Superconductivity with peculiar upper critical fields in quasi-one-dimensional Cr-based pnictides
Cao Guang-Han1, †, Zhu Zeng-Wei2
Department of Physics, Zhejiang University, Hangzhou 310027, China
Wuhan National High Magnetic Field Center, School of Physics, Huazhong University of Science and Technology, Wuhan 430074, China

 

† Corresponding author. E-mail: ghcao@zju.edu.cn

Abstract
Abstract

The discovery of superconductivity in quasi-one-dimensional Cr-based pnictides A2Cr3As3 (A = alkali metals) has generated considerable research interest, primarily owing to their reduced dimensionality, significant electron correlations, and possible unconventional superconductivity. The upper critical field () provides important information on the superconducting pairing. In this paper, we first briefly overview the latest research progress on the Cr-based superconductors. Then, we introduce typical behaviors of type-II superconductors in relation with the pair-breaking mechanisms. After a description of the measurement method for , we focus on the analysis of data, especially for the temperature and angle dependence, in K2Cr3As3 crystals. The result indicates (i) an absence of Pauli-paramagnetic pair breaking for field perpendicular to the Cr3As3 chains, and (ii) a unique threefold modulation for the in-plane profile. Finally we conclude with remarks on the possible unconventional superconducting pairing symmetry.

1. Introduction

Superconductivity is an intriguing macroscopic quantum state in solids due to the condensation of electronic Cooper pairs. In a phonon-mediated BCS-type conventional superconductor, the superconducting order parameter is mostly of an s-wave in which the underlying Cooper pairs are in a spin-singlet state. In a so-called unconventional superconductor, however, the electron pairing interaction is basically non-phonon-mediated and, the superconducting order parameter is mostly of a non-s-wave, which breaks additional symmetries apart from the global U(1) gauge symmetry.[1] If the Cooper pairs are of spin triplet, which is analogous to those in 3He superfluid phases, the superconductor is always considered as being unconventional. Nonetheless, such a spin-triplet superconductor appears to be rare. Up to 2014, the main candidates of spin-triplet superconductors include some organic superconductors,[2,3] UPt3,[4,5] Sr2RuP4,[6] and U-based superconducting ferromagnets.[7] Note that the superconducting transition temperature Tc of these superconductors is relatively low (below 1.5 K), which is undesirable for in-depth investigations.

Following the discovery of superconductivity at ∼ 2 K under an external pressure of ∼ 1 GPa in CrAs,[8,9] quasi-one-dimensional (Q1D) Cr-based arsenides A2Cr3As3 (A = K, Rb, and Cs) were discovered in 2015.[1014] The superconducting transition temperature Tc is 6.1, 4.8, and 2.2 K, respectively, for A = K, Rb, and Cs. The normal-state ρ(T) behavior is very different for polycrystalline and single-crystalline samples, primarily because of the anisotropy (see below). The polycrystalline samples show somewhat linear temperature dependence of resistivity [Fig. 1(a)]. In contrast, for high-quality crystals of K2Cr3As3, the normal-state ρ(T) below 50 K follows a power-law relation with an exponent close to 3 [Fig. 1(b)]. The Tc value of high-quality crystals is also increased a little, compared with that of the polycrystalline samples. The defect/impurity effect on Tc can be seen elsewhere,[15] which shows substantial suppression of superconductivity due to nonmagnetic scattering, suggestive of a non-s-wave pairing in K2Cr3As3.

Figure 1. (color online) (a) Resistive superconducting transitions in polycrystals of A2Cr3As3 (A =K, Rb, Cs). The inset shows the close-up. (b) Superconducting transition in electrical resistivity (with current flowing along the c-axis direction) for K2Cr3As3 single crystals. The inset shows the derivative of resistivity, defining and .

Note that the Tc value increases monotonically with the decrease of ionic radii of the alkali-metal elements, which implies that Tc could be further increased if K+ ions are replaced with smaller cations such as Na+ and Li+. Whereas high-temperature synthesis of these target compounds were unsuccessful,[14] very recently, it was reported that the expected sister compound Na2Cr3As3 could be obtained through an ion-exchange process.[16] The Tc value was indeed increased (up to 8.6 K). This finding further corroborates that Tc increases with decreasing the a axis (interchain distance), as shown in Fig. 2. Interestingly, the Sommerfeld coefficient, derived from the specific-heat measurements, follows a similar dependence.[14]

Figure 2. (color online) Superconducting transition temperature Tc (left) and the electronic specific-heat coefficient (right) as functions of the lattice parameter a for A2Cr3As3 (A = Na, K, Rb, Cs).

If the increase of Tc above is considered as a consequence of chemical pressure, then the effect of physical pressure is surprisingly in an opposite tendency. Earlier preliminary data show that Tc of K2Cr3As3 decreases at a rate of −0.34 K/GPa under hydrostatic pressures up to 0.7 GPa.[17] Later high-pressure measurements together with crystal-structure investigations were done for both K2Cr3As3 and Rb2Cr3As3.[18] The result indicates that Tc decreases almost linearly to below 2 K and, by extrapolation, superconductivity disappears at about 8 GPa and 6 GPa, respectively, for K2Cr3As3 and Rb2Cr3As3. A similar result was independently reported for K2Cr3As3.[19] In-situ structural determinations reveal that the extent of non-centrosymmetry, defined by the difference between bond angles of As2–Cr2–As2 (α) and As1–Cr1–As1 (β), seems to be in relation with superconductivity. Note that for the centrosymmetric non-superconducting 133 series ACr3As3 (A = K, Rb, Cs),[20,21] the (αβ) value becomes zero since there is only one Cr site.

