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

Intrinsic Superconducting Diode Effect

Akito Daido Department of Physics, Graduate School of Science, Kyoto University, Kyoto 606-8502, Japan daido@scphys.kyoto-u.ac.jp    Yuhei Ikeda Department of Physics, Graduate School of Science, Kyoto University, Kyoto 606-8502, Japan    Youichi Yanase Department of Physics, Graduate School of Science, Kyoto University, Kyoto 606-8502, Japan Institute for Molecular Science, Okazaki 444-8585, Japan
Abstract

Stimulated by the recent experiment [F. Ando et al., Nature 584, 373 (2020)], we propose an intrinsic mechanism to cause the superconducting diode effect (SDE). SDE refers to the nonreciprocity of the critical current for the metal-superconductor transition. Among various mechanisms for the critical current, the depairing current is known to be intrinsic to each material and has recently been observed in several superconducting systems. We clarify the temperature scaling of the nonreciprocal depairing current near the critical temperature and point out its significant enhancement at low temperatures. It is also found that the nonreciprocal critical current shows sign reversals upon increasing the magnetic field. These behaviors are understood by the nonreciprocity of the Landau critical momentum and the change in the nature of the helical superconductivity. The intrinsic SDE unveils the rich phase diagram and functionalities of noncentrosymmetric superconductors.

Introduction. — Rectification by the semiconductor diode is one of the central building blocks of electronic devices. Apart from the nonreciprocity induced by asymmetric junctions, it has been revealed that nonreciprocal transport can be obtained as a bulk property of materials Tokura and Nagaosa (2018); Ideue and Iwasa (2021). Magnetochiral anisotropy (MCA) Rikken et al. (2001); Krstić et al. (2002); Pop et al. (2014); Rikken and Wyder (2005); Ideue et al. (2017); Wakatsuki and Nagaosa (2018); Hoshino et al. (2018); Wakatsuki et al. (2017); Qin et al. (2017); Yasuda et al. (2019); Itahashi et al. (2020) is an example, described by the equation R(j)=R0(1+γjh).R(j)=R_{0}(1+\gamma j{h}). Here RR, jj, and h{h} are the resistance, electric current, and the magnetic field, respectively. The coefficient γ\gamma gives rise to different resistance for rightward and leftward electric currents and can be finite in noncentrosymmetric materials. MCA has been observed in (semi)conductors Rikken and Wyder (2005); Ideue et al. (2017); Krstić et al. (2002); Pop et al. (2014) as well as in superconductors Wakatsuki et al. (2017); Qin et al. (2017); Yasuda et al. (2019); Itahashi et al. (2020), and allows us to access various aspects of noncentrosymmetric materials: from spin-orbit splitting in the band structure Ideue et al. (2017) to the spin-singlet and -triplet mixing of Cooper pairs Wakatsuki et al. (2017); Wakatsuki and Nagaosa (2018); Hoshino et al. (2018)

MCA is the inequivalence of R(j)R(j) and R(j)R(-j), where both R(±j)R(\pm j) usually take finite values. On the other hand, such a drastic situation is possible in superconductors that either one of R(±j)R(\pm j) vanishes while the other remains finite [Fig. 1]. Such a superconducting diode effect (SDE) has recently been observed in the Nb/V/Ta superlattice without an inversion center and is controlled by the applied inplane magnetic field Ando et al. (2020). This is the first report of the SDE in a bulk material, while similar effects have been recognized in engineered systems Reynoso et al. (2008); Zazunov et al. (2009); Margaris et al. (2010); Yokoyama et al. (2014); Silaev et al. (2014); Campagnano et al. (2015); Dolcini et al. (2015); Chen et al. (2018); Minutillo et al. (2018); Pal and Benjamin (2019); Kopasov et al. (2021) and followed by recent SDE experiments Baumgartner et al. (2021); Lyu et al. (2021). SDE is a promising building block of the dissipationless electric circuits, and is a fascinating phenomenon manifesting the interplay of the inversion breaking and superconductivity. One of the remaining issues is to identify suitable materials providing the best performance; however, the mechanisms to cause SDE in a bulk material Ando et al. (2020) have not been clarified, while the SDE in artificial devices Lyu et al. (2021); Baumgartner et al. (2021) has been well simulated by Bogoliubov-de Gennes (BdG) Baumgartner et al. (2021) and time-dependent Ginzburg-Landau (GL) theories Lyu et al. (2021).

Refer to caption
Figure 1: Schematic figure for the SDE. The system has zero and finite resistance for the rightward and leftward current, respectively, and vice versa when the magnetic field hh is reversed.

SDE is the nonreciprocity of the critical current for the resistive transition. In usual situations, in particular, under out-of-plane magnetic fields, the resistive transition is caused by the vortex motion. The details of the vortex motion depend on the device setup such as impurity concentrations Blatter et al. (1994), and in turn, has an advantage of tunability by the nanostructure engineering Villegas et al. (2003); Lyu et al. (2021). Apart from the extrinsic mechanisms to cause resistivity, the depairing current is known as the critical current unique to each superconducting material. Here, the metal-superconductor transition is literally caused by the dissociation of the flowing Cooper pairs Tinkham (2004); Dew-Hughes (2001). The depairing current always gives the upper limit of the critical current and is an important material parameter characterizing superconductors Blatter et al. (1994). The depairing limit generally requires a huge current density, but is within the scope of experimental techniques. Indeed, the depairing limit has recently been achieved in the microbridge superconducting devices of YBa2Cu3O7-δ Nawaz et al. (2013), Ba0.5K0.5Fe2As2 Li et al. (2013), and Fe1+yTe1-xSex Sun et al. (2020).

In this Letter, as a first step of the theoretical research on SDE of bulk materials, we propose the intrinsic mechanism of SDE by studying the nonreciprocity in the depairing current. The results can be tested with the microbridge experiments and establish the foundation of the future study on the bulk SDE. Furthermore, it is revealed that the intrinsic SDE is closely related to the Flude-Ferrell-Larkin-Ovchinnikov state Larkin and Ovchinnikov (1964); Fulde and Ferrell (1964). While the Larkin-Ovchinnikov (LO) state with the spatially inhomogeneous pair potential Δ(x)=Δcosqx\Delta(x)=\Delta\cos qx has been discussed for FeSe Kasahara et al. (2014, 2020), CeCoIn5 Matsuda and Shimahara (2007) and organic superconductors Wosnitza (2018), the Flude-Fellel (FF) type order parameter Δ(x)=Δeiqx\Delta(x)=\Delta e^{iqx} is known to ubiquitously appear in noncentrosymmetric superconductors and is particularly called the helical superconductivity Bauer and Sigrist (2012); Smidman et al. (2017); Agterberg (2003); Barzykin and Gor’kov (2002); Dimitrova and Feigel’man (2003); Kaur et al. (2005); Agterberg and Kaur (2007); Dimitrova and Feigel’man (2007); Samokhin (2008); Yanase and Sigrist (2008); Bauer and Sigrist (2012); Michaeli et al. (2012); Sekihara et al. (2013); Houzet and Meyer (2015). Implications of the helical superconductivity have been obtained in thin films of Pb Sekihara et al. (2013) and doped SrTiO3 Schumann et al. (2020), and a heavy-fermion superlattice Naritsuka et al. (2017, 2021). We show that the intrinsic SDE works as a probe to study the phase diagram of helical superconductivity. Relation to the recent experiments Ando et al. (2020); Ono et al. is also discussed.

Model. — We consider the critical current in two-dimensional (2D) superconductors with a polar axis due to the substrate and/or the crystal structure. The magnetic field is applied along the yy direction, which makes the critical current nonreciprocal in the xx direction [Fig. 1]. The system is modeled by the Rashba-Zeeman Hamiltonian with the attractive Hubbard interaction,

H^\displaystyle\hat{H} =𝒌σσ[ξ(𝒌)δσσ+𝒈(𝒌)𝝈σσh(σy)σσ]c𝒌σc𝒌σ\displaystyle=\sum_{\bm{k}\sigma\sigma^{\prime}}\bigl{[}\xi(\bm{k})\delta_{\sigma\sigma^{\prime}}+\bm{g}(\bm{k})\cdot\bm{\sigma}_{\sigma\sigma^{\prime}}-h(\sigma_{y})_{\sigma\sigma^{\prime}}\bigr{]}c_{\bm{k}\sigma}^{\dagger}c_{\bm{k}\sigma^{\prime}}
UV𝒌1+𝒌2+𝒌3+𝒌4=𝟎c𝒌1c𝒌2c𝒌3c𝒌4.\displaystyle\qquad\quad-\frac{U}{V}\sum_{\bm{k}_{1}+\bm{k}_{2}+\bm{k}_{3}+\bm{k}_{4}=\bm{0}}c^{\dagger}_{\bm{k}_{1}\uparrow}c^{\dagger}_{\bm{k}_{2}\downarrow}c_{\bm{k}_{3}\downarrow}c_{\bm{k}_{4}\uparrow}. (1)

