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

,

ss-wave superconductivity in the noncentrosymmetric W3Al2C superconductor: An NMR study

D Tay1, T Shang2∗, Y P Qi3, T P Ying4,5, H Hosono4, H-R Ott1,6, and T Shiroka1,6∗ 1Laboratorium für Festkörperphysik, ETH Zürich, CH-8093 Zurich, Switzerland 2Key Laboratory of Polar Materials and Devices (MOE), School of Physics and Electronic Science, East China Normal University, Shanghai 200241, China 3School of Physical Science and Technology, ShanghaiTech University, Shanghai 201210, China 4Materials Research Center for Element Strategy, Tokyo Institute of Technology, Yokohama 226-8503, Japan 5Beijing National Laboratory for Condensed Matter Physics, Institute of Physics, Chinese Academy of Sciences, Beijing 100190, China 6Paul Scherrer Institut, CH-5232 Villigen PSI, Switzerland [email protected] [email protected]
Abstract

We report on a microscopic study of the noncentrosymmetric superconductor W3Al2C (with Tc=7.6T_{c}=7.6 K), mostly by means of 27Al- and 13C nuclear magnetic resonance (NMR). Since in this material the density of states at the Fermi level is dominated by the tungsten’s 5dd orbitals, we expect a sizeable spin-orbit coupling (SOC) effect. The normal-state electronic properties of W3Al2C resemble those of a standard metal, but with a Korringa product 1/(T1T)1/(T_{1}T) significantly smaller than that of metallic Al, reflecting the marginal role played by ss-electrons. In the superconducting state, we observe a reduction of the Knight shift and an exponential decrease of the NMR relaxation rate 1/T11/T_{1}, typical of ss-wave superconductivity. This is further supported by the observation of a small but distinct coherence peak just below TcT_{c} in the 13C NMR relaxation-rate, in agreement with the fully-gapped superconducting state inferred from the electronic specific-heat data well below TcT_{c}. The above features are compared to those of members of the same family, in particular, Mo3Al2C, often claimed to exhibit unconventional superconductivity. We discuss why, despite the enhanced SOC, W3Al2C does not show spin-triplet features in its superconducting state and consider the broader consequences of our results for noncentrosymmetric superconductors in general.

: J. Phys.: Condens. Matter

Keywords: Noncentrosymmetric superconductors, electronic correlations, ss-wave superconductivity, NMR

\ioptwocol

1 Introduction

Noncentrosymmetric superconductors (NCSCs) belong to a class of materials that miss a key symmetry, such as parity [1]. In NCSCs, the lack of inversion symmetry of the crystal lattice often induces an antisymmetric spin-orbit coupling (ASOC), which lifts the degeneracy of the conduction-electron bands and splits the Fermi surface. Consequently, both intra- and inter-band Cooper pairs can be formed and an admixture of spin-singlet and spin-triplet pairings is possible. The extent of the singlet-triplet admixture is notably determined not just by the strength of the ASOC, but also by other microscopic parameters [1, 2]. In superconductors with spin-triplet pairing, time-reversal symmetry (TRS) breaking is not a strict requirement [3]. Hence, they are particularly interesting for studying TRS, should it occur [1]. A notable example is that of UPt3, where claims of triplet superconductivity (SC) have been confirmed by various experimental methods [4], including measurements of zero-field muon-spin relaxation [5] and optical Kerr effect [6].

A particular NCSC family, which exhibits unconventional superconducting properties, is that of M3X2YM_{3}X_{2}Y compounds. For the definition of “unconventional” superconductor (basically one that does not conform to ss-wave pairing), we follow the definition put forward in ref. [7]. Here, MM = Mo, Pd, W, Pt; XX = Li, Al; and YY = B, C, N. Five members of this family have already been studied and, in ascending order of atomic number ZZ of the metal atom MM, they are: Rh2Mo3N, Mo3Al2C, Li2Pt3B, W3Al2C, and Li2Pd3B. From general principles, it is expected that as ZZ increases, the ASOC strength increases as well. Consequently, a superconductor of an unconventional type is more likely to appear in metals with a higher atomic number. Indeed, from the existing literature, it is known that, while Li2Pd3B is a conventional superconductor [8], its high-ZZ Pt counterpart, Li2Pt3B, exhibits unconventional superconductivity [9], here identified by the presence of gap nodes [10]. The superconducting properties of the isostructural Mo3Al2C vs W3Al2C compounds are, however, still under study, with various groups reporting clearly contradictory types of superconductivity, for either of them, both conventional [11, 12, 13, 14, 15, 16, 17] and unconventional [18, 19]. Thus, in Mo3Al2C, a power-law behavior of the 27Al NMR spin-lattice relaxation rates possibly suggests superconducting gap nodes [18], while the exponential temperature dependence of the magnetic penetration depth and the absence of time-reversal symmetry breaking are more consistent with conventional nodeless SC [16, 20]. More experimental data is required to fully establish the behavior of these systems.

