Introduction

Nonreciprocal transport properties1 near or inside the superconducting phase of electronic systems have been attracting a lot of research attention recently. It may manifest itself in nonreciprocal paraconductivity (NPC)2,3,4,5,6,7,8 or in so-called supercurrent diode effect (SDE)9,10.

In superconductors (SCs) or Josephson junctions with broken inversion (\({{{{{{{\mathcal{P}}}}}}}}\)) and time-reversal (\({{{{{{{\mathcal{T}}}}}}}}\)) symmetries, the critical currents along opposite directions, Jc±, may be unequal, leading to the SDE. This effect has been found in various experimental systems10,11,12,13,14,15,16,17,18, part of which may be understood by combining spin-orbit coupling (SOC) and Zeeman field19,20,21, which break \({{{{{{{\mathcal{P}}}}}}}}\) and \({{{{{{{\mathcal{T}}}}}}}}\), respectively. The SOC-Zeeman mechanism also works in one-dimension22,23 and in systems with disorders24. Supercurrent interferometers may also give rise to the SDE25,26,27, where the fractional Josephson effect of Majorana fermions can play a crucial role25. There also exist theories that consider symmetry breakings by internal magnetic28,29,30,31,32, electric33,34 or valley35 orders, finite momentum pairing36,37, unconventional superconductivity38,39, etc. However, systems with magnetic orders may be understood in a way similar to those under Zeeman fields, and superconductors with spontaneous \({{{{{{{\mathcal{P}}}}}}}}\)- or \({{{{{{{\mathcal{T}}}}}}}}\)-breaking pairing are not conveniently found in nature. Thus, it remains an open question whether there exist a new mechanism to generate the SDE in state-of-the-art experimental systems. Finding such a mechanism shall greatly enrich the choice of platforms to investigate the SDE and promote the research in this direction.

While the SDE is a manifestation of a nonreciprocal SC below its transition temperature Tc, the nonreciprocity can also be seen slightly above Tc, where Cooper pairs start to form but coherent superconductivity is not reached yet. In this regime, the trend of forming Cooper pairs makes a large contribution to the conductivity, which is called the paraconductivity40,41. In systems where \({{{{{{{\mathcal{P}}}}}}}}\) and \({{{{{{{\mathcal{T}}}}}}}}\) are broken, the paraconductivity in opposite directions may differ significantly, leading to the NPC. Although nonreciprocal conductance may also exist in the normal state at TTc, this effect can be enhanced by several orders of magnitude as the temperature approaches Tc4. Theories have shown that the NPC can also originate from a combination of SOC and Zeeman field2,4. Despite the similarity in the conditions to realize SDE and the NPC, current theories have not discussed the two in the same framework to the best of our knowledge.

In the research works on nonreciprocal transport phenomena in superconductors, both the understanding of current experimental results and the proposals of future platforms focus on systems with magnetization or spin-orbit coupling. A mechanism of generating SDE or NPC in non-magnetic materials without spin-orbit coupling remained elusive.

Here, we reveal such a mechanism with the chiral structure being the key element and predict nanotubes as realistic experimental platforms. We show that both the SDE and the NPC exist in a chiral nanotube under a magnetic field along its axial direction, and they can be obtained with the same generalized Ginzburg-Landau theory. The inversion symmetry is broken by the chiral structure of the nanotube without any SOC, and the magnetic field plays its role through the orbital effect, i.e., Aharonov-Bohm effect, instead of the Zeeman coupling. The resulting nonreciprocal signals strongly depend on the magnetic flux, the nanotube radius, and the chiral angle. There exist a periodicity in the magnetic flux through the tube, similar to the Little-Parks oscillation42, as well as a periodicity in the chiral angle. The interplay of the magnetic flux and the chiral structure is the origin of both the SDE and the NPC.

The NPC in nanotubes has been observed by Qin et al. in ref. 3 where the nanotubes are formed by transition metal dichalcogenides WS2. A strong SOC exists in this material which may also contribute to the NPC. Our theory is useful to clarify the origin of the observed NPC in ref. 3 and, on the other hand, shows the existence of SDE in chiral structures without SOC. While helping to understand the existing experimental results, this study also serves as a basis for future material choice. Its unified picture of non-reciprocal transport phenomena below and above the superconductivity transition temperature Tc shall be beneficial to the research in both regimes.

