dai, “Proposed design of a josephson diode,” Phys. Rev. Lett. 99, 067004 (2007).
  • Buzdin [2008] A. Buzdin, “Direct coupling between magnetism and superconducting current in the josephson φ0{\varphi}_{0} junction,” Phys. Rev. Lett. 101, 107005 (2008).
  • Yokoyama et al. [2013] Tomohiro Yokoyama, Mikio Eto,  and Yuli V. Nazarov, “Josephson current through semiconductor nanowire with spin–orbit interaction in magnetic field,” Journal of the Physical Society of Japan 82, 054703 (2013).
  • Yokoyama et al. [2014] Tomohiro Yokoyama, Mikio Eto,  and Yuli V. Nazarov, “Anomalous josephson effect induced by spin-orbit interaction and zeeman effect in semiconductor nanowires,” Phys. Rev. B 89, 195407 (2014).
  • Dolcini et al. [2015] Fabrizio Dolcini, Manuel Houzet,  and Julia S. Meyer, “Topological josephson ϕ0{\phi}_{0} junctions,” Phys. Rev. B 92, 035428 (2015).
  • Szombati et al. [2016] D. B. Szombati, S. Nadj-Perge, D. Car, S. R. Plissard, E. P. A. M. Bakkers,  and L. P. Kouwenhoven, “Josephson ϕ\phi0-junction in nanowire quantum dots,” Nature Physics 12, 568–572 (2016).
  • Assouline et al. [2019] Alexandre Assouline, Cheryl Feuillet-Palma, Nicolas Bergeal, Tianzhen Zhang, Alireza Mottaghizadeh, Alexandre Zimmers, Emmanuel Lhuillier, Mahmoud Eddrie, Paola Atkinson, Marco Aprili,  and Hervé Aubin, “Spin-orbit induced phase-shift in bi2se3 josephson junctions,” Nature Communications 10, 126 (2019).
  • Alidoust and Linder [2013] Mohammad Alidoust and Jacob Linder, “φ\varphi-state and inverted fraunhofer pattern in nonaligned josephson junctions,” Phys. Rev. B 87, 060503 (2013).
  • Bobkova et al. [2016] I. V. Bobkova, A. M. Bobkov, Alexander A. Zyuzin,  and Mohammad Alidoust, “Magnetoelectrics in disordered topological insulator josephson junctions,” Phys. Rev. B 94, 134506 (2016).
  • Alidoust and Hamzehpour [2017] Mohammad Alidoust and Hossein Hamzehpour, “Spontaneous supercurrent and φ0{\varphi}_{0} phase shift parallel to magnetized topological insulator interfaces,” Phys. Rev. B 96, 165422 (2017).
  • Alidoust et al. [2018] Mohammad Alidoust, Morten Willatzen,  and Antti-Pekka Jauho, “Strain-engineered majorana zero energy modes and φ0{\varphi}_{0} josephson state in black phosphorus,” Phys. Rev. B 98, 085414 (2018).
  • Alidoust [2020] Mohammad Alidoust, “Critical supercurrent and φ0{\varphi}_{0} state for probing a persistent spin helix,” Phys. Rev. B 101, 155123 (2020).
  • Alidoust et al. [2021] Mohammad Alidoust, Chenghao Shen,  and Igor Žutić, “Cubic spin-orbit coupling and anomalous josephson effect in planar junctions,” Phys. Rev. B 103, L060503 (2021).
  • Halterman et al. [2022] Klaus Halterman, Mohammad Alidoust, Ross Smith,  and Spencer Starr, “Supercurrent diode effect, spin torques, and robust zero-energy peak in planar half-metallic trilayers,” Phys. Rev. B 105, 104508 (2022).
  • Misaki and Nagaosa [2021] Kou Misaki and Naoto Nagaosa, “Theory of the nonreciprocal josephson effect,” Phys. Rev. B 103, 245302 (2021).
  • He et al. [2022] James Jun He, Yukio Tanaka,  and Naoto Nagaosa, “A phenomenological theory of superconductor diodes,” New Journal of Physics 24, 053014 (2022).