In this paper, we explore the electronic properties of W3Al2C, in both the normal- and the superconducting states. Our results indicate that this system is only weakly correlated. Most importantly, based on our NMR experiments, we provide evidence of BCS-type ss-wave pairing in the noncentrosymmetric W3Al2C superconductor. It is surprising that unconventional superconductivity is not observed, even though W has a higher atomic number than Mo (and, hence, a larger ASOC). We argue why, despite the presence of the heavier element, W3Al2C exhibits only conventional SC behaviour.

Refer to caption
Refer to caption
Figure 1: (a) 1/T11/T_{1} vs temperature measured in different fields. Above TcT_{c}, the field independent spin-lattice relaxation rate is linear in temperature and changes slope at T103T^{\star}\sim 103 K. (b) 1/T1T1/T_{1}T vs temperature. For T>TcT>T_{c} (here 5.3 K), the product is almost constant (0.017 s-1K-1). For comparison, we also show the data for Mo3Al2[14], Li2Pd3[21] and Li2Pt3[21] — see text for the discussion.

2 Experimental details

Polycrystalline samples of W3Al2C were prepared via high-temperature, high-pressure solid-state reaction. To achieve a homogeneous mixture, high-purity W, Al, and C powders in a quasi-stoichiometric ratio 3:1.8:0.8 were ball-milled for two days in a glove box under Ar atmosphere. The mixture was then pressed into a pellet and placed in an hh-BN capsule, where the sample was heated up to 2173 K under 5 GPa for 24 hours. Room-temperature x-ray powder diffraction (XRD) measurements were performed in a Bruker D8 diffractometer with Cu Kα\alpha radiation. Rietveld refinements via the FullProf suite confirm the cubic structure of W3Al2C, with a space group P4132P4_{1}32 (No. 213) [22]. Magnetic susceptibility and heat-capacity measurements were performed on, respectively, a vibrating sample magnetometer (VSM) and a physical property measurement system (PPMS), both by Quantum Design. Further details on sample preparation and characterization have been reported in [11].

The 27Al NMR measurements, including lineshapes and spin-lattice relaxation times, were performed on W3Al2C in powder form in different applied magnetic fields (4–7 T). To cover the 1.8 to 300 K temperature range we used a continuous-flow CF-1200 cryostat by Oxford Instruments, with temperatures below 4.2 K being achieved under pumped 4He conditions. Preliminary resonance detuning experiments confirmed a TcT_{c} of 7.6 K at 0 T (5.3 K at 5 T). The 27Al NMR signal was detected by means of a standard spin-echo sequence, consisting of π/2\pi/2 and π\pi pulses of 3 and 6 μ\mus, with recycling delays ranging from 1 to 60 s, in the 1.8–300 K temperature range. Despite an echo delay of 100 μ\mus, 2 to 32 scans were sufficient to acquire a good-quality signal. The lineshapes were obtained via fast Fourier transform (FFT) of the echo signal. Spin-lattice relaxation times T1T_{1} were measured via the inversion recovery method, using a π\piπ/2\pi/2π\pi pulse sequence. Subsequently, the T1T_{1} values for the central transition of the spin-5/2\nicefrac{{5}}{{2}} 27Al nucleus were obtained by using the relevant formula for the relaxation of quadrupole nuclei [23].

Refer to caption
Refer to caption
Figure 2: Evidence of conventional BCS superconductivity in W3Al2C. (a) Below TcT_{c}, the relaxation rate (T11T_{1}^{-1}) can be described by a single-gap exponential function exp(Δ/T)(-\Delta/T). (b) The NMR shift KsK_{s} (scaled by its normal-state value KnK_{n} at TcT_{c}) remains constant down to TcT_{c}, but decreases sharply below it. Inset: 27Al NMR line shapes at selected temperatures measured in a 5-T applied magnetic field. The vertical line indicates the reference frequency ν0\nu_{0}. Above TcT_{c}, the lineshapes are practically temperature-independent.

3 Results and discussion

3.1 27Al NMR in the normal- and superconducting phases

Refer to caption
Figure 3: Temperature dependence of the stretching coefficient β\beta. Inset: Temperature dependence of the linewidth. The behaviour of both parameters suggests the presence of intrinsic disorder in W3Al2C.
Refer to caption
Refer to caption
Figure 4: 13C NMR relaxation data in W3Al2C. Temperature dependence of the relaxation rate T11T_{1}^{-1} (a) (the HS peak is circled in black) and of the Korringa product (T1T)1(T_{1}T)^{-1} (b). Fits to equation (1) are also shown for the c=0c=0 and c=1c=1 cases (see text for details).