Here, ξ(𝒌)=2t1(coskx+cosky)+4t2coskxcoskyμ\xi(\bm{k})=-2t_{1}(\cos k_{x}+\cos k_{y})+4t_{2}\cos k_{x}\cos k_{y}-\mu and 𝒈(𝒌)=αg(sinky,sinkx, 0)\bm{g}(\bm{k})=\alpha_{g}(-\sin k_{y},\,\sin k_{x},\,0) represent the hopping energy and the Rashba spin-orbit coupling, respectively. The magnetic field in the yy direction is introduced by the Zeeman term hμBHyh\equiv\mu_{B}H_{y}. The parameters are given by (t1,t2,μ,αg,U)=(1,0,1,0.3,1.5)(t_{1},t_{2},\mu,\alpha_{g},U)=(1,0,-1,0.3,1.5) unless mentioned otherwise. The next-nearest-neighbor hopping t2t_{2} is introduced for the latter use. The energy dispersion of the noninteracting part is given by ξχh(𝒌)=ξ(𝒌)+χ|𝒈(𝒌)hy^|ξχ0(𝒌𝒒χ(𝒌)/2)\xi_{\chi}^{h}(\bm{k})=\xi(\bm{k})+\chi|\bm{g}(\bm{k})-h\hat{y}|\simeq\xi^{0}_{\chi}(\bm{k}-\bm{q}_{\chi}(\bm{k})/2). Here, each band is labeled by the helicity χ=±\chi=\pm, and the momentum shift under hh is given by 𝒒χ(𝒌)/2=χgy(𝒌)h𝒗χ(𝒌)/|𝒗χ(𝒌)|2\bm{q}_{\chi}(\bm{k})/2=\chi g_{y}(\bm{k})h\bm{v}_{\chi}(\bm{k})/|\bm{v}_{\chi}(\bm{k})|^{2} with 𝒗χ(𝒌)ξχ0(𝒌)\bm{v}_{\chi}(\bm{k})\equiv\nabla\xi^{0}_{\chi}(\bm{k}). The momentum shift is estimated by its Fermi-surface (FS) average, qχx^𝒒χ(𝒌)FS2χh/|𝒗χ(𝒌)|FSq_{\chi}\equiv\hat{x}\cdot\braket{\bm{q}_{\chi}(\bm{k})}_{\mathrm{FS}}\sim 2\chi h/\braket{}{\bm{v}_{\chi}(\bm{k})}{}_{\mathrm{FS}}.

We solve the model (1) within the mean-field approximation. The attractive Hubbard interaction is approximated by

12𝒌σσΔ(iσy)σσc𝒌+𝒒σc𝒌σ+H.c.+Δ2/2U.\displaystyle{\frac{1}{2}}\sum_{\bm{k}\sigma\sigma^{\prime}}\Delta(i\sigma_{y})_{\sigma\sigma^{\prime}}c^{\dagger}_{\bm{k}+\bm{q}\sigma}c^{\dagger}_{-\bm{k}\sigma^{\prime}}+\mathrm{H.c.}+\Delta^{2}/2U. (2)

The ss-wave pair potential Δ\Delta is considered with a center-of-mass momentum 𝒒=qx^\bm{q}=q\hat{x} to describe the current-flowing state. For a given qq, the value of Δ=Δ(q)\Delta=\Delta(q) is determined self-consistently by the gap equation with the temperature TT. To describe the superconducting transitions and the supercurrent, it is convenient to introduce the condensation energy F(q)F(q) for each qq, that is, the difference of the free energy per unit area in the normal and superconducting states. The sheet current density is obtained by j(q)=2qF(q)j(q)=2\partial_{q}F(q), which coincides with the expectation value of the current operator Sup .

When an electric current jexj_{\mathrm{ex}} is applied, the superconducting state with qq satisfying j(q)=jexj(q)=j_{\mathrm{ex}} should be realized. However, no superconducting state can sustain jexj_{\mathrm{ex}} when jex<jcj_{\mathrm{ex}}<j_{c-} or jex>jc+j_{\mathrm{ex}}>j_{c+}, with jc+maxqj(q)j_{c+}\equiv\max_{q}j(q) and jcminqj(q).j_{c-}\equiv\min_{q}j(q). Thus, the depairing current in the positive and negative directions is given by the maximum jc+j_{c+} and minimum jcj_{c-} of j(q)j(q), respectively. In particular, the nonreciprocal component is given by

Δjcjc++jc=jc+|jc|.\Delta j_{c}\equiv j_{c+}+j_{c-}=j_{c+}-|j_{c-}|. (3)

The SDE is identified with a finite Δjc\Delta j_{c} of the system. We also define the averaged critical current j¯c(jc+jc)/2\bar{j}_{c}\equiv(j_{c+}-j_{c-})/2, by which the strength of the nonreciprocal nature can be expressed as rΔjc/j¯cr\equiv\Delta j_{c}/\bar{j}_{c}.

GL analysis. — First, we discuss the SDE by the GL theory. The GL free energy f(Δ,q)=α(q)Δ2+β(q)2Δ4f(\Delta,q)=\alpha(q)\Delta^{2}+\frac{\beta(q)}{2}\Delta^{4} gives a good approximation of F(q)F(q) near the transition temperature TcT_{c} when the optimized order parameter Δ=Δ(q)\Delta=\Delta(q) is substituted. The GL coefficients are assumed to have the following form: α(q)=α0+α1q+12α2q2+16α3q3\alpha(q)=\alpha_{0}+\alpha_{1}q+{\frac{1}{2}}\alpha_{2}q^{2}+{\frac{1}{6}}\alpha_{3}q^{3}, and β(q)=β0+β1q\beta(q)=\beta_{0}+\beta_{1}q, which is valid for the description up to O(TcT)5/2O(T_{c}-T)^{5/2}. When the higher-order gradient terms α3,β1\alpha_{3},\beta_{1} are neglected, the broken inversion and time-reversal symmetries are encoded solely into α10\alpha_{1}\neq 0, which shifts the minimum of f(q)=f(Δ(q),q)f(q)=f(\Delta(q),q) from q=0q=0 to q0=α1/2α2q_{0}=-\alpha_{1}/2\alpha_{2}. Thus, the superconducting state with a finite q0q_{0}, namely the helical superconductivity is realized Agterberg (2003); Smidman et al. (2017). The helical superconducting state with q=q0q=q_{0} does not carry a supercurrent Agterberg (2003); Dimitrova and Feigel’man (2003, 2007), j(q0)=2q0f(q0)=0,j(q_{0})=2\partial_{q_{0}}f(q_{0})=0, as the most stable state generally should be.

It is convenient to rewrite the GL coefficients as α(q)=α~0+α~22(qq~0)2+α36(qq~0)3\alpha(q)={\tilde{\alpha}}_{0}+\frac{\tilde{\alpha}_{2}}{2}(q-\tilde{q}_{0})^{2}+\frac{{\alpha}_{3}}{6}(q-\tilde{q}_{0})^{3} and β(q)=β~0+β1(qq~0)\beta(q)={\tilde{\beta}}_{0}+\beta_{1}(q-\tilde{q}_{0}), where the linear term in α(q)\alpha(q) is erased. Clearly, f(Δ,q+q~0)f(\Delta,q+\tilde{q}_{0}) for α3=β1=0\alpha_{3}=\beta_{1}=0 is equivalent to the GL free energy of a centrosymmetric superconductor, leading to a reciprocal critical current Smidman et al. (2017). Thus, the SDE is caused by the higher-order terms, α3\alpha_{3} and β1\beta_{1},

Δjc\displaystyle\Delta j_{c} =(1627β~0α~2α389β~02β1)α~02,\displaystyle=\left(\frac{16}{27{\tilde{\beta}}_{0}{\tilde{\alpha}}_{2}}\alpha_{3}-\frac{8}{9{\tilde{\beta}}_{0}^{2}}\beta_{1}\right){\tilde{\alpha}}_{0}^{2}, (4)

up to first order in α3\alpha_{3} and β1\beta_{1} Sup . Note that Δjc(TcT)2\Delta j_{c}\propto(T_{c}-T)^{2} in contrast to the averaged critical current j¯c(TcT)3/2\bar{j}_{c}\propto(T_{c}-T)^{3/2} Tinkham (2004); Sup , since α~0TTc{\tilde{\alpha}}_{0}\propto T-T_{c}. Thus, a small but finite Δjc\Delta j_{c} is predicted by the GL theory, while a larger Δjc\Delta j_{c} is expected at low temperatures. The result obtained here is valid for general noncentrosymmetric superconductors without orbital depairing effect, e.g., superconducting thin films under inplane magnetic fields.