Dramatically, Ren and co-workers[22] recently reported that the 133-type KCr3As3 could superconduct at 5 K provided it was in single crystalline form. Additionally, the same group observed superconductivity at 7.3 K in RbCr3As3 single crystals.[23] The Tc value of the latter case is significantly higher than that of Rb2Cr3As3 (4.8 K), which basically rules out the possibility that superconductivity may come from Rb2Cr3As3 that resides in the sample. So far, however, it is not clear whether CsCr3As3 and NaCr3As3 crystals superconduct or not.

Before ending this section, we introduce efforts on explorations of superconductivity in quasi-two-dimensional (Q2D) Cr-based systems. Actually, it was the attempt to look for superconductivity in “Ba1−xKxCr2As2” that led to the discovery of the Q1D K2Cr3As3 superconductor.[14] Along this line, we also discovered Sr2Cr3As2O2[24] and Ba2Ti2Cr2As4O,[25] unfortunately, neither of them are superconducting above 1.8 K. Some theoretical studies predict that doped Q2D Cr-based pnictides could host unconventional superconductivity.[26,27] So far, however, we have not yet observed any trace of superconductivity in the Q2D systems.

2. Basic properties of A2Cr3As3

As shown in Fig. 3(a), the A2Cr3As3 family is structurally characterized by the double-walled linear sub-nanotubes separated by the K+ counterions. Consistent with the Q1D crystal structure, the resistivity data [Fig. 3(b)] indeed show a strong anisotropy with an anisotropic ratio, , up to 120 at 50 K. The shows a quasi-linear behavior below ∼30 K, which may account for the large resistivity value and linearity in polycrystalline samples [Fig. 1(a)]. At around 50 K, interestingly, exhibits a resistivity maximum, characteristic of 3D-to-1D dimension crossover as observed in Q1D so-called Bechgaard slats.[28] This result suggests that a Tomonaga–Luttinger liquid (TLL) state could be realized at high temperatures.[28,29]

Figure 3. (color online) (a) A top view of the crystal structure of K2Cr3As3 with the c axis along the linear Cr3As3 chain. (b) Anisotropic resistivity of K2Cr3As3 single crystals, from which a 3D-to-1D dimension crossover can be seen in the case of (i.e., electric current flowing perpendicular to the chain direction). (c) Calculated band structure of nonmagnetic K2Cr3As3. The inset shows comparison between relativistic (in dark violet) and non-relativistic (in grey) results (from Ref. [30]). (d) Fermi-surface sheets (non-relativistic result).

The inner Cr3 sub-nanotubes can be theoretically modelled in terms of a twisted Hubbard tube, which gives rise to 1D molecular-orbital bands.[31] Electronic band-structure calculations reveal that the electronic states near the Fermi energy are dominated by the Cr-3d atomic orbitals.[30] Note that the asymmetrical spin–orbit interactions are sizable around K and Γ points, as shown in the inset of Fig. 3(c). There are two Q1D Fermi-surface sheets (FSs) in the α and β bands and one three-dimensional (3D) FS pockets in the γ band [Fig. 3(d)].[30,32,33] The two hole-like Q1D FSs are experimentally observed in a recent angle-resolved photoemission spectroscopy study,[34] from which a TLL behavior is signatured. The TLL physics has been earlier suggested both experimentally[35] and theoretically.[31,36,37]

For the prototype compound K2Cr3As3, the total density of states (DOS) at EF is calculated to be ,[30] corresponding to a ‘bare’ electronic specific-heat coefficient of . The experimental Sommerfeld coefficient () is 3.5 times larger,[10] indicating remarkable electron-mass renormalization (electron–phonon coupling constant is generally less than 1.0 owing to the lattice instability). Similar comparison between calculations and experiments for Rb2Cr3As3 and Cs2Cr3As3[32,33] suggests that the electron-mass renormalization factor decreases significantly with the interchain distance. Furthermore, there is a positive correlation between Tc and the Sommerfeld coefficient (Fig. 2). What controls the electron correlations? How do the electron correlations determine Tc as well as the superconducting pairing? What is the role of one dimensionality on superconductivity? These questions remains open at present.

Remarkably, the Cr-based Q1D arsenides show evidence of unconventional superconductivity with a possible spin-triplet pairing. Theoretical investigations based on an effective theoretical model[3841] show that the most attractive pairing interactions are always in spin-triplet channels. Spin-triplet superconductivity is also suggested by the theoretical analyses in terms of the multiband TLL scenario[31,36] and melting of macroscopic spin polarization in a weakly coupled odd-gon-unit system.[42] Nevertheless, spin-singlet pairing with a -like gap function is obtained from a phenomenological two-band model.[37] Note that conventional electron-phonon mediated superconductivity is also proposed for K2Cr3As3.[43] The calculated total electron-phonon coupling constant is as large as , which seems too high to maintain the lattice.

Experimentally, the 75As Knight shift data[44] indicate the existence of ferromagnetic spin fluctuations, in favor of spin-triplet pairing, consistent with the theoretical calculations and analyses.[30,32] Additionally, the zero-field SR measurements show evidence of the spontaneous appearance of an internal magnetic field below Tc.[45] The NMR investigations show no Hebel–Slichter coherence peak,[35,44,46] against conventional superconductivity. Furthermore, the Cr-spin fluctuations are captured from the nuclear spin-lattice relaxation rate 1/, implying non-phonon-mediated superconductivity. It was found that, with increasing the interchain distance, the strength of Cr spin fluctuations is progressively suppressed, together with the decrease in Tc.[46] The inelastic neutron scattering in K2Cr3As3 also indicates the existence of short-range magnetic fluctuations, which suggests proximity to a magnetic instability.[47] The incipient magnetic order is considered to couple strongly to the lattice and competes with superconductivity as well. As for the properties of the superconducting gap function, most experiments (including NMR,[35] SR,[45,48] penetration depth,[49,50] and specific heat[51]) mostly suggest the existence of line nodes, again favoring unconventional superconductivity.