Results

Chiral nanotubes near T c

A nanotube near its superconductivity transition temperature Tc may be described by the following free energy,

$$F=\int\,{d}^{2}{{{{{{{\bf{r}}}}}}}}{\psi }^{*}({{{{{{{\bf{r}}}}}}}})\left[\alpha+\xi (\hat{{{{{{{{\bf{p}}}}}}}}})+\frac{\beta }{2}|\psi \left({{{{{{{\bf{r}}}}}}}}\right){|}^{2}\right]\psi ({{{{{{{\bf{r}}}}}}}}),$$
(1)

where α ~ T − Tc and β are the conventional Ginzburg-Landau parameters. The displacement vector r = (x, y) is defined so that the nanotube aligns alone the x-direction and the transverse coordinate y circulates around the tube, as illustrated in Fig. 1. The term \(\xi (\hat{{{{{{{{\bf{p}}}}}}}}})={\sum }_{ij}{\xi }_{ij}{\hat{p}}_{x}^{i}{\hat{p}}_{y}^{j}\) is the kinetic energy of a Cooper pair. Apparently, a periodic boundary condition should be applied along the y-direction. The momentum operator is \(\hat{{{{{{{{\bf{p}}}}}}}}}=-i\hslash {\nabla }_{{{{{{{{\bf{r}}}}}}}}}+2e{{{{{{{\bf{A}}}}}}}}({{{{{{{\bf{r}}}}}}}})\). Considering a uniform magnetic field applied along the x-direction, i.e \({{{{{{{\bf{H}}}}}}}}={H}_{x}\hat{{{{{{{{\bf{x}}}}}}}}}\), and assuming the nanotube wall thickness to be negligible, the vector potential becomes \({{{{{{{\bf{A}}}}}}}}=\frac{\phi }{2\pi R}\hat{{{{{{{{\bf{y}}}}}}}}}\), where ϕ = πR2Hx is the magnetic flux through the nanotube and R is its radius. This is equivalent to a boundary condition \(\psi ({{{{{{{\bf{r}}}}}}}})=\psi ({{{{{{{\bf{r}}}}}}}}+2\pi R\hat{{{{{{{{\bf{y}}}}}}}}})\exp \{-2\pi i\phi /{\phi }_{0}\},{\phi }_{0}=h/2e\) being the magnetic flux quantum.

Fig. 1: A schematic of a chiral nanotube formed by rolling a two-dimensional sheet.
figure 1

The two coordinate systems, (x0, y0) and (x, y), are connected by a rotation of the chiral angle θ. A magnetic field H is applied along the tube to generate nonreciprocal effects.

A Fourier transformation (taking into account the magnetic flux) leads to the following equivalent form of Eq. (1),

$$F=2\pi R\mathop{\sum}\limits_{n}\int\,dq\left[\alpha+\xi \left({{{{{{{\bf{p}}}}}}}}\right)+\frac{\beta }{2}{(2\pi R)}^{2}|{\psi }_{n}{|}^{2}\right]|{\psi }_{n}{|}^{2},$$
(2)

where q is the wavenumber along the tube and p = (q, [n − ϕ/ϕ0]/R). The integer n labels the transverse Fourier components. It is quantized due to the small circumference of the tube. We have neglected the coupling between different q-components in the ψ4 term, which does not affect the results of this study. It is clear from Eq. (2) that F is a periodic function of ϕ, leading to the Little-Parks oscillation, as will be seen later.

The chiral structure of the nanotube is reflected in the functional form of ξ(p). To see that, imagine a nano-ribbon obtained by cutting and flattening the nanotube. When the local continuous rotational symmetry (\({{{{{{{{\mathcal{C}}}}}}}}}_{\infty }\)) of this ribbon is reduced a discrete \({{{{{{{{\mathcal{C}}}}}}}}}_{n}\), a chiral nanotube can be obtained if the rolling direction mismatch all the high-symmetry directions. For simplicity, we consider here a system with \({{{{{{{{\mathcal{C}}}}}}}}}_{2}\) and the kinetic term may be written as (up to the 4-th order in the momentum)