Refer to caption
Figure 2: The temperature dependence of Δjc\Delta j_{c} at h=0.03h=0.03. The red closed circles indicate Δjc(T)\Delta j_{c}(T), while the open blue circles indicate Δ(T)\Delta(T) (a.u.). The dashed lines show the fitting curve of Δjc(T)\Delta j_{c}(T) and Δ(T)\Delta(T) near TcT_{c} with (TcT)2(T_{c}-T)^{2} and TcT\sqrt{T_{c}-T}, respectively. The inset shows the enlarged figure near Tc0.036T_{c}\simeq 0.036.

Critical current under low fields. — Equipped with the insight of the GL theory, we discuss the temperature dependence of Δjc\Delta j_{c} based on the model (1). The result is shown in Fig. 2 for h=0.03h=0.03. As shown in the inset, the temperature scaling Δjc(TcT)2\Delta j_{c}\propto(T_{c}-T)^{2} is confirmed near the transition temperature TcT_{c}. The scaling law becomes inaccurate as TcTT_{c}-T gets large, where Δ(T)\Delta(T) also deviates from Δ(T)TcT\Delta(T)\propto\sqrt{T_{c}-T}. Importantly, Δjc(T)\Delta j_{c}(T) is strongly enhanced at low temperatures.

To clarify the origin of the SDE, we show j(q)j(q) by red lines in Figs. 3. In Fig. 3 (a) for T=0.03TcT=0.03\simeq T_{\rm c}, j(q)j(q) is a smooth curve and its tiny asymmetry gives rise to Δjc\Delta j_{c}, as is illustrated by the difference of the solid and dashed horizontal lines (indicating jc+j_{c+} and jc-j_{c-}). This is consistent with the GL picture where Δjc\Delta j_{c} is caused by the asymmetry factors α3,β10\alpha_{3},\beta_{1}\neq 0. Two curves, j(q)j(q) and j(q)-j(q), cross at q0<0q_{0}<0, indicating the helical superconductivity. In Fig. 3 (c), Δ(q)\Delta(q) and the minimum excitation energy ΔE(q)\Delta E(q) are shown in addition to j(q)j(q), by the blue and black lines, respectively. The superconducting state remains stable even after the spectrum becomes gapless, and reaches the maximum and minimum of j(q)j(q) in the gapless region.

As shown in Fig. 3 (b), the dispersion of j(q)j(q) at T=0.001TcT=0.001\ll T_{\rm c} is significantly different from that at T=0.03T=0.03, and a large Δjc\Delta j_{c} is realized. The maximum and minimum of j(q)j(q) are achieved at the ends of the region where j(q)j(q) is almost linear in qq. These momenta approximately coincide with the Landau critical momenta, qR>0q_{R}>0 and qL<0q_{L}<0, i.e. the first qq’s satisfying ΔE(q)=0\Delta E(q)=0, as is clear from Fig. 3 (d). Actually, the depairing effect takes place after q>qRq>q_{R} or q<qLq<q_{L}: The excited quasiparticles reduce |j(q)||j(q)| and Δ(q)\Delta(q) and finally cause a first-order phase transition into the normal state. From these observations, we obtain the formula

Δjc=nxxs(qR+qL2q0)/2,\Delta j_{c}=n^{s}_{xx}(q_{R}+q_{L}-2q_{0})/2, (5)

by using, e.g., jc+=nxxs(qRq0)/2j_{c+}=n^{s}_{xx}(q_{R}-q_{0})/2 with the superfluid weight nxxs=2q0j(q0)n_{xx}^{s}=2\partial_{q_{0}}j(q_{0}). Thus, the nonreciprocal Landau critical momentum qR+qLq_{R}+q_{L} measured from 2q02q_{0} gives rise to the SDE at extremely low temperatures. As TT gets larger, the maximum and minimum of j(q)j(q) deviate from j(qR)j(q_{R}) and j(qL)j(q_{L}), and Eq. (5) becomes no longer valid. The mechanism of the SDE at low temperatures is not captured by the GL theory.

Refer to caption
Figure 3: (a), (b): The qq dependence of the supercurrent at h=0.03h=0.03 and (a) T=0.03T=0.03 and (b) T=0.001T=0.001. In addition to j(q)j(q) (red lines), j(q)-j(q) (dashed red lines) is shown. The black (dashed) horizontal lines indicate jc+j_{c+} (jc-j_{c-}). The position of q0q_{0} is indicated by arrows. (c), (d): The order parameter Δ(q)\Delta(q) (blue lines) and the excitation gap ΔE(q)\Delta E(q) (black lines) are shown together with j(q)j(q) (a.u.). The parameters for the panels (c) and (d) are the same as the panel (a) and (b), respectively. The vertical dashed black lines indicate the momentum qq where j(q)=jc±j(q)=j_{c\pm}. Landau critical momentum qRq_{R} and qLq_{L} are indicated by arrows.
Refer to caption
Figure 4: The magnetic-field and temperature dependence of the nonreciprocal component of the critical current Δjc(h,T)\Delta j_{c}(h,T) for (a) t2=0t_{2}=0 and (b) t2=0.2t_{2}=0.2. The red and blue colors indicate positive and negative values of Δjc\Delta j_{c}, respectively. The transition temperature Tc(h)T_{c}(h) determined with the TT mesh (a) 0.045/210.045/21 and (b) 0.12/210.12/21 is shown with the black line for the guide of the eye.

Phase diagram. — In Figs. 4 (a) and (b), we show the temperature and magnetic-field dependence of the nonreciprocal component Δjc\Delta j_{c} for t2=0t_{2}=0 and t2=0.2t_{2}=0.2. Let us focus on the low-field region, where positive and negative values of Δjc\Delta j_{c} are widely obtained for Fig. 4 (a) and (b), respectively. The sign reversal of Δjc\Delta j_{c} by t2t_{2} can be understood based on Eq. (5). Indeed, we show in the Supplemental Material Sup that qR+qL2q0q_{R}+q_{L}-2q_{0} causes a sign reversal as t2t_{2} increases, leading to that of Δjc\Delta j_{c} as well. It is also shown that for large values of t2t_{2}, qR+qL2q0q_{R}+q_{L}-2q_{0} is dominated by the nonreciprocal Landau critical momentum qR+qLq_{R}+q_{L}, while it is dominated by 2q0-2q_{0} for small values of t2t_{2}. A relatively large SDE for t20.2t_{2}\sim 0.2 is explained by large values of qR+qL2q0q_{R}+q_{L}-2q_{0} as a result of the anisotropy Sup . The pronounced aspect of Fig. 4 is the sign reversals prevailing under moderate and high magnetic fields. This point will be discussed in the following.

Critical current under high fields.

Refer to caption
Figure 5: (a)-(d) j(q)j(q) (red lines), F(q)F(q) (blue lines), and ΔE(q)\Delta E(q) (black lines) normalized to [1,1][-1,1] at T=0.001T=0.001 for (a) h=0.043h=0.043, (b) h=0.058h=0.058, (c) h=0.063h=0.063 and (d) h=0.075h=0.075. (e),(f) q0q_{0} and rr for various values of hh and TT.

To see the origin of the high-field behavior, we show the condensation energy F(q)F(q) at T=0.001T=0.001 for various values of hh in Fig. 5. To be specific, the case of Fig. 4 (a) is considered. The condensation energy F(q)F(q) shown by the blue line has a single-well structure under low magnetic fields [panel (a)]. The structure near |q|0.05|q|\sim 0.05 is developed under higher magnetic fields [panel (b)], to form two local minima [panel (c)], where the left one becomes most stable. These side wells are the precursor of the high-field helical superconducting states [panel (d)], where the central minimum finally disappears. Such a change is most evident in q0(T,h)q_{0}(T,h) shown in Fig. 5 (e). Under low fields, q0q_{0} is determined by the balance of two Fermi surfaces shifted in the opposite directions, resulting in |q0|102|q_{0}|\lesssim 10^{-2}; on the other hand, under high fields, q0101q_{0}\sim 10^{-1} almost coincides with qχq_{\chi} of the Fermi surface with a larger density of states Dimitrova and Feigel’man (2003); Agterberg and Kaur (2007); Smidman et al. (2017); Yanase and Sigrist (2008). This determines the “crossover line” 111This terminology is named after the crossover generally seen in noncentrosymmetric superconductors, while the line changes to the first-order transition at lower temperature T0.12T\lesssim 0.12 in this model, as in Figs.5(b)-5(d). of helical superconductivity visible at h0.06h\sim 0.06 in Fig. 5 (e).