The upper critical field () of A2Cr3As3 appears to be very unusual,[14] which also gives valuable information on the superconducting pairing. Nevertheless, there were very different interpretations to the experimental data.[17,5254] To clarify the related issues clearly, in this paper, we present some basic knowledge first, then focus on the details of temperature and angular dependence of , and finally try to discuss the result on an equal footing.

3. versus pair breaking

To understand the electron pairing in a superconductor, an alternative method is to investigate how the pairs are broken. The upper critical magnetic field of a type-II superconductor may serve as an indicator for unconventional superconductivity.[3] In the following, we first address how Cooper pairs are broken in a type-II superconductor if an external magnetic field is applied. Then, we present the general experimental method for the measurements of . We will focus on the data as functions of temperature and field direction in the A2Cr3As3 superconductors. The implication on the possible superconducting pairing symmetry is discussed. Note that this article has some overlap with our recent publications.[14,53,54]

As is known, superconductors can be classified into type-I and type-II superconductors, according to their responses to external magnetic fields. Unlike type-I superconductors, which show full diamagnetism (Meissner effect) until the external field exceeds the thermodynamic critical field , type-II superconductors are characterized by an additional phase, the mixed state (or vortex state), which contains quantized magnetic fluxons. The two phase boundaries (Meissner to mixed state and mixed state to normal state) define lower and upper critical fields, namely, and . According to Ginzburg–Landau (GL) theory, the borderline of type-I and type-II superconductors locates at , where κ is the so-called GL parameter, λ and ξ denote magnetic penetration depth and superconducting coherence length, respectively. In the case of orbital pair breaking, relates to ξ by the formula .

In general, a magnetic field suppresses superconductivity through the following two mechanisms. The first is the orbital pair breaking due to electromagnetic interactions, which is described by GL equations.[57] The purely orbital pair breaking limits the in a way characterized by a linear relation at , as shown in Fig. 4(a). The second is the so-called paramagnetic effect,[58,59] which is due to the difference in spin susceptibility between superconducting and normal states (the spin susceptibility of a spin-singlet superconductor is zero). If there were no orbital effects, then the spin paramagnetic effect by itself would yield (here the electron g factor is given to be 2.0), where Δ is the superconducting energy gap. Given the BCS result with , one obtains the so-called Pauli-paramagnetic (or Clogston–Chandrasekhar) limit, (in unit Tesla).[58,59] Note that paramagnetic pair breaking is featured by a concave-down curvature, as shown in Fig. 2(b), and holds near Tc. Since grows much faster than with decreasing T, therefore, an orbital pair breaking mostly dominates at .

Figure 4. (color online) Simulations of temperature dependence of upper critical field with different pair-breaking mechanisms. Panel (a) represents the scenario of purely orbital limitation. Panel (b) describes the purely paramagnetic-limited case without spin–orbit scattering. Panel (c) gives an example in which both mechanisms take effect and, the spin–orbit scattering effect is included. The results in panels (a) and (c) are based on Werthammer–Helfand–Hohenberg theory.[55] denotes the strength of spin–orbit scattering. The Maki parameter[56] α is defined by , where is the purely orbitally limited value at zero temperature, denotes the Pauli-paramagnetic limit.

The Pauli-paramagnetic limit can be exceeded either by strong coupling or by spin–orbit scattering. The latter effect can be parameterized by the Werthammer–Helfand–Hohenberg (WHH) theory,[55] which simultaneously includes both pair-breaking effects together with spin–orbit scattering [Fig. 4(c)]. It is found that spin–orbit scattering counteracts the paramagnetic effect in limiting . As a result, the can be significantly enhanced through spin–orbit scattering. Note that, in certain circumstances, a novel Fulde–Ferrel–Larkin–Ovchinnikov (FFLO) state with spatial modulations of superconducting order parameter may appear at high fields. The appearance of FFLO often accompanies an upturn in .[60,61]

In fact, the pair-breaking effect displayed in Fig. 4 only applies for simple isotropic, s-wave gap, and single-band cases, without the considerations of strong coupling, electron-mass anisotropy, nodal gap, and multiband scenarios. In particular, if the Cooper pairs are in a spin-triplet configuration, which can also be spin-polarized by the external field, then no paramagnetic effect is expected at certain field directions. In a spin-triplet pairing, the superconducting order parameter is expressed using a so-called vector,[1] thus only a certain field direction may avoid the paramagnetic pair breaking. In contrast, orbital pair breaking by electromagnetic interactions always exists (with exceptions for an extremely large anisotropy) regardless of the spin structure of the superconducting Cooper pairs. We will come back to this point later.

4. Measurement of

In general, is measured by the sharp superconducting transition in a physical quantity under external magnetic fields. The physical quantities can be resistivity, ac susceptibility, dc magnetization, and heat capacity, etc. Two experimental modes can be employed, namely, field-sweep at a fixed temperature and temperature-sweep at a fixed field, which give and , respectively. One should find a suitable technique for the measurement, compatible with the specific cases and experimental conditions.

Among the different techniques mentioned above, measurement of resistivity under magnetic fields is the most common one, since it is relatively easily operated and accessible. Figure 5 shows an example for the anisotropic measurement of Rb2Cr3As3 crystals. Note that different crystals had to be used for different field directions owing to the sample’s extreme sensitivity to moisture (tiny moisture may deteriorate the sample) and also the apparatus’ incapability to rotate the sample in situ. What we emphasize here is about the criterion of in T-sweep (or in H-sweep). can be regarded as the lowest field in which the first magnetic vortex is allowed to enter the sample’s interior. Therefore, the junction between superconducting and normal states should be the position of (or ), if superconducting thermal fluctuations are not remarkable. That is why here we employ onset superconducting temperature for the determination of .