NMR is a powerful local technique for investigating the electronic properties of materials, in particular, their electronic correlations, magnetic order, and superconductivity [24]. Although the magnetic field reduces the TcT_{c} value, a high upper critical field (above 50 T [12]), still allows us to explore the superconducting behaviour of W3Al2C. In addition, we also investigated the NMR response in the normal state. The 27Al NMR measurements were performed in different external magnetic fields, from 4 to 7 T, corresponding to TcT_{c} values between 5.9 K and 4.6 K. In all cases, the 27Al NMR reference frequency ν0\nu_{0} was calculated with respect to a standard Al(NO3)3 reference. The T1T_{1} relaxation times were calculated by using a stretched-exponential relaxation model [23, 25], whereby disorder is modelled by a stretching coefficient β\beta:

Mz(t)Mz0=1f(135e(t/T1)β+845e(6t/T1)β+5063e(6t/T1)β)\frac{M_{z}(t)}{M_{z}^{0}}=1-f\left(\frac{1}{35}e^{(-t/T_{1})^{\beta}}+\frac{8}{45}e^{(-6t/T_{1})^{\beta}}+\frac{50}{63}e^{(-6t/T_{1})^{\beta}}\right)

Here Mz0M_{z}^{0} is the magnetization value at thermal equilibrium, while ff reflects the efficiency of population inversion (ideally 2). In general, the experimental NMR data — shifts and relaxation rates — clearly reflect the two phases of W3Al2C: a normal metallic phase above TcT_{c} and a BCS-type superconducting phase below it.

3.1.1 Metallic behaviour in the normal phase

In the normal phase, we observe an almost ideal metallic behaviour for all the applied magnetic fields (see figures 1(a) and 2(b)). For instance, the relaxation-rate data in figure 1(a) are well described by a power-law function with a scaling exponent of 1.0(1), i.e., the relaxation rate is perfectly proportional to temperature. We can also compute the Korringa product 1/T1T1/T_{1}T (here, only 0.017 s-1K-1), which is proportional to the electronic density of states at the Fermi level, N(EF)N(E_{\mathrm{F}}) (see figure 1(b)). As expected for an ideal metal, in the normal phase of W3Al2C, N(EF)N(E_{\mathrm{F}}) is practically constant with temperature and independent of the applied magnetic field, but approximately thirty times smaller than the Korringa product in metallic Al (0.54 s-1K-1[26, 27]). These widely differing relaxation rates are partly reflected also in the different Knight shifts: 450 ppm in W3Al2C vs 1636 ppm in metallic Al [28], resulting in a T1TK2T_{1}TK^{2} of 1.19×1051.19\times 10^{-5} sK and 4.98×1064.98\times 10^{-6} sK in W3Al2C and in metallic Al, respectively. The different T1TK2T_{1}TK^{2} values suggest that, although W3Al2C is an ordinary metal in its normal phase, it could still exhibit weak electronic correlations. The strong reduction in shift in W3Al2C suggests a very small contribution of ss-electrons (responsible for the contact hyperfine interaction) at the Fermi energy. This is consistent with previous NMR studies of Al-based alloys [29] and with the electronic structure calculations of W3Al2C [11]. Indeed, the latter indicate that the ss-electron band is shifted to higher energies, whereas the valence dd-electron band of the W transition metal is prominent at the Fermi level.

3.1.2 NMR evidence of BCS superconductivity in W3Al2C

In the superconducting phase, the electronic properties of the system can be described by the standard BCS theory, which predicts a gap ratio 2Δ0/kBTc=3.532\Delta_{0}/k_{\mathrm{B}}T_{c}=3.53 (see figure 2(a)) and a reduction of the NMR shift at low temperatures (see figure 2(b)). Here, the reduced shift reflects an electron pairing forming spin-compensated (typically ss-wave) Cooper pairs. This result is consistent with a recent muon-spin rotation (μ\muSR) study of W3Al2[12], where ss-wave superconductivity with the same gap ratio was reported. The superconducting gap value we find in W3Al2C is similar to the average SC gap reported in Mo3Al2C, where an NMR study claimed a 2Δ0/kBTc2\Delta_{0}/k_{\mathrm{B}}T_{c} ratio of 2.8(2)2.8(2) [14], while specific-heat measurements implied a ratio of 4.03 [13].

3.1.3 Detection of the Hebel-Slichter peak via 13C NMR

According to the Korringa relation [30], T1T_{1} in metals depends strongly on N(EF)N(E_{\mathrm{F}}). Hence, upon entering the superconducting state, we expect an exponential decay of T1T_{1}. Yet, preceding this decay, just below TcT_{c} the spin-lattice relaxation rate first increases above its normal-state Korringa value, showing a so-called coherence (or Hebel-Slichter — HS) peak [31, 32]. Such increase in relaxation rate results partly from the enhanced density of electronic states, which pile up near the SC gap edges [33, 34]. Considering its key role in the experimental confirmation of the BCS theory, the observation of a Hebel-Slichter peak is recognized as a clear indication of BCS-type ss-wave superconductivity.