$$\xi ({{{{{{{{\bf{p}}}}}}}}}_{0})= \frac{|{{{{{{{{\bf{p}}}}}}}}}_{0}{|}^{2}}{2{m}_{0}}+\frac{|{{{{{{{{\bf{p}}}}}}}}}_{0}{|}^{4}}{4{m}_{0}^{2}{\zeta }_{0}}+\frac{{p}_{x0}^{2}-{p}_{y0}^{2}}{2{m}_{1}}+\frac{{\left({p}_{x0}^{2}-{p}_{y0}^{2}\right)}^{2}}{4{m}_{1}^{2}{\zeta }_{1}}\\ +\frac{{p}_{x0}^{2}\left({p}_{x0}^{2}-3{p}_{y0}^{2}\right)+{p}_{y0}^{2}\left({p}_{y0}^{2}-3{p}_{x0}^{2}\right)}{4{m}_{2}^{2}{\zeta }_{2}}$$
(3)

where p0 is defined in a coordinate system whose axes align with the high-symmetry directions. It is generally different from that of p defined in the previous coordinate system whose x-axis is along the nanotube. They are connected by a rotation of the chiral angle θ, as shown in Fig. 1. The first two terms in Eq. (3) preserves \({{{{{{{{\mathcal{C}}}}}}}}}_{\infty }\) while the third term reduces it to \({{{{{{{{\mathcal{C}}}}}}}}}_{2}\). Note that m1 > m0 must hold for the mass along arbitrary direction to be positive. The last two terms are \({{{{{{{{\mathcal{C}}}}}}}}}_{4}\) symmetric. The inclusion of quartic terms is necessary to reveal the nonreciprocal properties, similar to the case where such an effect is caused by magnetochiral anisotropy2,4,20,21.

Equation (3) can be rewritten as

$$\xi ({{{{{{{\bf{p}}}}}}}})=\frac{{p}_{x}^{2}}{2{m}_{x}}+\frac{{p}_{y}^{2}}{2{m}_{y}}+\frac{{p}_{x}{p}_{y}}{{m}_{xy}}+\mathop{\sum }\limits_{n=0}^{4}{\kappa }_{n}{p}_{x}^{n}{p}_{y}^{4-n}$$
(4)

with, mx, my, mxy and κn being functions (see Materials and Methods) of the original parameters in Eq. (3). To see how a chiral nanotube breaks \({{{{{{{\mathcal{P}}}}}}}}\), note that py = (ny − ϕ/ϕ0)/R is defined along a circular coordinate and behaves as angular momentum (rather than the usual momentum in a flat space). It remains unchanged under \({{{{{{{\mathcal{P}}}}}}}}\) operation, consistent with the symmetry property of the magnetic flux ϕ which should not change under spatial inversion. As a result, the nanotube geometry leads to the symmetry operation \(({p}_{x},{p}_{y})\mathop{\to }\limits^{{{{{{{{\mathcal{P}}}}}}}}}(-{p}_{x},{p}_{y})\), and thus the px-odd terms in Eq. (4) break \({{{{{{{\mathcal{P}}}}}}}}\).

The supercurrent is

$${J}_{x}=-2e\int\,dy{\psi }^{*}({{{{{{{\bf{r}}}}}}}})\frac{d\xi }{d{\hat{p}}_{x}}\psi ({{{{{{{\bf{r}}}}}}}})$$
(5)
$$=-2e\mathop{\sum}\limits_{n}\frac{2\pi R}{L}\int\,dq\frac{\partial \xi ({{{{{{{\bf{p}}}}}}}})}{\partial {p}_{x}}|{\psi }_{n}(q){|}^{2},$$
(6)

where L →  is the length of the nanotube. With Eqs. (2), (4) and (6), we study the SDE when T < Tc and the NPC when T > Tc in the following.

Supercurrent diode effect

When a supercurrent passes through the nanotube, the Cooper pairs acquire a momentum p and a kinetic energy ξ(p). The order parameter is determined by the Ginzburg-Landau equation as \(|{\psi }_{n}(q){|}^{2}=|\alpha|{\beta }^{-1}{(2\pi R)}^{-2}\left(1-\xi ({{{{{{{\bf{p}}}}}}}})/|\alpha|\right)\) and the supercurrent is

$${J}_{x}(n,q)=\frac{-2eR}{{L}^{2}}\frac{|\alpha|}{\beta {R}^{2}}\left(1-\frac{\xi ({{{{{{{\bf{p}}}}}}}})}{|\alpha|}\right)\frac{\partial \xi ({{{{{{{\bf{p}}}}}}}})}{\partial {p}_{x}}.$$
(7)

Note that α < 0 since T < Tc. The critical currents Jc± are the absolute values of the maximum and minimum, respectively, of Jx(n, q) as n and q are varied.