Figure 5. (color online) Measurement of anisotropic upper critical fields exemplified by temperature dependence of resistivity for Rb2Cr3As3 crystals under magnetic fields along (a) and perpendicular to (b) the crystallographic c axis. The resulting data are plotted in panel (c), yielding the superconducting phase diagrams (from Ref. [54]). The crossing between 4.5 K and 5 K comes from different samples (with slightly different Tc values) employed.

In this work we employed a pulsed magnetic field up to 60 T in Wuhan’s National High Magnetic Field Center for the measurement of K2Cr3As3 crystals. For the high-field measurement of Rb2Cr3As3, we also employed the Cell 1 Water-Cooling Magnet of the High Magnetic Field Laboratory of the Chinese Academy of Sciences, which generates the maximum magnetic field up to 38.5 T and allows the magneto-resistance measurement down to 0.3 K. Note that it is very important to avoid the sample’s deterioration and keep the electrodes on the sample effective during the measurement.

5. Temperature dependence of
5.1. K2Cr3As3 and Na2Cr3As3

Figure 6 shows the magneto-resistivity measurement result for K2Cr3As3 polycrystals, which gives an initial slope of .[10] By using the WHH formula without spin–orbit coupling,[55] one obtains a zero-temperature upper critical field of , which is three times the Pauli-paramagnetic limit . Thus the Maki parameter,[56] , is above 4, suggesting either a dominant Pauli-limiting effect at low temperatures or a novel spin-triplet superconductivity. In the case of Na2Cr3As3, the value is estimated to be 54 T,[16] 3.4 times of its Pauli limit (16 T). The result contrasts with another Q1D superconductor Tl2Mo6Se6 with , whose values are merely 5.8 T and 0.47 T, respectively, for magnetic fields parallel and perpendicular to the c axis.[62] Nevertheless, a large beyond the Pauli limit is seen in Q1D Li0.9Mo6O17,[63] which is regarded as evidence of spin-triplet superconductivity.[64]

Figure 6. (color online) The preliminary measurement for K2Cr3As3 polycrystals, giving an initial slope of −7.43 T/K. The Pauli limit is marked with the horizontal line in the inset (adapted from Ref. [10]).

Recalling that Tc decreases monotonically with applying pressures, very interestingly, the of the pressurized sample of K2Cr3As3 decreases much faster than Tc itself.[19] The value decreases to below the Pauli limit at above 2 GPa. The estimated Maki parameter drops from 4 at ambient pressure to well below 1 at , indicating a crossover from Pauli-paramagnetic limitation to orbital limitation. If the high value at ambient pressure represents a sign of spin-triplet pairing, the pressure-induced suppression of may suggest a breakdown of the spin-triplet pairing channel upon applying pressures.

Kong, Bud’ko, and Canfield[17] first reported the anisotropic in K2Cr3As3 crystals. The initial slopes were measured to be −7 T/K and −12 T/K, respectively, for and . Meanwhile, the results obtained by Zuo et al.[53] are −4.2 T/K and −10.3 T/K, and those by Wang et al.[65] are −5 T/K and −16.1 T/K, respectively. Although these data look scattered, primarily because of the sample’s sensitivity and the difference in the temperature range of linear fitting, the result points to a relatively weak effective-mass anisotropy with near Tc. Nevertheless, the upper bound of the effective-mass anisotropy is consistent with the resistivity anisotropy value at .[14]

Owing to the expected high at low temperatures, measurements under high magnetic fields are necessary to reveal the whole behavior. Balakirev et al.[52] measured and for K2Cr3As3 at high magnetic fields up to 60 T, using a contactless technique based on a proximity detector oscillator. The results show a Pauli-limited behavior for , yet exhibits an orbitally limited behavior with no signs of paramagnetic pair breaking. Consequently, an anomalous reversal in appears at ∼4 K.

We independently conducted the anisotropic measurements for K2Cr3As3 crystals by direct magnetoresistance measurement for the first time. A standard four-probe technique was employed, and the sample was installed on a rotator that bears pulsed magnetic fields up to 60 T.[53] Figure 7(a) and 7(b) show the field dependence of the magnetoresistance with and , respectively. can be clearly identified by the junction points between superconducting and normal states, although no zero-resistance is achieved due to the sample’s extreme sensitivity to the environment. The obtained data, plotted in Fig. 7(c), basically agree with those of the previous report.[52] Indeed, shows a downward curvature, in contrast with the upward curvature for . One also sees that and cross at 3.1 K. The low-temperature values for both field directions exceed the Pauli limit by a factor of 2–3. We will discuss the origin and its implications later on.

Figure 7. (color online) Field dependence of resistance at fixed temperatures in K2Cr3As3 crystals with the field directions (a) and (b). Panels (c) and (d) plot the extracted and the anisotropic parameters as functions of temperature (adapted from Ref. [53]).
5.2. Rb2Cr3As3 and Cs2Cr3As3

The anisotropic measurement for Rb2Cr3As3 under relatively low fields is shown in Fig. 5, from which the data can be directly obtained. As is seen, the superconducting transition moves to lower temperatures with applying magnetic fields, and Tc is suppressed more rapidly for . The resultant data are plotted in Fig. 5(c). One sees that exhibits a steeper increase with decreasing temperature, but the magnitude of the slope tends to decrease. By contrast, increases slowly near Tc, yet the magnitude of the slope increases with decreasing temperature. The result in Cs2Cr3As3 shown in Fig. 8 is quite similar, although the values are close to the Pauli limit. Therefore, the contrasting and behavior is universal to all three members of A2Cr3As3. As illustrated in Fig. 4, the concave-down feature in indicates a dominant Pauli-limited pair-breaking mechanism. The almost linear characteristic meets an orbitally limiting mechanism. The slight concave-up behavior may reflect multiband superconductivity since there are three or four FSs in the system.[32,33]

Figure 8. (color online) Anisotropic versus temperature in Cs2Cr3As3 crystals. The solid line is the WHH fitting for (see text for the details).