The evolution of j(q)j(q) by hh follows that of F(q)F(q). Overall, j(q)j(q) consists of several almost-straight lines and their interpolation, since F(q)F(q) is approximated by the square function of qq around each local minimum. Comparing Figs. 5 (a) and (b), the qq point achieving jcj_{c-} is changed from qLq_{L} to a critical momentum of the left well, which we name q,Lq_{-,L}. In the panel (c), q,Lq_{-,L} remains to give jcj_{c-}, whose value is significantly enhanced owing to the development of the left minimum. This causes the sign reversal of Δjc\Delta j_{c}. In the panel (d), Δjc\Delta j_{c} is determined by the tiny asymmetry of the left well. It should be noticed that the ratio r=Δjc/j¯cr=\Delta j_{c}/\bar{j}_{c} shown in Fig. 5 (f) is quite large around the crossover line, takes values up to |r|0.4|r|\lesssim 0.4, as is understood from Figs. 5 (a)-(c). Thus, the sign reversals and huge values of r=Δjc/j¯cr=\Delta j_{c}/\bar{j}_{c} under magnetic fields are caused by the change in the helical superconducting states. According to Figs. 5 (e) and (f), the sign reversal also occurs near Tc(h)T_{c}(h) by the crossover.

Figure 4 (b) can be understood similarly. In this case, the crossover line is identified to be h0.17h\sim 0.17 Sup , where Δjc\Delta j_{c} changes its sign. The high-field helical superconducting states span only a small fraction in the phase diagram. The difference from Fig. 4 (a) is another sign reversal at h0.09h\sim 0.09. In this region, Δjc\Delta j_{c} is determined by the nonreciprocal Landau critical momentum, and qR+qL2q0q_{R}+q_{L}-2q_{0} turns out to change its sign by increasing hh. This is because qR+qL(<0)q_{R}+q_{L}(<0) shows nonmonotonic behavior, while 2q0(>0)-2q_{0}(>0) grows linearly and finally becomes dominant Sup . The sign reversal survives at higher temperatures and reaches the transition temperature.

Discussion — We have revealed the sign reversals of the SDE, which is closely connected with the change in the helical superconducting states. Thus, the intrinsic SDE is a promising bulk probe directly unveiling the crossover line. This probe is complementary to the junction Kaur et al. (2005) and spectroscopy Smidman et al. (2017) experiments proposed to detect the helical superconductivity.

In the end, we briefly discuss the connection with the experimental results of SDE Ando et al. (2020). The sign reversals of Δjc\Delta j_{c} by increasing the magnetic field at low temperatures have recently been observed Ono et al. , which might be explained by our results for the intrinsic SDE. An “inverse effect,” the nonreciprocity of the critical magnetic field under applied electric current, has also been reported Miyasaka et al. (2021), implying the nonreciprocity as a bulk property of the superconductor. Thus, the SDE with sign reversals implies the crossover in the superconducting state of the Nb/V/Ta superlattice. On the other hand, Δjc(h)\Delta j_{c}(h) near TcT_{c} seems to be at variance with the intrinsic SDE Ando et al. (2020). This point might be overcome by considering the effect of vortices, which is left as an intriguing future issue.

Acknowledgements.
We appreciate helpful discussions with T. Ono, Y. Miyasaka, R. Kawarazaki, H. Narita, and H. Watanabe. This work was supported by JSPS KAKENHI (Grants No. JP18H05227, No. JP18H01178, No. 20H05159, No. 21K13880, and No. 21J14804), JSPS research fellowship, WISE Program MEXT, and SPIRITS 2020 of Kyoto University. Note added.— During finalizing the manuscript, we became aware of independent overlapping works. A recent arXiv post by N. Yuan and L. Fu Yuan and Fu (2021) studies the depairing currrent of the Rashba-Zeeman model mainly using the GL theory. However, sign reversals of the SDE are not obtained. The work by J. He and N. Nagaosa et al. He et al. (2021) studies the related topic independently of ours. We thank J. He for coordinating submission to arXiv.