However, its absence does not necessarily rule out the possibility of a standard ss-wave pairing. This seems to be also the case for W3Al2C where, as illustrated in figure 1(a), we do not observe the expected peak in the 27Al NMR relaxation rates below TcT_{c}. Several hypotheses can be put forward to explain this experimental result.

One possibility, relevant to type-II superconductors, was suggested by Goldberg and Weger [35]. Here, the total nuclear spin-relaxation rate consists of the sum of two terms, the first of which describes the relaxation in the normal-state vortex cores, while the other captures the relaxation in the remaining superconducting volume. This theory predicts that the contribution from the normal-state cores is proportional to the external magnetic field HH. To test it, we measured the relaxation rates at different magnetic fields, in the range from 4 to 7 T. As can be seen in figure 2(a), the (scaled) relaxation times generally fall on the same exponential curve, indicating that no field-dependent term arises from vortex cores. Consequently, this hypothesis cannot explain the absence of an HS peak.

Another possibility is that the HS peak is suppressed by disorder, caused by intrinsic defects and/or magnetic impurities. To investigate the role of disorder, we studied the linewidths and the stretching coefficient β\beta. As shown in figure 3, β\beta decreases continuously as the temperature is lowered, while remaining always smaller than 1. This indicates a distribution of relaxation rates, reflecting the inequivalence of NMR sites, in turn due to the intrinsic disorder (see, for instance, refs. [36, 37]). However, we find that both parameters depend only weakly on temperature and show no discontinuities or an unusual behaviour near TcT_{c}. Thus, we can rule out the possibility that disorder, although present, is affecting the HS peak.

Yet another possibility is that, for quadrupolar nuclei such as the spin-5/2\nicefrac{{5}}{{2}} 27Al, also the quadrupole interaction contributes to the spin-lattice relaxation, tending to smear out the HS peak [38]. To test this possibility, we measured the 13C NMR relaxation rate in W3Al2C. Since 13C is a spin-1/2\nicefrac{{1}}{{2}} nucleus, it represents a purely dipolar probe, not undergoing any quadrupole interaction. As can be seen in figure 4(a), we indeed observe a small feature just below TcT_{c}. Here, the exact TcT_{c} value at 5 T was determined by means of a standard resonance-detuning experiment (data not shown).

To model the feature near TcT_{c}, we begin with the usual expression for the HS peak [33]:

RsRn=2kBΔ[Ns2(E)+Ms2(E)]f(E)[1f(E)]dE.\frac{R_{s}}{R_{n}}=\frac{2}{k_{\mathrm{B}}}\int_{\Delta}^{\infty}\left[N_{s}^{2}(E)+M_{s}^{2}(E)\right]f(E)[1-f(E)]\,\mathrm{d}E. (1)

Here, RsR_{s} and RnR_{n} are the relaxation rates in the superconducting and the normal state, respectively, Ms(E)=N0Δ/E2Δ2M_{\mathrm{s}}(E)=N_{0}\Delta/\sqrt{E^{2}-\Delta^{2}} is the anomalous density of states (DOS) due to the coherence factor [33], and Ns(E)=N0E/E2Δ2N_{\mathrm{s}}(E)=N_{0}E/\sqrt{E^{2}-\Delta^{2}} is the DOS in the superconducting state.

In our case, two modifications were made. Firstly, we chose to convolute Ms(E)M_{s}(E) and Ns(E)N_{s}(E) with a triangular broadening function B(E)B(E), characterized by a width 2δ2\delta and a height 1/δ1/\delta, with δ=0.2Δ\delta=0.2\Delta. Secondly, we model the degree of coherence by the parameter cc, by making the substitution ΔcΔ\Delta\rightarrow c\Delta. The conventional BCS theory [39] predicts that, depending on the details of the scattering operator, the scattering matrix element can adopt a positive or a negative sign. The modelling of the HS peak typically requires a positive sign, but in other situations, e.g., in ultrasound absorption, the negative sign has to be used. Here, by setting cc as an adjustable parameter, one can model the case where the scattering matrix consists of both positive and negative components, which may cause a (partly) suppression of the HS peak. After these two modifications, Ns(E)N_{s}(E) and Ms(E)M_{s}(E) can be written as [40]:

Ns(E)\displaystyle N_{s}(E) =B(EE)E(E2Δ2)1/2dE,\displaystyle=\int B\left(E^{\prime}-E\right)\frac{E^{\prime}}{\left(E^{\prime 2}-\Delta^{2}\right)^{1/2}}\,\mathrm{d}E^{\prime}, (2)
Ms(E)\displaystyle M_{s}(E) =B(EE)cΔ(E2Δ2)1/2dE.\displaystyle=\int B\left(E^{\prime}-E\right)\frac{c\Delta}{\left(E^{\prime 2}-\Delta^{2}\right)^{1/2}}\,\mathrm{d}E^{\prime}. (3)

The fit results are shown in figure 4(b). We find that the fit obtained by fixing c=1c=1 drastically overestimates the relaxation rates below TcT_{c}. On the other hand, the fit obtained by fixing c=0c=0 reproduces adequately the feature near TcT_{c}, although with a slightly overestimated relaxation below TcT_{c}.