Generally, orbital pair breaking dominates at . This means that the initial slope of is a measure of the orbitally-limited upper critical field at zero temperature, . From the data in Fig. 5(c), we obtain the initial slopes of and , respectively, for Rb2Cr3As3. The data of Cs2Cr3As3 (in Fig. 8) show and . According to GL theory, the effective-mass anisotropy, , are about 30 and 50, respectively, for Rb2Cr3As3 and Cs2Cr3As3. The result suggests stronger Q1D anisotropy in Rb2Cr3As3 and Cs2Cr3As3, consistent with the larger interchain distance in the crystal structure.[14]

Figure 9(a) and 9(b) show the field dependence of resistance at fixed temperatures with different field directions in Rb2Cr3As3. We employ the identical criterion for the determination of . While it is evident to determine the for sample C (with ) which shows a single sharp superconducting transition, however, it turns out to be ambiguous for sample D (with ) because it exhibits two main transitions (possibly arising from the electrode contact problem and/or sample’s partial deterioration.[20,21]). Figure 9(c) presents an additional data set for , from which we find the steepest transition to use for the determination. The resultant data are displayed in Fig. 9(d). One sees that different sets of data basically meet each other. The phase diagram looks similar to that of K2Cr3As3. Firstly, shows an obvious downward curvature, characteristic of Pauli-limiting behavior. tends to saturate to 17.5 T, twice of , without any sign of the formation of the FFLO state. The enhanced seems to be mostly related to the strong coupling superconductivity as evidenced by the large specific-heat jump .[10] Secondly, is approximately linear in the whole temperature range, which strongly suggests the absence of the Pauli-paramagnetic pair breaking effect. In particular, keeps increasing linearly down to 0.075Tc. This suggests that, apart from the orbitally pair-breaking effect with the typical profile shown in Fig. 4(a), there could be a positive role played by external magnetic fields for superconductivity. Spin-triplet with equal-spin pairing may account for such a novel behavior. Thirdly, the data for K2Cr3As3 and Rb2Cr3As3 almost coincide with each other, although the latter shows a larger anisotropy close to Tc. In both cases, the apparent reversal at about 0.4Tc actually reflects different pair breaking mechanisms for different field directions. We will further address this issue with the measurement of polar-angle dependence.

Figure 9. (color online) (a) and (b) Field dependence of resistance at fixed temperatures below Tc, for Rb2Cr3As3 crystals with and , respectively. Panel (c) shows an additional data set for another crystal with . Panel (d) summarizes the result in Rb2Cr3As3, using as the horizontal axis. The Pauli paramagnetic limit is marked. The inset plots the anisotropic parameter in a logarithmic scale (adapted from Ref. [54]).

The phenomenon of reversal, i.e., crossing between and , has been witnessed in several other systems including the Q1D organic superconductor (TMTSF)2PF6,[66] iron-based chalcogenides,[61,67,68] and the heavy-fermion superconductor UPt3.[69] At first sight, the reversal in K2Cr3As3 and Rb2Cr3As3 suggests a dimension crossover. However, it seems unlikely from the crystal and electronic structure points of view. Moreover, the dimension crossover scenario can be ruled out by our data (see below) which show modulations at the “crossover” temperature of 3.1 K. This is definitely inconsistent with the expected angle dependence, , in which a constant is given for the effective-mass ratio .

We note that details of the behaviors differ in different systems. In (TMTSF)2PF6, shows a pronounced upward curvature for the field directions perpendicular to at low temperatures.[66] This arises from the dramatic reduction of the orbitally pair-breaking effect owing to a field-induced dimension crossover.[3] In iron-based superconductors, the data reveal that the Pauli-limiting effect is always at work regardless of the field directions,[61,67] consistent with a spin-singlet pairing. The anisotropy inversion appears only accidentally, and generally approaches 1.0 at ,[70] which actually reflects the Pauli-paramagnetic pair breaking effect associated with anisotropic spin–orbit coupling. By contrast, the reversed anisotropy ratio in UPt3 is sizable at zero temperature,[69] similar to the case here in A2Cr3As3 (A = K and Rb). As for Cs2Cr3As3, although the Maki parameter decreases to 1.8 for , the data show a linear dependence, suggesting that there is still a possibility of an absence of the Pauli-limiting effect for .

6. Angle dependence of in K2Cr3As3
6.1. Polar-angle dependence

To further clarify if there is any “residual” Pauli-limiting effect for , we performed the angle dependence measurement. Namely, the sample rotates in situ such that the field direction tilts a “polar” angle θ from the c axis, as shown in the inset of Fig. 10. First of all, at temperatures near , decreases monotonously with increasing θ (from 0° to 90°). This is understandable since there is a uniaxial effective-mass anisotropy. As the temperature is lowered, an additional maximum in appears at θ = 90°. The enhanced is consistent with the reduced Pauli-limiting effect for .

Figure 10. (color online) Polar angle (θ) dependent at fixed temperatures in K2Cr3As3. The solid lines are the fitted result using Eq. (2) (see text for details), and the dashed pink curve (for 5-K data only) is the fitted result with Eq. (1) (adapted from Ref. [53]).

Below we show that the whole data set can be perfectly explained by the combination of a fully anisotropic Pauli-limiting effect (i.e., absence of Pauli-limiting effect for ) with a uniaxial effective-mass anisotropy. According to GL theory, the effective-mass anisotropy leads to the anisotropy of the orbitally limited ,

where denotes the presumed orbitally limited upper critical field for and, is the effective-mass ratio. Indeed, equation (1) can basically describe the experimental data at 5 K (albeit of obvious deviations at ). However, it completely fails to meet the data at lower temperatures.