References

  • Tokura and Nagaosa (2018) Y. Tokura and N. Nagaosa, Nat. Commun. 9, 3740 (2018).
  • Ideue and Iwasa (2021) T. Ideue and Y. Iwasa, Annu. Rev. Condens. Matter Phys. 12, 201 (2021).
  • Rikken et al. (2001) G. L. Rikken, J. Fölling,  and P. Wyder, Phys. Rev. Lett. 87, 236602 (2001).
  • Krstić et al. (2002) V. Krstić, S. Roth, M. Burghard, K. Kern,  and G. L. J. A. Rikken, J. Chem. Phys. 117, 11315 (2002).
  • Pop et al. (2014) F. Pop, P. Auban-Senzier, E. Canadell, G. L. J. A. Rikken,  and N. Avarvari, Nat. Commun. 5, 3757 (2014).
  • Rikken and Wyder (2005) G. L. J. A. Rikken and P. Wyder, Phys. Rev. Lett. 94, 016601 (2005).
  • Ideue et al. (2017) T. Ideue, K. Hamamoto, S. Koshikawa, M. Ezawa, S. Shimizu, Y. Kaneko, Y. Tokura, N. Nagaosa,  and Y. Iwasa, Nat. Phys. 13, 578 (2017).
  • Wakatsuki and Nagaosa (2018) R. Wakatsuki and N. Nagaosa, Phys. Rev. Lett. 121, 026601 (2018).
  • Hoshino et al. (2018) S. Hoshino, R. Wakatsuki, K. Hamamoto,  and N. Nagaosa, Phys. Rev. B Condens. Matter 98, 054510 (2018).
  • Wakatsuki et al. (2017) R. Wakatsuki, Y. Saito, S. Hoshino, Y. M. Itahashi, T. Ideue, M. Ezawa, Y. Iwasa,  and N. Nagaosa, Science Advances 3, e1602390 (2017).
  • Qin et al. (2017) F. Qin, W. Shi, T. Ideue, M. Yoshida, A. Zak, R. Tenne, T. Kikitsu, D. Inoue, D. Hashizume,  and Y. Iwasa, Nat. Commun. 8, 14465 (2017).
  • Yasuda et al. (2019) K. Yasuda, H. Yasuda, T. Liang, R. Yoshimi, A. Tsukazaki, K. S. Takahashi, N. Nagaosa, M. Kawasaki,  and Y. Tokura, Nat. Commun. 10, 2734 (2019).
  • Itahashi et al. (2020) Y. M. Itahashi, T. Ideue, Y. Saito, S. Shimizu, T. Ouchi, T. Nojima,  and Y. Iwasa, Science advances 6, eaay9120 (2020).
  • Ando et al. (2020) F. Ando, Y. Miyasaka, T. Li, J. Ishizuka, T. Arakawa, Y. Shiota, T. Moriyama, Y. Yanase,  and T. Ono, Nature 584, 373 (2020).
  • Reynoso et al. (2008) A. A. Reynoso, G. Usaj, C. A. Balseiro, D. Feinberg,  and M. Avignon, Phys. Rev. Lett. 101, 107001 (2008).
  • Zazunov et al. (2009) A. Zazunov, R. Egger, T. Jonckheere,  and T. Martin, Phys. Rev. Lett. 103, 147004 (2009).
  • Margaris et al. (2010) I. Margaris, V. Paltoglou,  and N. Flytzanis, J. Phys. Condens. Matter 22, 445701 (2010).
  • Yokoyama et al. (2014) T. Yokoyama, M. Eto,  and Y. V. Nazarov, Phys. Rev. B Condens. Matter 89, 195407 (2014).
  • Silaev et al. (2014) M. A. Silaev, A. Y. Aladyshkin, M. V. Silaeva,  and A. S. Aladyshkina, J. Phys. Condens. Matter 26, 095702 (2014).
  • Campagnano et al. (2015) G. Campagnano, P. Lucignano, D. Giuliano,  and A. Tagliacozzo, J. Phys. Condens. Matter 27, 205301 (2015).
  • Dolcini et al. (2015) F. Dolcini, M. Houzet,  and J. S. Meyer, Phys. Rev. B Condens. Matter 92, 035428 (2015).
  • Chen et al. (2018) C.-Z. Chen, J. J. He, M. N. Ali, G.-H. Lee, K. C. Fong,  and K. T. Law, Phys. Rev. B Condens. Matter 98, 075430 (2018).
  • Minutillo et al. (2018) M. Minutillo, D. Giuliano, P. Lucignano, A. Tagliacozzo,  and G. Campagnano, Phys. Rev. B Condens. Matter 98, 144510 (2018).
  • Pal and Benjamin (2019) S. Pal and C. Benjamin, EPL 126, 57002 (2019).
  • Kopasov et al. (2021) A. A. Kopasov, A. G. Kutlin,  and A. S. Mel’nikov, Phys. Rev. B Condens. Matter 103, 144520 (2021).
  • Baumgartner et al. (2021) C. Baumgartner, L. Fuchs, A. Costa, 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, Nat. Nanotechnol. , 10.1038/s41565 (2021).
  • Lyu et al. (2021) Y.-Y. Lyu, J. Jiang, Y.-L. Wang, Z.-L. Xiao, S. Dong, Q.-H. Chen, M. V. Milošević, H. Wang, R. Divan, J. E. Pearson, P. Wu, F. M. Peeters,  and W.-K. Kwok, Nat. Commun. 12, 2703 (2021).
  • Blatter et al. (1994) G. Blatter, M. V. Feigel’man, V. B. Geshkenbein, A. I. Larkin,  and V. M. Vinokur, Rev. Mod. Phys. 66, 1125 (1994).
  • Villegas et al. (2003) J. E. Villegas, S. Savel’ev, F. Nori, E. M. Gonzalez, J. V. Anguita, R. García,  and J. L. Vicent, Science 302, 1188 (2003).
  • Tinkham (2004) M. Tinkham, Introduction to Superconductivity (Courier Corporation, 2004).
  • Dew-Hughes (2001) D. Dew-Hughes, Low Temp. Phys. 27, 713 (2001).
  • Nawaz et al. (2013) S. Nawaz, R. Arpaia, F. Lombardi,  and T. Bauch, Phys. Rev. Lett. 110, 167004 (2013).
  • Li et al. (2013) J. Li, J. Yuan, Y.-H. Yuan, J.-Y. Ge, M.-Y. Li, H.-L. Feng, P. J. Pereira, A. Ishii, T. Hatano, A. V. Silhanek, L. F. Chibotaru, J. Vanacken, K. Yamaura, H.-B. Wang, E. Takayama-Muromachi,  and V. V. Moshchalkov, Appl. Phys. Lett. 103, 062603 (2013).
  • Sun et al. (2020) Y. Sun, H. Ohnuma, S.-Y. Ayukawa, T. Noji, Y. Koike, T. Tamegai,  and H. Kitano, Phys. Rev. B 101, 134516 (2020).
  • Larkin and Ovchinnikov (1964) A. I. Larkin and Y. N. Ovchinnikov, Zh. Eksperim. i Teor. Fiz. 47, 1136 (1964).
  • Fulde and Ferrell (1964) P. Fulde and R. A. Ferrell, Phys. Rev. 135, A550 (1964).
  • Kasahara et al. (2014) S. Kasahara, T. Watashige, T. Hanaguri, Y. Kohsaka, T. Yamashita, Y. Shimoyama, Y. Mizukami, R. Endo, H. Ikeda, K. Aoyama, T. Terashima, S. Uji, T. Wolf, H. von Löhneysen, T. Shibauchi,  and Y. Matsuda, Proc. Natl. Acad. Sci. U. S. A. 111, 16309 (2014).
  • Kasahara et al. (2020) S. Kasahara, Y. Sato, S. Licciardello, M. Čulo, S. Arsenijević, T. Ottenbros, T. Tominaga, J. Böker, I. Eremin, T. Shibauchi, J. Wosnitza, N. E. Hussey,  and Y. Matsuda, Phys. Rev. Lett. 124, 107001 (2020).
  • Matsuda and Shimahara (2007) Y. Matsuda and H. Shimahara, J. Phys. Soc. Jpn. 76, 051005 (2007).
  • Wosnitza (2018) J. Wosnitza, Ann. Phys. 530, 1700282 (2018).
  • Bauer and Sigrist (2012) E. Bauer and M. Sigrist, Non-Centrosymmetric Superconductors: Introduction and Overview (Springer Science & Business Media, 2012).
  • Smidman et al. (2017) M. Smidman, M. B. Salamon, H. Q. Yuan,  and D. F. Agterberg, Rep. Prog. Phys. 80, 036501 (2017).
  • Agterberg (2003) D. F. Agterberg, Physica C Supercond. 387, 13 (2003).
  • Barzykin and Gor’kov (2002) V. Barzykin and L. P. Gor’kov, Phys. Rev. Lett. 89, 227002 (2002).
  • Dimitrova and Feigel’man (2003) O. V. Dimitrova and M. V. Feigel’man, JETP Lett. 78, 637 (2003).
  • Kaur et al. (2005) R. P. Kaur, D. F. Agterberg,  and M. Sigrist, Phys. Rev. Lett. 94, 137002 (2005).
  • Agterberg and Kaur (2007) D. F. Agterberg and R. P. Kaur, Phys. Rev. B Condens. Matter 75, 064511 (2007).
  • Dimitrova and Feigel’man (2007) O. Dimitrova and M. V. Feigel’man, Phys. Rev. B Condens. Matter 76, 014522 (2007).
  • Samokhin (2008) K. V. Samokhin, Phys. Rev. B Condens. Matter 78, 224520 (2008).
  • Yanase and Sigrist (2008) Y. Yanase and M. Sigrist, J. Phys. Soc. Jpn. 77, 342 (2008).
  • Michaeli et al. (2012) K. Michaeli, A. C. Potter,  and P. A. Lee, Phys. Rev. Lett. 108, 117003 (2012).
  • Sekihara et al. (2013) T. Sekihara, R. Masutomi,  and T. Okamoto, Phys. Rev. Lett. 111, 057005 (2013).
  • Houzet and Meyer (2015) M. Houzet and J. S. Meyer, Phys. Rev. B Condens. Matter 92, 014509 (2015).
  • Schumann et al. (2020) T. Schumann, L. Galletti, H. Jeong, K. Ahadi, W. M. Strickland, S. Salmani-Rezaie,  and S. Stemmer, Phys. Rev. B Condens. Matter 101, 100503 (2020).
  • Naritsuka et al. (2017) M. Naritsuka, T. Ishii, S. Miyake, Y. Tokiwa, R. Toda, M. Shimozawa, T. Terashima, T. Shibauchi, Y. Matsuda,  and Y. Kasahara, Phys. Rev. B Condens. Matter 96, 174512 (2017).
  • Naritsuka et al. (2021) M. Naritsuka, T. Terashima,  and Y. Matsuda, J. Phys. Condens. Matter 33, 273001 (2021).
  • (57) T. Ono, Y. Miyasaka,  and R. Kawarazaki, Private communication.
  • (58) See Supplemental Material for more details.
  • Note (1) This terminology is named after the crossover generally seen in noncentrosymmetric superconductors, while the line changes to the first-order transition at lower temperature T0.12T\lesssim 0.12 in this model, as in Figs.5(b)-5(d).
  • Miyasaka et al. (2021) Y. Miyasaka, R. Kawarazaki, H. Narita, F. Ando, Y. Ikeda, R. Hisatomi, A. Daido, Y. Shiota, T. Moriyama, Y. Yanase,  and T. Ono, Appl. Phys. Express 14, 073003 (2021).
  • Yuan and Fu (2021) N. F. Q. Yuan and L. Fu,   (2021), arXiv:2106.01909 [cond-mat.supr-con] .
  • He et al. (2021) J. J. He, Y. Tanaka,  and N. Nagaosa, “A phenomenological theory of superconductor diodes in presence of magnetochiral anisotropy,”  (2021), arXiv:2106.03575 [cond-mat.supr-con] .

I Formulation to evaluate the electric current

Here, we show the details of the formulation to calculate the current expectation values in the superconducting states. The free energy per unit volume in the superconducting state is given by

Ω(Δ,q)\displaystyle\Omega(\Delta,q) TVlnTreH^MFq(Δ)/T\displaystyle\equiv-\frac{T}{V}\ln\mathrm{Tr}\,e^{-\hat{H}_{\text{MF}}^{q}(\Delta)/T}
=12V𝒌trN[Δ2U+HN(𝒌)]\displaystyle=\frac{1}{2V}\sum_{\bm{k}}\mathrm{tr}\,_{N}\left[\frac{\Delta^{2}}{U}+H_{N}(\bm{k})\right]
T2V𝒌tr[ln(1+eH(𝒌,q)/T)].\displaystyle\qquad-\frac{T}{2V}\sum_{\bm{k}}\mathrm{tr}\,\left[\ln(1+e^{-H(\bm{k},q)/T})\right]. (6)

Here, trN\mathrm{tr}\,_{N} represents the trace over the spin degrees of freedom, while tr\mathrm{tr}\, represents that over both the spin and the Nambu degrees of freedom. V=LxLyV=L_{x}L_{y} represents the system size with LiL_{i} the diameter in the i=x,yi=x,y direction. We introduced the Bogoliubov-de Genens (BdG) Hamiltonian H(𝒌,q)H(\bm{k},q) by