From the 13C NMR relaxation rates in the superconducting state and the above analysis we infer with some confidence that W3Al2C indeed exhibits ss-wave superconductivity, as evidenced by the observation of an (albeit reduced) HS peak. Yet, it appears that there are some decoherence effects, both at and below TcT_{c}, most likely due to a strong electron-phonon coupling in this material (see ref. [41] and the section below). In general, the exact mechanism of such decoherence effects is yet to be understood.

Refer to caption
Figure 5: Normalized electronic specific heat Ce/γTC_{\mathrm{e}}/\gamma T for W3Al2C versus temperature. Inset: The measured specific heat C/TC/T as a function of T2T^{2}. The dashed line in the inset is a fit to C/T=γ+βT2+δT4C/T=\gamma+\beta T^{2}+\delta T^{4}, yielding γ\gamma = 10(1) mJ/mol-K2, β\beta = 0.49(4) mJ/mol-K4, and δ\delta = 1.0(3) ×\times 10-3 mJ/mol-K6. The solid lines in the main panel are the electronic specific heat calculated by considering ss-, pp- and dd-wave models. Specifc heat data were take from ref. [11].

3.2 Electronic specific heat

To get further insight into the superconducting state of W3Al2C, the temperature dependence of its electronic specific heat was evaluated and analyzed using different models. After subtracting the phonon contribution from the measured specific heat (see inset in figure 5), the electronic specific heat divided by the electronic specific-heat term, i.e., Ce/γTC_{\mathrm{e}}/\gamma T, is obtained and presented in figure 5. The temperature-dependent superconducting-phase contribution to the entropy was calculated by means of the expression [42]:

S(T)=6γnπ2kB0[flnf+(1f)ln(1f)]dϵ,S(T)=-\frac{6\gamma_{\mathrm{n}}}{\pi^{2}k_{\mathrm{B}}}\int^{\infty}_{0}[f\mathrm{ln}f+(1-f)\mathrm{ln}(1-f)]\,\mathrm{d}\epsilon, (4)
Parameter Rh2Mo3N Mo3Al2C Li2Pd3B W3Al2C Li2Pt3B
Z(M)Z(M) 42 42 46 74 78
TcT_{c} (K) 4.3 [43] 9.0 [18] 7.2 [44] 7.6 [11] 2.6 [44]
γ\gamma (mJ/mol K2) 24.15 [43] 17.8 [18] 9.5 [44] 10 9.6 [44]
ΔCp/γTc\Delta C_{p}/\gamma T_{c} 1.5 2.28 [18] 1.75 [44] 2.3 0.75 [44]
ξ\xi (nm) 6.77 [43] 4.6 [18] 7.6 [44] <5.7 [11] 11.8 [44]
N(EF)N(E_{\mathrm{F}}) (states/eV/f.u.) 5.48 [18] 2.24 [45] 2.37 [11] 2.9 [45]
1/(T1T)1/(T_{1}T)(sK)-1 0.34(27Al)[14] 0.075(27Al)[18] 0.012(27Al)[19] 0.045(11B)[21] 0.017(27Al) 0.034(11B)[21]
2Δ/kBTc2\Delta/k_{\mathrm{B}}T_{c} (CpC_{p}) 3.62 [43] 4.03 [13] 4.50 non ss-wave
2Δ/kBTc2\Delta/k_{\mathrm{B}}T_{c} (NMR) 2.8 [14] 2.2 [8] 3.5 non ss-wave
2Δ/kBTc2\Delta/k_{\mathrm{B}}T_{c} (μ\muSR) 3.46 (dirty) [46] 2.59 [20] 3.5 [12] non ss-wave
3.84 (clean) [46]
Table 1: Comparison of the superconducting parameters of selected noncentrosymmetric superconductors with the chemical formula M3X2YM_{3}X_{2}Y, where MM is a metal. In all the cited references, N(EF)N(E_{\mathrm{F}}) includes the contribution from SOC.

where γn\gamma_{\mathrm{n}} is the normal-state electronic specific-heat coefficient, f=(1+eE/kBT)1f=(1+e^{E/k_{\mathrm{B}}T})^{-1} is the Fermi function and E(ϵ)=ϵ2+Δ2(T)E(\epsilon)=\sqrt{\epsilon^{2}+\Delta^{2}(T)} is the excitation energy of quasiparticles, with ϵ\epsilon the electron energies measured relative to the chemical potential (Fermi energy) [42, 47]. Here, Δ(T)\Delta(T) is the temperature dependent gap function, which in the BCS ss-wave model can be written as Δ(T)=Δ0tanh{1.82[1.018(Tc/T1)]0.51}\Delta(T)=\Delta_{0}\mathrm{tanh}\{1.82[1.018(T_{\mathrm{c}}/T-1)]^{0.51}\} [48], with Δ0\Delta_{0} the gap value at zero temperature. In case of pp- and dd-wave models, the temperature-dependent gap functions are Δ(T)\Delta(T)sinθ\theta, and Δ(T)\Delta(T)cos2ϕ\phi, respectively, exhibiting point- and line nodes in the respective gap function. The temperature-dependent electronic specific heat in the superconducting state can be calculated from Ce=TdSdTC_{\mathrm{e}}=T\frac{dS}{dT}.