For general parameters, Jc± can be determined numerically and the resulting diode efficiency, \(\eta \equiv \frac{{I}_{c+}-{I}_{c-}}{{I}_{c+}+{I}_{c-}}\), is shown in Fig. 2 as functions of the magnetic flux ϕ, the angle θ and the temperature, respectively. Figure 2a shows a periodicity in ϕ, similar to the Little-Parks oscillation. Different curves are for various values of the ratio r = R/l0, with R being the radius of the nanotube and \({l}_{0}=\hslash /\sqrt{2{m}_{0}{T}_{c}}\). When r is small and ϕ/ϕ0 is close to a half-integer, the transverse momentum, py = (n − ϕ/ϕ0)/R ≈ /(2rl0), costs so high a kinetic energy ξ(p) that it kills the superconductivity (i.e., ψn → 0), leading to vanishing Jc±. We define η in this case to be zero, resulting in the curve with r = 1 in Fig. 2 (a). As r increases, Jc± becomes nonzero for arbitrary magnetic flux and discontinuities occur as ϕ/ϕ0 changes across half-integers, which originates from the quantization of the transverse index n in Eq. (7). When r 1, discontinuities disappear while non-smooth kinks remain and η decreases. From Fig. 2b, one finds that η vanishes whenever θ becomes a multiple of π/2. This is expected because the nanotubes in these cases are not chiral and the inversion symmetry is preserved, forbidding the SDE. As θ/π deviate from half-integers, η increases sharply and extreme values of η are reached quickly. Note that the positions of the extreme points depend on the ratio m0/m1, which measures the strength (and the sign) of inversion symmetry breaking. The temperature dependence has the usual feature \(\eta \sim \sqrt{{T}_{c}-T}\), as shown in Fig. 2 (c).

Fig. 2: The diode efficiency, η = (Jc+ − Jc)/(Jc+ + Jc), obtained by numerically solving for the critical currents Jc± with Eq. (7).
figure 2

a The dependence on the magnetic flux (ϕ0 = h/2e is the flux quantum). The solid curves are for various values of the nanotube radius R, normalized so that r = R/l0, where \({l}_{0}=\hslash /\sqrt{2{m}_{0}{T}_{c}}\). The dashed curve is the approximate result given by Eq. (8) with r = 30. b Dependence on the angle θ which corresponds to the chiral structure of the nanotube. c The temperature dependence. The parameters are m0 = 1, m1 = 2, ζ2m2 → ∞, ζ0/Tc = 10, ζ1/Tc = 20, r = 2, θ = 0.6π, ϕ/ϕ0 = 0.3 and T/Tc = 0.9 for all the results unless specified otherwise.

It is helpful to obtain the analytical form of η, which is possible when ζ0,1,2Tc and thus the terms with κn in Eq. (4) can be treated as perturbations. We also assume r to be small, and then varying the transverse quantum number n costs so much energy that Jc± are obtained with a fixed n in Eq. (7). Under these conditions, the diode efficiency is

$$\eta = \frac{-4}{\sqrt{3}}\left(4{\kappa }_{0}\frac{{m}_{x}^{2}}{{m}_{xy}}+{\kappa }_{1}{m}_{x}\right){m}_{0}{T}_{c}\\ \times b\sqrt{\frac{|\alpha|}{{T}_{c}}\frac{{m}_{x}}{{m}_{0}}-{b}^{2}\left(\frac{{m}_{x}}{{m}_{y}}-\frac{{m}_{x}^{2}}{{m}_{xy}^{2}}\right)},$$
(8)