H^MFq(Δ)=12𝒌Ψ(𝒌,q)H(𝒌,q)Ψ(𝒌,q)+const.,\displaystyle\hat{H}_{\text{MF}}^{q}(\Delta)=\frac{1}{2}\sum_{\bm{k}}\Psi^{\dagger}(\bm{k},q)H(\bm{k},q)\Psi(\bm{k},q)+\text{const.}, (7)
H(𝒌,q)=(HN(𝒌+𝒒)ΔiσyΔiσyHNT(𝒌)),\displaystyle H(\bm{k},q)=\begin{pmatrix}H_{N}(\bm{k}+\bm{q})&\Delta i\sigma_{y}\\ -\Delta i\sigma_{y}&-H_{N}^{T}(-\bm{k})\end{pmatrix}, (8)

with the momentum 𝒒=qx^\bm{q}=q\hat{x} and the Nambu spinor Ψ(𝒌,q)=(c𝒌+𝒒,c𝒌+𝒒,c𝒌,c𝒌)\Psi(\bm{k},q)^{\dagger}=(c^{\dagger}_{\bm{k}+\bm{q}\uparrow},c^{\dagger}_{\bm{k}+\bm{q}\downarrow},c_{-\bm{k}\uparrow},c_{-\bm{k}\downarrow}). Here, we choose qq to be compatible with the periodic boundary conditions, q2π/Lxq\in 2\pi\mathbb{Z}/L_{x}. The constant term in Eq. (7) is equivalent to the first term of Eq. (6). The normal-state Bloch Hamiltonian is given by HN(𝒌)=ξ(𝒌)+(𝒈(𝒌)𝒉)𝝈H_{N}(\bm{k})=\xi(\bm{k})+(\bm{g}(\bm{k})-\bm{h})\cdot\bm{\sigma}.

The electric current (the sheet current density) is defined by

j(Δ,q)=tr[j^xeH^MFq(Δ)/T]tr[eH^MFq(Δ)/T].\displaystyle j(\Delta,q)=\frac{\mathrm{tr}\,[\hat{j}_{x}e^{-\hat{H}^{q}_{\text{MF}}(\Delta)/T}]}{\mathrm{tr}\,[e^{-\hat{H}^{q}_{\text{MF}}(\Delta)/T}]}. (9)

Here, the current operator is given by

j^x\displaystyle\hat{j}_{x} =1V𝒌σσkxHN(𝒌)σσc𝒌σc𝒌σ\displaystyle=\frac{1}{V}\sum_{\bm{k}\sigma\sigma^{\prime}}\partial_{k_{x}}H_{N}(\bm{k})_{\sigma\sigma^{\prime}}c^{\dagger}_{\bm{k}\sigma}c_{\bm{k}\sigma^{\prime}} (10)
=1V𝒌Ψ(𝒌,q)qH(𝒌,q)Ψ(𝒌,q).\displaystyle=\frac{1}{V}\sum_{\bm{k}}\Psi^{\dagger}(\bm{k},q)\partial_{q}H(\bm{k},q)\Psi(\bm{k},q). (11)

After some calculations, we obtain

j(Δ,q)\displaystyle j(\Delta,q) =1V𝒌tr[qH(𝒌,q)f(H(𝒌,q))]\displaystyle=\frac{1}{V}\sum_{\bm{k}}\mathrm{tr}\,[\partial_{q}H(\bm{k},q)f(H(\bm{k},q))] (12)
=2qΩ(Δ,q),\displaystyle=2\partial_{q}\Omega(\Delta,q), (13)

with the Fermi distribution function f(ϵ)=(eϵ/T+1)1f(\epsilon)=(e^{\epsilon/T}+1)^{-1}.

The gap equation is given by

ΔΩ(Δ,q)=0,\partial_{\Delta}\Omega(\Delta,q)=0, (14)

which determines the pair potential Δ\Delta self-consistently. The solution is written as Δ(q)\Delta(q), and satisfies Eq. (14), or equivalently,

Δ(q)=UVkckck+q|Δ=Δ(q).\Delta(q)=-\frac{U}{V}\sum_{k}\braket{c_{-k\downarrow}c_{k+q\uparrow}}|_{\Delta=\Delta(q)}. (15)

By using Δ(q)\Delta(q), the condensation energy defined in the main text is written as

F(q)=Ω(Δ(q),q)Ω(0,q),F(q)=\Omega(\Delta(q),q)-\Omega(0,q), (16)

where Ω(0,q)=Ω(0,0)\Omega(0,q)=\Omega(0,0) holds as is easily confirmed with Eqs. (6) and (8). By using Eqs. (13) and (14), we obtain

2qF(q)\displaystyle 2\partial_{q}F(q) =limΔΔ(q)[2qΩ(Δ,q)+2qΔ(q)ΔΩ(Δ,q)]\displaystyle=\lim_{\Delta\to\Delta(q)}\bigl{[}2\partial_{q}\Omega(\Delta,q)+2\partial_{q}\Delta(q)\,\partial_{\Delta}\Omega(\Delta,q)\bigr{]}
=j(Δ(q),q)\displaystyle=j(\Delta(q),q)
j(q).\displaystyle\equiv j(q). (17)

The obtained equality goes along with the standard expression j=AΩj=-\partial_{A}\Omega with AA the uniform vector potential since qq changes by 2δA-2\delta A when AA changes by δA\delta A.

II GL analysis

Here we show the details of the GL analysis of the superconducting diode effect. Let us start from the expression

f(Δ,q)=α(q)Δ2+β(q)2Δ4,f(\Delta,q)=\alpha(q)\Delta^{2}+\frac{\beta(q)}{2}\Delta^{4}, (18)

keeping the order parameter of the form Δ(x)eiqx\Delta(x)\propto e^{iqx} in mind. The coefficients are given by

α(q)=α~0+α~22(qq0)2+α36(qq0)3,\displaystyle\alpha(q)=-{\tilde{\alpha}}_{0}+\frac{{\tilde{\alpha}}_{2}}{2}(q-q_{0})^{2}+\frac{\alpha_{3}}{6}(q-q_{0})^{3},
β(q)=β~0+β1(qq0).\displaystyle\beta(q)={\tilde{\beta}}_{0}+\beta_{1}(q-q_{0}). (19)

Here we omit the tilde of q~0\tilde{q}_{0} in the main text for simplicity, and redefine α~0α~0{\tilde{\alpha}}_{0}\to-{\tilde{\alpha}}_{0}. The order parameter is optimized by

f/Δ2=α(q)+β(q)Δ2=0.\partial f/\partial{\Delta^{2}}=\alpha(q)+\beta(q)\Delta^{2}=0. (20)

Assuming β(q)>0\beta(q)>0 for the range of qq we are interested in, Δ\Delta has a nontrivial real solution only when α(q)<0\alpha(q)<0. Thus, the GL free energy is given by

f(q)=f(Δ(q),q)=α(q)22β(q)θ(α(q)),f(q)=f(\Delta(q),q)=-\frac{\alpha(q)^{2}}{2\beta(q)}\theta(-\alpha(q)), (21)

with θ(x)\theta(x) the Heaviside step function. Since the minimum of α(q)\alpha(q) is α~0-{\tilde{\alpha}}_{0}, the transition from the normal to helical superconducting state occurs when the sign of α~0{\tilde{\alpha}}_{0} changes from negative to positive as lowering the temperature. Thus, we conclude α~0TcT{\tilde{\alpha}}_{0}\propto T_{c}-T.

We first consider the case α3=β1=0\alpha_{3}=\beta_{1}=0. The supercurrent is given by j(q)=2qf(q)j(q)=2\partial_{q}f(q),

β~0j(q)/2=qα(q)2/2=α(q)qα(q).{\tilde{\beta}}_{0}j(q)/2=-\partial_{q}\alpha(q)^{2}/2=-\alpha(q)\partial_{q}\alpha(q). (22)

This is an odd function of qq0q-q_{0}, and thus the critical current is reciprocal. Actually, The maximum and minimum of j(q)j(q) are achieved at q=qcq=q_{c} satisfying

0\displaystyle 0 =[qcβ~0j(qc)/2]α3=β1=0\displaystyle=[\partial_{q_{c}}{\tilde{\beta}}_{0}j(q_{c})/2]_{\alpha_{3}=\beta_{1}=0}
=3α~222(2α~03α~2(qcq0)2).\displaystyle=\frac{3{\tilde{\alpha}}_{2}^{2}}{2}\left(\frac{2{\tilde{\alpha}}_{0}}{3{\tilde{\alpha}}_{2}}-(q_{c}-q_{0})^{2}\right). (23)

Accordingly, |qcq0||q_{c}-q_{0}| scales as TcT\sqrt{T_{c}-T}, as is the inverse of the coherence length. Thus, the reciprocal critical current is given by

[jc+]α3=β1=0\displaystyle[j_{c+}]_{\alpha_{3}=\beta_{1}=0} =[jc]α3=β1=0\displaystyle=[-j_{c-}]_{\alpha_{3}=\beta_{1}=0}
=46α~29β~0α~03/2.\displaystyle=\frac{4\sqrt{6{\tilde{\alpha}}_{2}}}{9{\tilde{\beta}}_{0}}{\tilde{\alpha}}_{0}^{3/2}. (24)

Note that this coincides with j¯c\bar{j}_{c} up to first order in α3\alpha_{3} and β1\beta_{1}. Thus, the well-known scaling law jc(TcT)3/2j_{c}\sim(T_{c}-T)^{3/2} is reproduced for j¯c\bar{j}_{c}.