  • Yuan and Fu [2022] Noah F. Q. Yuan and Liang Fu, “Supercurrent diode effect and finite-momentum superconductors,” Proceedings of the National Academy of Sciences 119, e2119548119 (2022).
  • Daido et al. [2022] Akito Daido, Yuhei Ikeda,  and Youichi Yanase, “Intrinsic superconducting diode effect,” Phys. Rev. Lett. 128, 037001 (2022).
  • Davydova et al. [2022] Margarita Davydova, Saranesh Prembabu,  and Liang Fu, “Universal josephson diode effect,” Science Advances 8, eabo0309 (2022).
  • Scammell et al. [2022] Harley D Scammell, J I A Li,  and Mathias S Scheurer, “Theory of zero-field superconducting diode effect in twisted trilayer graphene,” 2D Materials 9, 025027 (2022).
  • Zhang et al. [2022] Yi Zhang, Yuhao Gu, Pengfei Li, Jiangping Hu,  and Kun Jiang, “General theory of josephson diodes,” Phys. Rev. X 12, 041013 (2022).
  • Souto et al. [2022] Rubén Seoane Souto, Martin Leijnse,  and Constantin Schrade, “Josephson diode effect in supercurrent interferometers,” Phys. Rev. Lett. 129, 267702 (2022).
  • Fulde and Ferrell [1964] Peter Fulde and Richard A. Ferrell, “Superconductivity in a strong spin-exchange field,” Phys. Rev. 135, A550–A563 (1964).
  • Larkin and Ovchinnikov [1965] A.I. Larkin and Yu.N. Ovchinnikov, “Nonuniform state of superconductors,” Sov. Phys. JETP 20, 762 (1965).
  • 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,” Reports on Progress in Physics 80, 036501 (2017).
  • Gor’kov and Rashba [2001] Lev P. Gor’kov and Emmanuel I. Rashba, “Superconducting 2d system with lifted spin degeneracy: Mixed singlet-triplet state,” Phys. Rev. Lett. 87, 037004 (2001).
  • Barzykin and Gor’kov [2002] Victor Barzykin and Lev P. Gor’kov, “Inhomogeneous stripe phase revisited for surface superconductivity,” Phys. Rev. Lett. 89, 227002 (2002).
  • Agterberg [2003] D.F. Agterberg, “Novel magnetic field effects in unconventional superconductors,” Physica C: Superconductivity 387, 13–16 (2003), proceedings of the 3rd Polish-US Workshop on Superconductivity and Magnetism of Advanced Materials.
  • Dimitrova and Feigel’man [2003] O. V. Dimitrova and M. V. Feigel’man, “Phase diagram of a surface superconductor in parallel magnetic field,” Journal of Experimental and Theoretical Physics Letters 78, 637–641 (2003).
  • Kaur et al. [2005] R. P. Kaur, D. F. Agterberg,  and M. Sigrist, “Helical vortex phase in the noncentrosymmetric cept3Si{\mathrm{cept}}_{3}\mathrm{Si},” Phys. Rev. Lett. 94, 137002 (2005).
  • Agterberg and Kaur [2007] D. F. Agterberg and R. P. Kaur, “Magnetic-field-induced helical and stripe phases in rashba superconductors,” Phys. Rev. B 75, 064511 (2007).
  • Dimitrova and Feigel’man [2007] Ol’ga Dimitrova and M. V. Feigel’man, “Theory of a two-dimensional superconductor with broken inversion symmetry,” Phys. Rev. B 76, 014522 (2007).
  • Samokhin [2008] K. V. Samokhin, “Upper critical field in noncentrosymmetric superconductors,” Phys. Rev. B 78, 224520 (2008).
  • Yanase and Sigrist [2008] Youichi Yanase and Manfred Sigrist, “Helical superconductivity in non-centrosymmetric superconductors with dominantly spin triplet pairing,” Journal of the Physical Society of Japan 77, 342–344 (2008).