where b = ϕ/ϕ0 − [ϕ/ϕ0] ([x] denotes the integer closest to x). From Eq. (8) it becomes clear that either \({m}_{xy}^{-1}\) or κ1 must be nonzero to achieve the SDE. The requirement, combined with Eqs. (13) and (15), becomes \({m}_{1}^{-1}\ne 0\) and \(\sin 2\theta \ne 0\), which is just equal to requiring the nanotube to have a chiral structure. When the magnetic filed Hx is small, η is linear in Hx (note that ϕ = πR2Hx). As the magnetic flux increases, the expression under the square root becomes negative for small α since \((\frac{{m}_{x}}{{m}_{y}}-\frac{{m}_{x}^{2}}{{m}_{xy}^{2}})\) is positive definite. This results in a decrease of the transition temperature to \({T}_{c}^{{\prime} }\) with \(\delta {T}_{c}={T}_{c}-{T}_{c}^{{\prime} } \sim {b}^{2}(\frac{{m}_{x}}{{m}_{y}}-\frac{{m}_{x}^{2}}{{m}_{xy}^{2}})\frac{{m}_{0}}{{m}_{x}}\). And the temperature dependence of Eq. (8) may be written as \(\eta \sim \sqrt{{T}_{c}^{{\prime} }-T}\). A substitution of Eqs. (1216) leads to the dashed curves in Fig. 2 (a) and (c), which show great agreement with previous numerical results except two situations, (i) r 1 and ϕ/ϕ0 is close to a half integer and (ii) The temperature is far below Tc. In both situations, the assumption that Jc± can be obtained with the same index n in Eq. (7) no longer holds.

The differences in the SDE between chiral nanotube SCs and previously studied spin-orbit coupled SCs19,20,21 is clear now. The diode efficiency here is controlled by the nanotube diameter and the chiral angle, while it is determined by the SOC strength in spin-orbit coupled SCs. The sign change of η happens in both kinds of systems as the magnetic field is tuned. However, the origins are rather different. In SOC SCs, η changes sign due to the higher-order (in momentum and in field strength) terms in the kinetic energy of the Cooper pairs. Here, it is because the transverse index n corresponding to the critical currents Jc± is shifted. The sign of η changes exactly at b = 1/2 here (i.e. when the number of flux quanta is a half-integer) while the sign-flipping field-strength in SOC SCs depends on multiple system parameters.

Nonreciprocal paraconductivity

The nonreciprocity of superconducting materials manifests itself not only in the SDE when T − Tc < 0, but also in the NPC when TcT − Tc > 0. In the latter case, although the average order parameter vanishes, its quantum fluctuations induce a significant contribution to the conductance, resulting in a drop of resistance above Tc before a finite order parameter is established. The relation between the two phenomena has not been discussed elsewhere although the symmetry requirements are very similar. In this section, we calculate the paraconductivity of the chiral nanotubes described by Eq. (2) and discuss it in the same framework as we discuss the SDE.

We calculate the paraconductivity using the time-dependent Ginzburg-Landau theory43 (see Materials and Methods). The resulting current density jx = σ1E + σ2E2 + O(E3) where the linear conductivity

$${\sigma }_{1}=\gamma \frac{T}{{T}_{c}}\frac{{e}^{2}}{4{\pi }^{2}\hslash }\frac{{l}_{0}}{R}\mathop{\sum}\limits_{n}\int\,dx\frac{{\partial }_{x}^{2}{f}_{n}(x)}{{[\alpha /{T}_{c}+{f}_{n}(x)]}^{2}},$$
(9)

and the nonreciprocal term

$${\sigma }_{2}={\gamma }^{2}\frac{T}{{T}_{c}^{2}}\frac{{e}^{3}}{6{\pi }^{2}\hslash }\frac{{l}_{0}^{2}}{R}\mathop{\sum}\limits_{n}\int\,dx\frac{{\partial }_{x}^{3}{f}_{n}(x)}{{[\alpha /{T}_{c}+{f}_{n}(x)]}^{3}}.$$
(10)

In the dimensionless function fn(x) = ξ(p)/Tc, we made a change of variables, p = [px, py] → [xq0, ynq0], where yn = (n − ϕ/ϕ0)/Rq0 and q0 = 1/l0. The substitution of Eq. (4) leads to

$${f}_{n}(x)= \frac{1}{{T}_{c}}\xi \left(x\hslash {q}_{0},{y}_{n}\hslash {q}_{0}\right)\\= \frac{{x}^{2}}{2{\tilde{m}}_{x}}+\frac{{y}_{n}^{2}}{2{\tilde{m}}_{y}}+\frac{x{y}_{n}}{{\tilde{m}}_{xy}}+\mathop{\sum }\limits_{i=0}^{4}{\tilde{\kappa }}_{n}{x}^{i}{y}_{n}^{4-i}$$
(11)

where \({\tilde{m}}_{x/y/xy}={m}_{0}^{-1}{m}_{x/y/xy}\) and \({\tilde{\kappa }}_{i}={\kappa }_{i}{m}_{0}{(\hslash {q}_{0})}^{2}\) are dimensionless parameters.