Let us consider the first-order change caused by α3\alpha_{3} and β1\beta_{1}. We obtain

α3[β~0j(q)/2]|α3=β1=0\displaystyle\partial_{\alpha_{3}}[{\tilde{\beta}}_{0}j(q)/2]\Bigr{|}_{\alpha_{3}=\beta_{1}=0}
=5α~2(qq0)212(6α~05α~2(qq0)2),\displaystyle=\frac{5{\tilde{\alpha}}_{2}(q-q_{0})^{2}}{12}\left(\frac{6{\tilde{\alpha}}_{0}}{5{\tilde{\alpha}}_{2}}-(q-q_{0})^{2}\right), (25)

and

β1[β~0j(q)/2]|α3=β1=0\displaystyle\partial_{\beta_{1}}[{\tilde{\beta}}_{0}j(q)/2]\Bigr{|}_{\alpha_{3}=\beta_{1}=0}
=5α~228β~0((qq0)22α~0α~2)((qq0)22α~05α~2).\displaystyle=\frac{5{\tilde{\alpha}}_{2}^{2}}{8{\tilde{\beta}}_{0}}\left((q-q_{0})^{2}-\frac{2{\tilde{\alpha}}_{0}}{{\tilde{\alpha}}_{2}}\right)\left((q-q_{0})^{2}-\frac{2{\tilde{\alpha}}_{0}}{5{\tilde{\alpha}}_{2}}\right). (26)

When the critical current jc+j_{c+} is realized at qc+=qc+δqcq_{c+}=q_{c}+\delta q_{c}, we obtain up to first order in α3\alpha_{3} and β1\beta_{1},

j(qc+)\displaystyle j(q_{c+}) =[j+δj](qc+δqc)\displaystyle=[j+\delta j](q_{c}+\delta q_{c})
=jc0+α3[α3j(qc)]α3=β1=0\displaystyle=j_{c0}+\alpha_{3}[\partial_{\alpha_{3}}j(q_{c})]_{\alpha_{3}=\beta_{1}=0}
+β1[β1j(qc)]α3=β1=0\displaystyle\quad+\beta_{1}[\partial_{\beta_{1}}j(q_{c})]_{\alpha_{3}=\beta_{1}=0}
+[qcj(qc)]α3=β1=0δqc\displaystyle\quad+[\partial_{q_{c}}j(q_{c})]_{\alpha_{3}=\beta_{1}=0}\delta q_{c}
=jc0+α3[α3j(qc)]α3=β1=0\displaystyle=j_{c0}+\alpha_{3}[\partial_{\alpha_{3}}j(q_{c})]_{\alpha_{3}=\beta_{1}=0}
+β1[β1j(qc)]α3=β1=0.\displaystyle\quad+\beta_{1}[\partial_{\beta_{1}}j(q_{c})]_{\alpha_{3}=\beta_{1}=0}. (27)

Here, jc0j_{c0} represents [j(qc)]α3=β1=0[j(q_{c})]_{\alpha_{3}=\beta_{1}=0}. Thus, we obtain the nonreciprocal component of the critical current,

Δjc\displaystyle\Delta j_{c} =2α3[α3j(qc)]α3=β1=0+2β1[β1j(qc)]α3=β1=0\displaystyle=2\alpha_{3}[\partial_{\alpha_{3}}j(q_{c})]_{\alpha_{3}=\beta_{1}=0}+2\beta_{1}[\partial_{\beta_{1}}j(q_{c})]_{\alpha_{3}=\beta_{1}=0}
=16α~0227β~0α~2α38α~029β~02β1.\displaystyle=\frac{16{\tilde{\alpha}}_{0}^{2}}{27{\tilde{\beta}}_{0}{\tilde{\alpha}}_{2}}\alpha_{3}-\frac{8{\tilde{\alpha}}_{0}^{2}}{9{\tilde{\beta}}_{0}^{2}}\beta_{1}. (28)

This scales as Δjc(TcT)2\Delta j_{c}\sim(T_{c}-T)^{2}.

(a) (b) (c) (d)
Refer to caption Refer to caption Refer to caption Refer to caption
Figure 6: Temperature and magnetic-field dependence of (a),(b) ΔE\Delta E, (c) q0q_{0}, and (d) rr. t2=0t_{2}=0 for the panel (a), while t2=0.2t_{2}=0.2 for the panels (b)-(d).

III Calculation details and Phase diagrams

Here we explain the details of the numerical calculations and show some additional figures related to the phase diagrams. All the calculations for the figures in the main text and those presented here are done with Lx=6000L_{x}=6000 and Ly=200L_{y}=200. Exceptionally, we adopt Lx=12000L_{x}=12000 for Fig. 5 and T>0.03T>0.03 in Fig. 2 to reduce the finite-size effect. To obtain jc±j_{c\pm}, j(q)j(q) is maximized/minimized among q2π/Lxq\in 2\pi\mathbb{Z}/L_{x}. The normalization of Figs. 5 (a)-(d) is done with max[jc+,|jc|]\max[j_{c+},|j_{c-}|], |minqF(q)||\min_{q}F(q)|, and max[0.003,maxqΔE(q)]\max[0.003,\max_{q}\Delta E(q)] for j(q)j(q), F(q)F(q), and ΔE(q)\Delta E(q), respectively. Note that we show in Figs. 5 (a)-(d) only the most stable state that minimizes F(q)F(q) for each qq. In particular, there is a metastable superconducting state for smaller (larger) qq’s of the left (right) peaks in Fig. 5 (d). The supercurrent sustained by these states might be observed when the experimental time scale is small. For Figs. 5 (a)-(d), it is confirmed that the superconducting solution is (if any) unique for each qq.

In Figs. 6 (a) and (b), we show the temperature and magnetic-filed dependence of the minimum excitation energy ΔE(q0)\Delta E(q_{0}) for t2=0t_{2}=0 and 0.20.2, respectively. The spectrum becomes gapless in the high-field helical superconducting states, which can be detected by scanning tunneling microscopy. Figures 6 (c) and (d) show q0q_{0} and rr for t2=0.2t_{2}=0.2. In Figs. 6 (c) and (d), the crossover line is seen to be h0.17h\sim 0.17, and a huge nonreciprocal nature r0.8r\sim 0.8 is observed. To be precise, Δjc\Delta j_{c} becomes positive in a tiny region near h0.19h\sim 0.19 and T0T\sim 0 for t2=0.2t_{2}=0.2. However, this is probably due to the peculiarity of the model, where the Lifshitz transition of the outer Fermi surface occurs around h0.185h\sim 0.185.

IV Evolution of Landau critical momenta

As discussed in the main text, the sign reversal of Δjc\Delta j_{c} by t2t_{2} can be understood based on the nonreciprocity of the Landau critical momenta. In Fig. 7, we show qR+qLq_{R}+q_{L} and q0q_{0} obtained from j(q)j(q) at h=0.03h=0.03 and T=0.001T=0.001, with varying t2t_{2} from 0 to 0.20.2. The sign reversal of qR+qL2q0q_{R}+q_{L}-2q_{0} (red closed circles) naturally explains that of Δjc\Delta j_{c}. For large values of t2t_{2}, qR+qL2q0q_{R}+q_{L}-2q_{0} is dominated by the nonreciprocal Landau critical momentum qR+qLq_{R}+q_{L} (blue closed triangles), while it is dominated by q0q_{0} (black closed squares) for small values of t2t_{2}. It should be noted that a large SDE is obtained for t20.2t_{2}\sim 0.2, although qR+qLq_{R}+q_{L} and 2q0-2q_{0} contribute destructively.