The fit results using the above-mentioned models are shown by solid lines in the main panel of figure 5. We find that, while the ss-wave model fits the electronic specific heat data across the entire temperature range, the pp- and dd-wave models deviate significantly from the data, implying the absence of any gap nodes in the superconducting state of W3Al2C. The fully-gapped ss-wave model gives a superconducting gap Δ0=2.25(5)\Delta_{0}=2.25(5)kBk_{\mathrm{B}}TcT_{c}, much larger than the expected weak-coupling BCS value (1.74 kBk_{\mathrm{B}}TcT_{c}), thus suggesting strongly-coupled [49] Cooper pairs in W3Al2C. This is confirmed also by the specific-heat discontinuity at TcT_{c}, i.e., Δ\DeltaCC/γ\gammaTcT_{c}, here around 2.5–2.7, which is higher than 1.43 predicted by the BCS theory.

3.3 Comparison with other members of the M3X2Y family

The superconducting properties of the M3X2YM_{3}X_{2}Y family of NCSCs are summarized in table 1. The data are somewhat challenging to interpret, because of the lack of a clear trend when arranged in order of increasing atomic number Z(M)Z(M). Most likely this reflects the fact that the superconducting behaviour of the M3X2YM_{3}X_{2}Y compounds is highly dependent on the details of their band structure, which differs widely between members of the same family. Despite this difficulty, we can still draw some conclusions.

First, we note that the SC gap values 2Δ/kBTc2\Delta/k_{\mathrm{B}}T_{c} seem to differ, depending on whether they were determined by means of NMR, μ\muSR, or CpC_{p} measurements. In general, the gaps estimated via NMR and μ\muSR, which are both local-probe techniques, tend to be in good agreement. Conversely, the SC gaps determined from CpC_{p} data tend to be higher than those calculated from NMR and μ\muSR data, and systematically indicate a strong electron-phonon coupling across the entire family of M3X2YM_{3}X_{2}Y compounds.

Refer to caption
Figure 6: 1/T1T1/T_{1}T vs temperature for Mo3Al2C as obtained from Koyama et al [14], Bauer et al [18] and Kuo et al [19]

Secondly, we note that the (T1T)1(T_{1}T)^{-1} values of W3Al2C are consistent with those of the analog compound Mo3Al2C obtained by Bauer et al [18]. Since Mo3Al2C adopts the same structure as W3Al2C, the only differences being that W is less electronegative than Mo (1.7 vs 2.16 on the Pauling scale) and the marginally larger atomic covalent radius of the W atoms compared to Mo (150 vs 145 pm), we expect the N(EF)N(E_{\mathrm{F}}) of Mo3Al2C and W3Al2C to be of a similar magnitude and, consequently, their (T1T)1(T_{1}T)^{-1} values to be comparable. Considering that the N(EF)N(E_{\mathrm{F}}) of Mo3Al2C is twice as large as that of W3Al2C (5.48 vs 2.37, see table 1), this results in an approximately fourfold increase in relaxation rate [(T1T)1=0.075(T_{1}T)^{-1}=0.075 vs 0.017 (sK)-1, see table 1]. On the other hand, the (T1T)1(T_{1}T)^{-1} values obtained by Kuo et al [19] and Koyama et al [14] are wildly different (see figure 6). Recently, it was shown that the deliberate introduction of vacancies (here Al) in M3X2YM_{3}X_{2}Y by soft chemical etching does not significantly influence the density of states (as from DFT calculations and heat-capacity data), but instead it dramatically alters the strength of electron-phonon coupling [41]. Hence, the presence of vacancies might justify the different (T1T)1(T_{1}T)^{-1} values obtained by the different groups. Furthermore, since W3Al2C intrisically contains more Al-vacancies than its Mo counterpart [41], this might explain the BCS-type character of the former.

Thirdly, our results in table 1 support the observation made by Bauer et al [1] that, in case of phonon-mediated superconductors, although a noncentrosymmetric crystalline structure is required for the occurrence of unconventional superconductivity, it does not necessarily imply it. Another remarkable example of this is Mo3[50] which, as W3Al2C reported here, is a normal BCS-type superconductor, despite being an NCSC. Further, even within the same family, isostructural compounds can display a different superconducting behavior. For instance, CePt3Si and LaPt3Si are isostructural (P4mmP4mm space group) and both display a large SOC. However, while CePt3Si exhibits unconventional superconductivity [51, 52], LaPt3Si does not [52]. Future studies could shed more light on this issue and clarify the link between unconventional SC and the lack of a structural inversion center.