The reason for the failure of Eq. (1) is because the paramagnetic pair breaking effect is not included. Given the contrasting behavior for and , the paramagnetic pair breaking is assumed to be fully anisotropic. Namely, only the magnetic-field component parallel to the c axis further suppresses the in Eq. (1). The paramagnetically pair-breaking effect can be parameterized by an effective Pauli-limiting field . Therefore, we have , in which (fully Pauli limiting) for θ = 0° and, (absence of Pauli limiting) for θ = 90°.

The effective Pauli-limiting field suppresses in a way of competition between the related energies.[55,58,59,71] Given , therefore, . Consequently, one obtains an explicit expression for ,

Remarkably, the whole experimental data set of satisfies the above equation very well, as shown in Fig. 10, which further clarifies the absence of the Pauli-limiting effect for . The data fitting yields three parameters, , ϵ, and at a given temperature. As expected, the fitted is basically linear. The linear extrapolation yields a zero-temperature orbitally limited of 51.5 T for , which is reasonably larger than the extrapolated value of for . The obtained and exceed the Pauli limit by a factor of 4.6 and 3.4, respectively. Using the GL relations, and , where is the flux quantum, the anisotropic coherence lengths at zero temperature can be estimated to be and . The value exceeds twice the interchain distance in K2 Cr3As3,[10] indicating the nature of anisotropic 3D superconductivity.

6.2. Azimuthal-angle dependence

Figure 11(a) shows derivative of the magnetoresistance for the field directions with different azimuthal ϕ angles. The data can be extracted by the peak value in dR/dH (here we use a different criterion so as to increase the measurement precision). The ϕ angle varied from 1.87° to 153.38°. According to the crystal symmetry with the point group of , the data set is rotated by 120° and −120°, respectively, in order to plot a complete polar diagram. One sees that the overlapped region matches well, which validates the rotation. Consequently, the obtained profile shows a three-fold modulation with an amplitude of 0.36 T. The maximum of appears for the field directions along the crystallographic a and b axes. Here the possibility of surface superconductivity and the influence of anisotropic normal-state magnetoresistance can be excluded (see the Supplementary Materials in Ref. [53]). Note that the modulation (with a relative magnitude up to ∼1.2%) in K2Cr3As3 is actually more obvious than those in UPt3[72] and MgB2,[73] both of which also crystallize in a hexagonal lattice.

Figure 11. (color online) In-plane azimuthal-angle dependence at 1.572 ± 0.001 K for K2Cr3As3. (a) The derivatives of for different ϕ angles, from which is obtained. (b) Polar plot of the extracted . The red (blue) triangles come from the rotation of the derived data with (−120°), respectively. Shown at the upper right is the photograph of the sample configuration in the measurement. The lower-right corner shows the possible Cr6-octahedron spin order for the magnetic ground state, according to first-principles calculations (adapted from Ref. [53]).

The sixfold modulation of in UPt3[72] can be explained in terms of a coupling to the symmetry breaking field.[74] In the case of K2Cr3As3, according to the first-principles calculations, the magnetic ground state appears to be a novel in-out co-planar magnetic order,[30,32] as shown at the lower-right corner in Fig. 11(b). The magnetic moments lie in the basal plane. Intuitively, this result seems to be compatible with the observed modulation that can be similarly explained by the coupling to the symmetry breaking field.

7. Discussion and conclusion

For a conventional superconductor with a high comparable to , the value is normally limited by the paramagnetic effect regardless of field directions. Considering the paramagnetically limited behavior of in K2Cr3As3, as well as the preliminary observation of the insensitivity of Tc on impurity scattering, Balakirev et al.[52] proposed a novel spin-singlet superconductivity. The absence of the Pauli-limiting effect for is explained by assuming electron-spin locking along the c direction. In fact, we have demonstrated that the apparent insensitivity of Tc on impurity scattering is due to the clean limit () for the superconductivity (see below). So, their staring point seems to have deviated. Secondly, according to the first-principles calculations,[30,32] the Cr moments likely align perpendicular to the chain direction that is possibly associated with the sizable spin–orbit coupling shown in Fig. 3(c), contradicting with the Ising-like spin-locking scenario. More importantly, our in-plane profile also opposes the Ising-like spin-singlet superconductivity.

We argue that spin-triplet pairing explains the overall data more naturally. The absence of the Pauli-limiting effect for and a large value (3.4 times of ) actually dictate a spin-triplet pairing scenario, as in the case of UPt3.[4] The latter was recently confirmed to host a spin-triplet odd-parity superconductivity.[75] The (pseudo)spins of the odd-parity Cooper pairs are with , which is equivalent to the spin state of . In this circumstance, the Zeeman energy breaks the Cooper pairs for , hence the Pauli-limiting behavior for . By contrast, the perpendicular field simply changes the population of Cooper pairs with spin directions and , and therefore, no paramagnetic pair-breaking is expected for . Indeed, by taking spin-triplet pairing into consideration, and with some simplifications and approximations, we are able to derive an equation for , whose solution is consistent with Eq. (2) (for details see the Supplementary Materials in Ref. [53]). Here we note that, owing to the non-centrosymmetric crystal structure in K2Cr3As3, the pairing symmetry is in principle a mixture of singlet and triplet states,[38] except for the case of a simple pz wave in which a purely triplet pairing is anticipated because of the mirror-plane reflection symmetry.[39]

Finally, we comment on the impurity scattering effect on Tc, which is important to judge the possibility of either the singlet or triplet pairing state. For an odd-parity unconventional superconductor, nonmagnetic scattering serves as a source of pair breaking even at zero field, hence Tc suppression is expected. However, such an impurity scattering effect will not be evident in the clean-limit regime, i.e., the electron mean free path l is much larger than the superconducting coherence length ξ. In K2Cr3As3, l is estimated to be ∼75 nm (after the electron-mass renormalization is considered) for the electron transport along the c axis in the sample with RRR = 61, and is only 3.5 nm, thus holds. A similar case is seen in Rb2Cr3As3.[54] This explains why the Tc keeps almost unchanged for single-crystal samples with different RRRs. Our recent study shows that, when introducing sufficient impurities (), Tc is indeed suppressed in K2Cr3As3.[15]