  • Michaeli et al. [2012] Karen Michaeli, Andrew C. Potter,  and Patrick A. Lee, “Superconducting and ferromagnetic phases in srtio3/laalo3{\mathrm{srtio}}_{3}/{\mathrm{laalo}}_{3} oxide interface structures: Possibility of finite momentum pairing,” Phys. Rev. Lett. 108, 117003 (2012).
  • Sekihara et al. [2013] Takayuki Sekihara, Ryuichi Masutomi,  and Tohru Okamoto, “Two-dimensional superconducting state of monolayer pb films grown on gaas(110) in a strong parallel magnetic field,” Phys. Rev. Lett. 111, 057005 (2013).
  • Houzet and Meyer [2015] Manuel Houzet and Julia S. Meyer, “Quasiclassical theory of disordered rashba superconductors,” Phys. Rev. B 92, 014509 (2015).
  • Yuan and Fu [2021] Noah F. Q. Yuan and Liang Fu, “Topological metals and finite-momentum superconductors,” Proceedings of the National Academy of Sciences 118, e2019063118 (2021).
  • Lee et al. [2019] Kyungmin Lee, Tamaghna Hazra, Mohit Randeria,  and Nandini Trivedi, “Topological superconductivity in dirac honeycomb systems,” Phys. Rev. B 99, 184514 (2019).
  • Zhang et al. [2021] Yi Zhang, Kun Jiang, Fuchun Zhang, Jian Wang,  and Ziqiang Wang, “Atomic line defects and topological superconductivity in unconventional superconductors,” Phys. Rev. X 11, 011041 (2021).
  • Chen et al. [2020] Cheng Chen, Kun Jiang, Yi Zhang, Chaofei Liu, Yi Liu, Ziqiang Wang,  and Jian Wang, “Atomic line defects and zero-energy end states in monolayer fe(te,se) high-temperature superconductors,” Nature Physics 16, 536–540 (2020).
  • Kitaev [2001] A Yu Kitaev, “Unpaired majorana fermions in quantum wires,” Physics-Uspekhi 44, 131–136 (2001).
  • Beenakker [1991] C. W. J. Beenakker, “Universal limit of critical-current fluctuations in mesoscopic josephson junctions,” Phys. Rev. Lett. 67, 3836–3839 (1991).
  • Mao and Sun [2022] Yue Mao and Qing-Feng Sun, “Spin phase regulated spin josephson supercurrent in topological superconductor,” Phys. Rev. B 105, 184511 (2022).
  • Golubov et al. [2004] A. A. Golubov, M. Yu. Kupriyanov,  and E. Il’ichev, “The current-phase relation in josephson junctions,” Rev. Mod. Phys. 76, 411–469 (2004).
  • Sauls [2018] J. A. Sauls, “Andreev bound states and their signatures,” Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences 376, 20180140 (2018).
  • Rozhkov and Arovas [1999] A. V. Rozhkov and Daniel P. Arovas, “Josephson coupling through a magnetic impurity,” Phys. Rev. Lett. 82, 2788–2791 (1999).
  • Siano and Egger [2004] F. Siano and R. Egger, “Josephson current through a nanoscale magnetic quantum dot,” Phys. Rev. Lett. 93, 047002 (2004).
  • Liu et al. [2022] Yanzhao Liu, Tianheng Wei, Guanyang He, Yi Zhang, Ziqiang Wang,  and Jian Wang, “Discovery of a pair density wave state in a monolayer high-Tc iron-based superconductor,” arXiv e-prints , arXiv:2209.04592 (2022), arXiv:2209.04592 [cond-mat.supr-con] .
  • Robinson et al. [2010] J. W. A. Robinson, J. D. S. Witt,  and M. G. Blamire, “Controlled injection of spin-triplet supercurrents into a strong ferromagnet,” Science 329, 59–61 (2010).
  • Khaire et al. [2010] Trupti S. Khaire, Mazin A. Khasawneh, W. P. Pratt,  and Norman O. Birge, “Observation of spin-triplet superconductivity in co-based josephson junctions,” Phys. Rev. Lett. 104, 137002 (2010).