4 Conclusion

Extensive NMR- and specific-heat measurements in the noncentrosymmetric W3Al2C superconductor, provide ample evidence about its weak electron correlation, yet with a strong electron-phonon coupling. Most importantly, we establish that W3Al2C shows a conventional BCS-type ss-wave superconductivity. This, apparently unnoteworthy result, is surprising in view of the enhanced SOC of W atoms, whose 5dd orbitals dominate the density of states at the Fermi level. This is all the more remarkable, if one considers that the analogous Mo3Al2C compound, hosting the much lighter Mo atoms, is claimed to be an unconventional superconductor. Such a counterintuitive outcome may be explained with the subtle role played by spin-orbit coupling, as well as by its competition with the electron bandwidth in the M3X2YM_{3}X_{2}Y family. Finally, by considering the similar case of Mo3P, we show that W3Al2C represents yet another example of a noncentrosymmetric material with conventional normal-state- and superconducting properties, thus emphasizing the role of a noncentrosymmetric structure as a beneficial (but not sufficient) condition in achieving unconventional superconductivity.

5 Acknowledgements

T.S. acknowledges support from the Natural Science Foundation of Shanghai (Grants No. 21ZR1420500 and 21JC1402300). This work was partially supported by the Schweizerische Nationalfonds zur Förderung der Wissenschaftlichen Forschung, SNF (Grant no. 200021-169455). Y.P. Qi was supported by the National Natural Science Foundation of China (Grants No. U1932217 and 11974246) and the Science and Technology Commission of Shanghai Municipality (19JC1413900).