In conclusion, we show that the detailed measurements supply very important information on the superconducting pairing of the Q1D Cr-based superconductors. The apparent “anisotropy-reversal” phenomenon actually reflects different pair-breaking mechanisms for different magnetic-field directions. The absence of paramagnetic pair breaking for is further clarified by the data set in K2Cr3As3. The three-fold modulation in implies time-reversal symmetry breaking in the superconducting state. The spin structure of the superconducting Cooper pairs probably involves Cr-spin locking within the ab plane, and the dominant pairing is likely to be of a spin-triplet state, which seems more natural to explain the full data. So far, the spin-triplet pairing scenario is consistent with the previous suggestions in experimental and theoretical investigations. Nevertheless, clarification of the specific pairing symmetry needs more investigations.

Unconventional superconductivity has been and remains one of the major topics in the area of condensed matter physics. According to the classification by Hirsch, Maple, and Marsiglio,[76] there are 11 classes of materials that can be categorized into “unconventional superconductors” in which superconductivity does not come from conventional electron-phonon interactions. The present status of the research, in particular, with the measurement, strongly suggests that the newly discovered A2Cr3As3 superconductors should belong to the unconventional category.

Acknowledgment

We thank Pi L, Xi C Y, Bao J K, Tang Z T, Liu Y, Zuo H K, Wang J H, Kang J, Xu Z A, Yuan H Q, Cao C, Ning F L, Imai T, Zheng G Q, and Sun L L for our fruitful collaborations. Thanks are also due to Agterberg D F, Zhang F C, Zhou Y, Dai J H, Luo J L, Cheng J G, Yang F, and Hu J P for their helpful discussions.