The integrals in Eqs. (9) and (10) can be done numerically and the resulting σ1/2 are shown in Fig. 3 as functions of the magnetic flux ϕ and the chiral angle θ. Little-Parks oscillations of both the linear and nonlinear conductivities are found in Fig. 3a. The maxima/minima of σ1 are at integer/half-integer values of ϕ/ϕ0 since σ1 is an even function of ϕ and finite flux suppresses superconductivity. On the hand, the nonreciprocal σ2 is odd in ϕ and it vanishes whenever ϕ/ϕ0 becomes a integer. The flux values for optimal σ2 depend on the system parameters such as the nanotube radius, as shown in Fig. 3b. The curves resemble those in Fig. 2a with the difference that they are smooth here because all the transverse components n ( − , ) of the order parameter contribute, unlike the supercurrent which is given by a certain n. Fig. 3c shows the effect of the chiral angle θ. The angle dependence of σ2 is of similar amplitude to the flux dependence in Fig. 3a. In Fig. 3d, we find that the temperature dependence of σ1 is rather linear, which is similar to higher-dimensional systems2,4,43. A difference here is a shifted transition temperature \({T}_{c}^{{\prime} }\), so that \({\sigma }_{1}^{-1} \sim (T-{T}_{c}^{{\prime} })\). The T-dependence of \({\sigma }_{2}^{-1}\) is clearly of higher order and we do not find any single power law.

Fig. 3: The linear and nonlinear paraconductivity of a chiral nanotube, σ1 and σ2, normalized by \({\overline{\sigma }}_{1}=\frac{{k}_{B}T}{{T}_{c}}\frac{{e}^{2}}{4{\pi }^{2}\hslash }\frac{\gamma {l}_{0}}{R}\) and \({\overline{\sigma }}_{2}=\frac{{k}_{B}T}{{T}_{c}^{2}}\frac{{e}^{3}}{6{\pi }^{2}\hslash }\frac{{\gamma }^{2}{l}_{0}^{2}}{R}\), respectively.
figure 3

a Magnetic-flux dependence, showing the Little-Parks oscillation. b The evolution of the flux dependence as the normalized radius r is varied. c Dependence of σ1/2 on the chiral angle θ. d The temperature dependence of the inverse of σ1/2. The parameters are the same as those in Fig. 2.

Discussion

We have shown that superconducting chiral nanotubes with trapped magnetic flux behave as supercurrent diodes, whose diode efficiency strongly depends on the chiral angle. We also found, in the same theoretical framework, that the paraconductivity of such chiral nanotubes near Tc contains a nonreciprocal part σ2, whose dependence on the system parameters is rather similar to that of the SDE efficiency η and oscillates periodically as the magnetic flux ϕ or the chiral angle θ is varied. The results show that a combination of inversion symmetry breaking by chiral structure and time-reversal symmetry breaking by magnetic flux can induce nonreciprocal transport properties, including the SDE and the NPC, in superconductors.

One may notice that actual nanotubes created in laboratories are mostly related to honeycomb or triangular lattices, while the nanotubes discussed here are obtained by rolling a sheet of rectangular lattice. This choice is for technical convenience. However, the main conclusions drew here shall generally apply. To quantitatively discuss a carbon nanotube (honeycomb) or a transition-metal-dichalcogenide nanotube as experimentally studied in ref. 3 (triangular), terms up to the 6-th order in momentum must be included when constructing their Ginzburg-Landau free energies, which is not really meaningful considering the condition for the validity of the Ginzburg-Landau theory itself. Thus, a study of realistic (carbon/NbSe2/WS2/...) nanotubes may need to use the microscopic BCS theory, which can be done numerically. Another difference of the theory from real materials is that realistic nanotubes may be multi-wall and have nonzero thickness, which we ignored here. Our theory is still valid as long as chiral structures are formed and the thickness is much smaller than the superconductivity coherence length. The former condition can be satisfied by sample choice without much difficulty, and the latter one is usually satisfied since the coherence length is quite large in comparison to atomic scales.