6 References

References

  • [1] Bauer E and Sigrist M (eds) 2012 Non-Centrosymmetric Superconductors vol 847 (Berlin: Springer Verlag)
  • [2] Smidman M, Salamon M B, Yuan H Q and Agterberg D F 2017 Rep. Prog. Phys. 80 036501
  • [3] Anderson P W 1984 Phys. Rev. B 30(7) 4000–4002
  • [4] Joynt R and Taillefer L 2002 Rev. Mod. Phys. 74 235 and references therein
  • [5] Luke G M, Keren A, Le L P, Wu W D, Uemura Y J, Bonn D A, Taillefer L and Garrett J D 1993 Phys. Rev. Lett. 71 1466
  • [6] Schemm E R, Gannon W J, Wishne C M, Halperin W P and Kapitulnik A 2014 Science 345 190
  • [7] Norman M R 2011 Science 332 196–200
  • [8] Nishiyama M, Inada Y and Zheng G q 2005 Phys. Rev. B 71(22) 220505
  • [9] Shimamura A, Furukawa Y, Kumagai K, Takeya H and Hirata K 2007 Physica C 460–462 663–665
  • [10] Yuan H Q, Agterberg D F, Hayashi N, Badica P, Vandervelde D, Togano K, Sigrist M and Salamon M B 2006 Phys. Rev. Lett. 97(1) 017006
  • [11] Ying T P, Qi Y P and Hosono H 2019 Phys. Rev. B 100(9) 094522
  • [12] Gupta R, Ying T P, Qi Y P, Hosono H and Khasanov R 2021 Phys. Rev. B 103(17) 174511
  • [13] Karki A B, Xiong Y M, Vekhter I, Browne D, Adams P W, Young D P, Thomas K R, Chan J Y, Kim H and Prozorov R 2010 Phys. Rev. B 82(6) 064512
  • [14] Koyama T, Maeda Y, Yamazaki T, Ueda K, Mito T, Kohara T, Waki T, Tabata Y, Tsunemi H, Ito M and Nakamura H 2013 J. Phys. Soc. Jpn. 82 073709
  • [15] Koyama T, Ozaki Y, Ueda K, Mito T, Kohara T, Waki T, Tabata Y, Michioka C, Yoshimura K, Suzuki M T and Nakamura H 2011 Phys. Rev. B 84(21) 212501
  • [16] Bonalde I, Kim H, Prozorov R, Rojas C, Rogl P and Bauer E 2011 Phys. Rev. B 84(13) 134506
  • [17] Zhigadlo N D, Logvinovich D, Stepanov V A, Gonnelli R S and Daghero D 2018 Phys. Rev. B 97(21) 214518
  • [18] Bauer E, Rogl G, Chen X Q, Khan R T, Michor H, Hilscher G, Royanian E, Kumagai K, Li D Z, Li Y Y, Podloucky R and Rogl P 2010 Phys. Rev. B 82(6) 064511
  • [19] Kuo C N, Liu H F and Lue C S 2012 Phys. Rev. B 85(5) 052501
  • [20] Bauer E, Sekine C, Sai U, Rogl P, Biswas P K and Amato A 2014 Phys. Rev. B 90(5) 054522
  • [21] Nishiyama M, Inada Y and Zheng G q 2007 Phys. Rev. Lett. 98(4) 047002
  • [22] Rodríguez-Carvajal J 1993 Physica B: Condens. Matter 192 55
  • [23] McDowell A F 1995 J. Magn. Reson. Ser. A 113 242–246
  • [24] Alloul H 2015 Scholarpedia 10 30632
  • [25] Narayanan A, Hartman J S and Bain A D 1995 J. Magn. Reson. Ser. A 112 58–65
  • [26] Carter G C, Bennett L H and Kahan D J 1977 Metallic shifts in NMR: A review of the theory and comprehensive critical data compilation of metallic materials (Progr. Mater. Sci. vol 20) (Oxford: Pergamon Press)
  • [27] Bennett L H, Watson R E and Carter G C 1970 J. Res. Nat. Bur. Stand. 74A 569–610
  • [28] Mehring M and Raber H 1973 Solid State Commun. 13 1637–1639
  • [29] Sandor M T, Kecskes L J, He Q, Xu J and Wu Y 2011 Chin. Sci. Bull. 56(36) 3937–3941 and references therein
  • [30] Korringa J 1950 Physica 16(7) 601–610
  • [31] Hebel L C and Slichter C P 1957 Phys. Rev. 107(3) 901–902
  • [32] Hebel L C and Slichter C P 1959 Phys. Rev. 113(6) 1504–1519
  • [33] Hebel L C 1959 Phys. Rev. 116(1) 79–81
  • [34] Alloul H 2014 Scholarpedia 9 32069
  • [35] Goldberg I B, Ehrenfreund E and Weger M 1968 Phys. Rev. Lett. 20(11) 539–540
  • [36] Shiroka T, Casola F, Glazkov V, Zheludev A, Prša K, Ott H R and Mesot J 2011 Phys. Rev. Lett. 106(13) 137202
  • [37] Shiroka T, Eggenschwiler F, Ott H R and Mesot J 2019 Phys. Rev. B 99(3) 035116
  • [38] Li Z, Jiao W H, Cao G H and Zheng G q 2016 Phys. Rev. B 94(17) 174511
  • [39] Bardeen J, Cooper L N and Schrieffer J R 1957 Phys. Rev. 108(5) 1175–1204
  • [40] Ding Q P, Wiecki P, Anand V K, Sangeetha N S, Lee Y, Johnston D C and Furukawa Y 2016 Phys. Rev. B 93(14) 140502
  • [41] Ying T, Muraba Y, Iimura S, Yu T, Cheng P, Kamiya T, Lu Y, Li J, Qi Y and Hosono H 2020 iScience 23 101196
  • [42] Tinkham M 1996 Introduction to Superconductivity 2nd ed (Mineola, NY: Dover Publications)
  • [43] Wei W, Zhao G J, Kim D R, Jin C, Zhang J L, Ling L, Zhang L, Du H, Chen T Y, Zang J, Tian M, Chien C L and Zhang Y 2016 Phys. Rev. B 94(10) 104503
  • [44] Takeya H, Kasahara S, El Massalami M, Mochiku T, Hirata K and Togano K 2007 Mater. Sci. Forum 561–565 2079–2082
  • [45] Lee K W and Pickett W E 2005 Phys. Rev. B 72(17) 174505
  • [46] Shang T, Smidman M, Ghosh S K, Baines C, Chang L J, Gawryluk D J, Barker J A T, Singh R P, Paul D M, Balakrishnan G, Pomjakushina E, Shi M, Medarde M, Hillier A D, Yuan H Q, Quintanilla J, Mesot J and Shiroka T 2018 Phys. Rev. Lett. 121 257002
  • [47] Padamsee H, Neighbor J E and Shiffman C A 1973 J. Low Temp. Phys. 12 387
  • [48] Carrington A and Manzano F 2003 Physica C 385 205
  • [49] Rainer D and Sauls J A 1995 Strong-coupling theory of superconductivity (Singapore: World Scientific) chap 2, pp 45–78
  • [50] Shang T, Philippe J, Verezhak J A T, Guguchia Z, Zhao J Z, Chang L J, Lee M K, Gawryluk D J, Pomjakushina E, Shi M, Medarde M, Ott H R and Shiroka T 2019 Phys. Rev. B 99(18) 184513
  • [51] Bauer E, Hilscher G, Michor H, Paul C, Scheidt E W, Gribanov A, Seropegin Y, Noël H, Sigrist M and Rogl P 2004 Phys. Rev. Lett. 92(2) 027003
  • [52] Ribeiro R L, Bonalde I, Haga Y, Settai R and Ōnuki Y 2009 J. Phys. Soc. Jpn. 78 115002