Reference
[1] Sigrist M Ueda K 1991 Rev. Mod. Phys. 63 239
[2] Jerome D 1991 Science 252 1509
[3] Zhang W De Melo C A R S 2007 Adv. Phys. 56 545
[4] Sauls J A 1994 Adv. Phys. 43 113
[5] Joynt R Taillefer L 2002 Rev. Mod. Phys. 74 235
[6] Mackenzie A P Maeno Y 2003 Rev. Mod. Phys. 75 657
[7] Huxley A D 2015 Physica 514 368
[8] Wu W Cheng J Matsubayashi K Kong P Lin F Jin C Wang N Uwatoko Y Luo J 2014 Nat. Commun. 5 5508
[9] Cheng J G 2017 Acta Phys. Sin. 66 037401
[10] Bao J K Liu J Y Ma C W Meng Z H Tang Z Sun Y L Zhai H F Jiang H Bai H Feng C M Xu Z A Cao G H 2015 Phys. Rev. 5 011013
[11] Tang Z T Bao J K Liu Y Sun Y L Ablimit A Zhai H F Jiang H Feng C M Xu Z A Cao G H 2015 Phys. Rev. 91 020506
[12] Tang Z T Bao J K Wang Z Bai H Jiang H Liu Y Zhai H F Feng C M Xu Z A Cao G H 2015 Sci. China-Materials 58 16
[13] Hu J 2015 Sci. Bull. 60 2140
[14] Cao G H Bao J K Tang Z T Liu Y Jiang H 2017 Phil. Mag. 97 591
[15] Liu Y Bao J K Zuo H K Ablimit A Tang Z T Feng C M Zhu Z W Cao G H 2016 Sci. China-Phys. Mech. Astron. 59 657402
[16] Mu Q G Ruan B B Pan B J Liu T Yu J Zhao K Chen G F Ren Z A 2018 Phys. Rev. Mater. 2 034803
[17] Kong T Budk´o S L Canfield P C 2015 Phys. Rev. 91 020507
[18] Wang Z Yi W Wu Q Sidorov V A Bao J Tang Z Guo J Zhou Y Zhang S Li H Shi Y Wu X Zhang L Yang K Li A Cao G Hu J Sun L Zhao Z 2016 Sci. Rep. 6 37878
[19] Sun J P Jiao Y Y Yang C L Wu W Yi C J Wang B S Shi Y G Luo J L Uwatoko Y Cheng J G 2017 J. Phys. Condens. Matter 29 455603
[20] Bao J K Li L Tang Z T Liu Y Li Y K Bai H Feng C M Xu Z A Cao G H 2015 Phys. Rev. 91 180404
[21] Tang Z T Bao J K Liu Y Bai H Jiang H Zhai H F Feng C M Xu Z A Cao G H 2015 Sci. China-Materials 58 543
[22] Mu Q G Ruan B B Pan B J Liu T Yu J Zhao K Chen G F Ren Z A 2017 Phys. Rev. 96 140504
[23] Liu T Mu Q G Pan B J Yu J Ruan B B Zhao K Chen G F Ren Z A 2017 Europhys. Lett. 120 27006
[24] Jiang H Bao J K Zhai H F Tang Z T Sun Y L Liu Y Wang Z C Bai H Xu Z A Cao G H 2015 Phys. Rev. 92 205107
[25] Ablimit A Sun Y L Jiang H Bao J K Zhai H F Tang Z T Liu Y Wang Z C Feng C M Cao G H 2017 J. Alloys Compd. 694 1149
[26] Wang W S Gao M Yang Y Xiang Y Y Wang Q H 2017 Phys. Rev. 95 144507
[27] Edelmann M Sangiovanni G Capone M de Medici L 2017 Phys. Rev. 95 205118
[28] Moser J Gabay M Auban-Senzier P Jérome D Bechgaard K Fabre J 1998 Eur. Phys. J. 1 39
[29] Biermann S Georges A Lichtenstein A Giamarchi T 2001 Phys. Rev. Lett. 87 276405
[30] Jiang H Cao G Cao C 2015 Sci. Rep. 5 16054
[31] Zhong H Feng X Y Chen H Dai J 2015 Phys. Rev. Lett. 115 227001
[32] Wu X X Le C C Yuan J Fan H Hu J P 2015 Chin. Phys. Lett. 32 057401
[33] Alemany P Canadell E 2015 Inorg. Chem. 54 8029
[34] Watson M D Feng Y Nicholson C W Monney C Riley J M Iwasawa H Refson K Sacksteder V Adroja D T Zhao J Hoesch M 2017 Phys. Rev. Lett. 118 097002
[35] Zhi H Z Imai T Ning F L Bao J K Cao G H 2015 Phys. Rev. Lett. 114 147004
[36] Miao J J Zhang F C Zhou Y 2016 Phys. Rev. 94 205129
[37] Wachtel G Kim Y B 2016 Phys. Rev. 94 104522
[38] Zhou Y Cao C Zhang F C 2017 Sci. Bull. 62 208
[39] Wu X Yang F Le C Fan H Hu J 2015 Phys. Rev. 92 104511
[40] Zhang L D Wu X Fan H Yang F Hu J 2016 Europhys. Lett. 113 37003
[41] Cao G H 2017 Sci. Bull. 62 206
[42] Reja S. Nishimoto S. 2018 arXiv: 1801.06204
[43] Subedi A 2015 Phys. Rev. 92 174501
[44] Yang J Tang Z T Cao G H Zheng G Q 2015 Phys. Rev. Lett. 115 147002
[45] Adroja D T Bhattacharyya A Telling M Feng Y Smidman M Pan B Zhao J Hillier A D Pratt F L Strydom A M 2015 Phys. Rev. 92 134505
[46] Zhi H Lee D Imai T Tang Z Liu Y Cao G 2016 Phys. Rev. 93 174508
[47] Taddei K M Zheng Q Sefat A S de la Cruz C 2017 Phys. Rev. 96 180506
[48] Adroja D Bhattacharyya A Smidman M Hillier A Feng Y Pan B Zhao J Lees M R Strydom A Biswas P K 2017 J. Phys. Soc. Jpn. 86 044710
[49] Pang G M Smidman M Jiang W B Bao J K Weng Z F Wang Y F Jiao L Zhang J L Cao G H Yuan H Q 2015 Phys. Rev. 91 220502
[50] Pang G M Smidman M Jiang W B Shi Y G Bao J K Tang Z T Weng Z F Wang Y F Jiao L Zhang J L Luo J L Cao G H Yuan H Q 2016 J. Magn. Magn. Mater. 400 84
[51] Shao Y.T. Wu X.X. Wang L. Shi Y.G. Hu J.P. Luo J.L. 2018 Phys Rev B submitted
[52] Balakirev F F Kong T Jaime M McDonald R D Mielke C H Gurevich A Canfield P C Budk´o S L 2015 Phys. Rev. 91 220505
[53] Zuo H Bao J K Liu Y Wang J Jin Z Xia Z Li L Xu Z Kang J Zhu Z Cao G H 2017 Phys. Rev. 95 014502
[54] Tang Z T Liu Y Bao J K Xi C Y Pi L Cao G H 2017 J. Phys.: Condens. Matter 29 424002
[55] Werthamer N R Helfand E Hohenberg P C 1966 Phys. Rev. 147 295
[56] Maki K 1966 Phys. Rev. 148 362
[57] Helfand E Werthamer N R 1966 Phys. Rev. 147 288
[58] Clogston A M 1962 Phys. Rev. Lett. 9 266
[59] Chandrasekhar B S 1962 Appl. Phys. Lett. 1 7
[60] Gurevich A 2010 Phys. Rev. 82 184504
[61] Gurevich A 2011 Rep. Prog. Phys. 74 124501
[62] Petrović A P Lortz R Santi G Decroux M Monnard H Fischer O Boeri L Andersen O K Kortus J Salloum D Gougeon P Potel M 2010 Phys. Rev. 82 235128
[63] Mercure J F Bangura A F Xu X Wakeham N Carrington A Walmsley P Greenblatt M Hussey N E 2012 Phys. Rev. Lett. 108 187003
[64] Sepper O Lebed A G 2014 Phys. Rev. 90 094509
[65] Wang X F Roncaioli C Eckberg C Kim H Yong J Nakajima Y Saha S R Zavalij P Y Paglione J 2015 Phys. Rev. 92 020508
[66] Lee I J Naughton M J Danner G M Chaikin P M 1997 Phys. Rev. Lett. 78 3555
[67] Khim S Kim J W Choi E S Bang Y Nohara M Takagi H Kim K H 2010 Phys. Rev. 81 184511
[68] Maiorov B Mele P Baily S A Weigand M Lin S Z Balakirev F F Matsumoto K Nagayoshi H Fujita S Yoshida Y Ichino Y Kiss T Ichinose A Mukaida M Civale L 2014 Supercond. Sci. Technol. 27 044005
[69] Shivaram B S Rosenbaum T F Hinks D G 1986 Phys. Rev. Lett. 57 1259
[70] Yuan H Q Singleton J Balakirev F F Baily S A Chen G F Luo J L Wang N L 2009 Nature 457 565
[71] Hake R R 1967 Appl. Phys. Lett. 10 189
[72] Keller N Tholence J L Huxley A Flouquet J 1994 Phys. Rev. Lett. 73 2364
[73] Tamegai T Shi Z Pradhan A K Nakamura H Ghosh A K Tokunaga M Takano Y Togano K Kito H Ihara H 2003 J. Low Temp. Phys. 131 1153
[74] Agterberg D F Walker M B 1995 Phys. Rev. Lett. 74 3904
[75] Schemm E R Gannon W J Wishne C M Halperin W P Kapitulnik A 2014 Science 345 190
[76] Hirsch J E Maple M B Marsiglio F 2015 Physica 54 1