To understand the behavior of qR+qLq_{R}+q_{L}, we discuss the Landau critical momenta with the help of the single-band formula |𝒒𝒗/2|=Δ|\bm{q}\cdot\bm{v}/2|=\Delta. In our case, 𝒒\bm{q} should be replaced with 𝒒𝒒χ(𝒌)\bm{q}-\bm{q}_{\chi}(\bm{k}) for each band, and we obtain,

Δ\displaystyle\Delta =maxχ=±,𝒌=𝒌Fχ|qx^𝒒χ(𝒌)2𝒗χ(𝒌)|\displaystyle=\max_{\chi=\pm,\ \bm{k}=\bm{k}_{F}^{\chi}}\left|\frac{q\,\hat{x}-\bm{q}_{\chi}(\bm{k})}{2}\cdot\bm{v}_{\chi}(\bm{k})\right|
=maxχ=±,𝒌=𝒌Fχ|qvχ,x(𝒌)/2χhg^y(𝒌)|,\displaystyle=\max_{\chi=\pm,\ \bm{k}=\bm{k}_{F}^{\chi}}\left|qv_{\chi,x}(\bm{k})/2-\chi h\hat{g}_{y}(\bm{k})\right|, (29)

whose positive and negative solutions for qq correspond to qRq_{R} and qLq_{L}, respectively. Here, 𝒌Fχ\bm{k}_{F}^{\chi} specifies the 𝒌\bm{k} points on the Fermi surface with the helicity χ\chi, while g^(𝒌)\hat{g}(\bm{k}) is the unit vector parallel to 𝒈(𝒌)\bm{g}(\bm{k}). Equation (29) well reproduces the result for qR+qLq_{R}+q_{L}, as shown by skyblue open triangles in Fig. 7. To go further, let us simplify the expression by replacing 𝒒χ(𝒌)\bm{q}_{\chi}(\bm{k}) in the first line of Eq. (29) with its average on the Fermi surface: qχx^𝒒χ(𝒌F,χ)q_{\chi}\hat{x}\equiv\braket{\bm{q}_{\chi}(\bm{k}_{F,\chi})}. We obtain for h>0h>0 [see the next section for the derivation],

qR+qL\displaystyle q_{R}+q_{L} ={q++q+2(ΔvΔv+),(|δv|/v¯h/Δ)2q+,(δv/v¯h/Δ)2q,(δv/v¯h/Δ)\displaystyle=\begin{cases}\begin{array}[]{l}q_{+}+q_{-}\\ \ +2\left(\frac{\Delta}{v_{-}}-\frac{\Delta}{v_{+}}\right),\end{array}&(|\delta v|/\bar{v}\lesssim h/\Delta)\\ 2q_{+},&(\delta v/\bar{v}\gtrsim h/\Delta)\\ 2q_{-},&(\delta v/\bar{v}\lesssim-h/\Delta)\end{cases} (30)

whose helicities are interchanged for h<0h<0. Here, we defined vχmax𝒌Fχvχ,x(𝒌Fχ)v_{\chi}\equiv\max_{\bm{k}_{F}^{\chi}}v_{\chi,x}(\bm{k}_{F}^{\chi}), δv=v+v\delta v=v_{+}-v_{-} and v¯(v++v)/2\bar{v}\sim(v_{+}+v_{-})/2. Equation (30) qualitatively agrees with qR+qLq_{R}+q_{L} for t20.15t_{2}\lesssim 0.15, as shown by the open purple inverted triangles in Fig. 7. In this regime, we have small δv\delta v and the first line of Eq. (30) is applied. Since q±±2h/v±q_{\pm}\sim\pm 2h/v_{\pm}, the difference of the Fermi velocities δv\delta v plays a key role to obtain a large qR+qLq_{R}+q_{L}. It is expected that the anisotropy of the system is advantageous to obtain a large value of δv\delta v. On the other hand, Eq. (30) underestimates qR+qLq_{R}+q_{L} around t2=0.2t_{2}=0.2, where the third line of Eq. (30) is applied. This indicates that the isotropic simplification 𝒒χ(𝒌)qχ\bm{q}_{\chi}(\bm{k})\to q_{\chi} is not valid for strongly anisotropic systems with large t2t_{2}. Thus, overall, large anisotropy of the system is expected to be the key to obtain a large SDE.

Refer to caption
Figure 7: t2t_{2} dependence of qR+qLq_{R}+q_{L}, q0q_{0}, and their combinations. The red closed circles, blue closed triangles, black closed squares indicate qR+qL2q0q_{R}+q_{L}-2q_{0}, qR+qLq_{R}+q_{L}, q0q_{0} evaluated from j(q)j(q), respectively. The open sky-blue triangles and open purple inverted triangles indicate qR+qLq_{R}+q_{L} calculated from Eq. (29) and that with the isotropic simplification 𝒒χ(𝒌)qχx^\bm{q}_{\chi}(\bm{k})\to q_{\chi}\hat{x}, respectively.
Refer to caption
Figure 8: Magnetic field dependence of qR+qLq_{R}+q_{L}, q0q_{0} and their combinations for t2=0.2t_{2}=0.2 and T=0.001T=0.001. The notations are the same as Fig. 7.

In Fig. 8, we show the magnetic-field dependence of qR+qLq_{R}+q_{L}, q0q_{0} and their combinations for t2=0.2t_{2}=0.2. The notations are the same as those of Fig. 7. While qR+qLq_{R}+q_{L} is nonmonotonic, 2q0-2q_{0} grows linearly and finally the sign reversal of qR+qL2q0q_{R}+q_{L}-2q_{0} occurs. “qR+qLq_{R}+q_{L} (aniso)”, i.e. Eq. (29), qualitatively captures the behavior of qR+qLq_{R}+q_{L}, whose slight deviation is probably due to the higher-order corrections of hh. The isotropic simplification does not work for t2=0.2t_{2}=0.2 as is clear in Fig. 8. The sign reversal of qR+qL2q0q_{R}+q_{L}-2q_{0} is the origin of that of Δjc\Delta j_{c} in the phase diagram for t2=0.2t_{2}=0.2 under moderate magnetic fields.

IV.1 Derivation of Eq. (30)

Here, we derive Eq. (30). By using 𝒒χ(𝒌)qχx^\bm{q}_{\chi}(\bm{k})\to q_{\chi}\hat{x}, we obtain

2Δ\displaystyle 2\Delta =maxχ=±[|qqχ|max𝒌Fχ|vχ,x(𝒌)|]\displaystyle=\max_{\chi=\pm}\Bigl{[}|q-q_{\chi}|\max_{\bm{k}_{F}^{\chi}}|v_{\chi,x}(\bm{k})|\Bigr{]} (31)
=maxχ=±[|qqχ|vχ].\displaystyle=\max_{\chi=\pm}\Bigl{[}|q-q_{\chi}|v_{\chi}\Bigr{]}. (32)

Let us first consider the positive solution q=qR>0q=q_{R}>0. We also fix h>0h>0. Then, we obtain

2Δ\displaystyle 2\Delta =max[(qR+|q|)v,(qR|q+|)v+].\displaystyle=\max\Bigl{[}(q_{R}+|q_{-}|)v_{-},\,(q_{R}-|q_{+}|)v_{+}\Bigr{]}. (33)

When (qR+|q|)v>(qR|q+|)v+(q_{R}+|q_{-}|)v_{-}>(q_{R}-|q_{+}|)v_{+}, we obtain

qR=q+2Δv.q_{R}=q_{-}+\frac{2\Delta}{v_{-}}. (34)

The consistency can be checked as follows. The above inequality reads

δvqR<|q|v+|q+|v+2h.\displaystyle\delta vq_{R}<|q_{-}|v_{-}+|q_{+}|v_{+}\sim 2h. (35)

Thus, this solution is valid for

2hδv2(Δh)v2δv(Δh)v¯.2h\gtrsim\delta v\frac{2(\Delta-h)}{v_{-}}\sim\frac{2\delta v(\Delta-h)}{\bar{v}}. (36)

Considering only the linear dependence in hh, we obtain

qR=q+2Δv,δvv¯hΔ.q_{R}=q_{-}+\frac{2\Delta}{v_{-}},\quad\frac{\delta v}{\bar{v}}\lesssim\frac{h}{\Delta}. (37)

In the same way, we obtain

qR=q++2Δv+,q_{R}=q_{+}+\frac{2\Delta}{v_{+}}, (38)

for δvqR2h\delta vq_{R}\gtrsim 2h, i.e. δv/v¯h/Δ\delta v/\bar{v}\gtrsim h/\Delta. The negative solutions q=qL<0q=q_{L}<0 are obtained as follows:

qL={q+2Δv+,δvv¯hΔq2Δv.δvv¯hΔq_{L}=\begin{cases}q_{+}-\frac{2\Delta}{v_{+}},&\frac{\delta v}{\bar{v}}\gtrsim\frac{-h}{\Delta}\\ q_{-}-\frac{2\Delta}{v_{-}}.&\frac{\delta v}{\bar{v}}\lesssim\frac{-h}{\Delta}\end{cases} (39)

Summing up qRq_{R} and qLq_{L}, we obtain Eq. (30).