Although single superconductors are considered here, the nonreciprocal effects discussed here shall apply to Josephson junctions where two conventional bulk superconductors (Al, Pb, Nb, NbSe2, etc.) are connected by a chiral nanotube. A study of such a system will be of great practical importance. In this manuscript, we aim to clarify the physical principles and general features of the nonreciprocal properties of superconducting chiral nanotubes, and leave more detailed and realistic studies to future works.

Although one needs to break both \({{{{{{{\mathcal{P}}}}}}}}\) and \({{{{{{{\mathcal{T}}}}}}}}\) to obtain unequal Jc±28,44, it should be noted that there also exist nonreciprocal properties in \({{{{{{{\mathcal{T}}}}}}}}\)-preserving Josephson junctions. The nonreciprocity may be observed in unequal retrapping currents Jr±45 or in ac Josephson effects9,28. The interaction between electrons plays an important role in these cases. The design or improvement of supercurrent diodes with strong electron interactions is a topic worth further investigation.

Methods

Parameters in the rotated coordinate system

By rotating the coordinate system by the chiral angle θ, one obtains the free energy form in Eq. (4) where the parameters are functions of those in Eq. (3). The functional forms are

$$\frac{1}{{m}_{x/y}}=\frac{1}{{m}_{0}}\pm \frac{\cos 2\theta }{{m}_{1}},$$
(12)
$$\frac{1}{{m}_{xy}}=-\frac{\sin 2\theta }{{m}_{1}},$$
(13)
$${\kappa }_{0}={\kappa }_{4}=\frac{1}{4}\left(\frac{1}{{\zeta }_{0}{m}_{0}^{2}}+\frac{{\cos }^{2}2\theta }{{\zeta }_{1}{m}_{1}^{2}}+\frac{\cos 4\theta }{{\zeta }_{2}{m}_{2}^{2}}\right),$$
(14)
$${\kappa }_{1}=-{\kappa }_{3}=\frac{\sin 4\theta }{2}\left(\frac{1}{{\zeta }_{1}{m}_{1}^{2}}+\frac{2}{{\zeta }_{2}{m}_{2}^{2}}\right),$$
(15)
$${\kappa }_{2}=\frac{1}{4}\left(\frac{2}{{\zeta }_{0}{m}_{0}^{2}}+\frac{1-3\cos 4\theta }{{\zeta }_{1}{m}_{1}^{2}}-\frac{6\cos 4\theta }{{\zeta }_{2}{m}_{2}^{2}}\right).$$
(16)

Time-dependent Ginzburg-Landau theory

At a temperature slightly above Tc, the fluctuation of the order parameter is determined by the following Langevin equation43,

$$\hslash \gamma {\partial }_{t}\psi ({{{{{{{\bf{r}}}}}}}},t)=-\left[\alpha+\xi (\hat{{{{{{{{\bf{p}}}}}}}}})\right]\psi ({{{{{{{\bf{r}}}}}}}},t)+\delta ({{{{{{{\bf{r}}}}}}}},t),$$
(17)

where δ(r, t) is an uncorrelated random force and γ is the inverse of damping constant. Note that α > 0 and the static order parameter vanishes, i.e. \({\langle {\psi }_{n,q}(t)\rangle }_{t}=0\). However, Eq. (17) leads to a nonzero \({\langle|{\psi }_{n,q}(t){|}^{2}\rangle }_{t}\), which is43

$$\langle|{\psi }_{n,q}(t){|}^{2}\rangle=\frac{2{k}_{B}T}{\hslash \gamma }\int\nolimits_{-\infty }^{t}d{t}^{{\prime} }{e}^{-\frac{2}{\hslash \gamma }\int\nolimits_{{t}^{{\prime} }}^{t}d{t}^{{\prime\prime} }[\alpha+\xi ({t}^{{\prime\prime} })]}.$$
(18)

It is nonzero when an electric field \({{{{{{{\bf{E}}}}}}}}=E\hat{{{{{{{{\bf{x}}}}}}}}}\) is applied, making ξ(p(t)) = ξn(q + 2eEt). Combining Eqs. (6) and (18), one obtains Eq. (9) and Eq. (10).