Introduction

Transition metal oxides are well known as one of the most fascinating solids because of their variety of physical properties1 such as metal-insulator transition2, giant magnetoresistance3, ferroelectricity4, and high-temperature superconductivity5. In transition metal oxides, anisotropic-shaped d-orbital electrons and strong electron correlation effect caused by Coulomb interaction between electrons play important roles for emerging the various physical phenomena6. The first discovery of superconductors in oxide compounds is a perovskite-type structural SrTiO3-x7, and the discovery has led to the development of many kinds of oxide superconductors, for example, BaPb1-xBixO38, YBa2Cu3O79, Li1+xTi2-xO410 and so on. Among the discoveries of oxide superconductors, the η-carbide-type superconductors A4B2X have attracted attention in recent studies; here, A and B are transition metals and X is a light element such as carbon, nitrogen, or oxygen11,12. Ma et al. reported the bulk superconductivity of Zr4Rh2Ox (x = 0.7 and 1)12 and Nb4Rh2C1-δ13 at transition temperatures of Tc = 2.8 K (x = 0.7), 4.8 K (x = 1) and 9.75 K, respectively. They also found that Ti4M2O with M = Co, Rh, and Ir show superconductivity at Tc = 2.7 K, 2.8 K, and 5.4 K, respectively14. The significant discovery of them was not only founding superconductors but also observing a large upper critical field μ0Hc2(0) violating the Pauli-Clogston limit (Pauli limit) for Nb4Rh2C1-δ, Ti4Co2O, and Ti4Ir2O. The value of Pauli limit μ0HP is determined with a certain magnetic field at which a gain of paramagnetic Zeeman energy at a normal state is equal to a superconducting condensation energy, given in the following formula15,16:

$$\begin{array}{c}{\mu }_{0}{{H}}_{\text{P}} \, \text{=} \, \frac{\Delta \text{(}{0}\text{)}}{\sqrt{g}{\mu }_{\text{B}}} \, \text{=} \, \text{1.86}{T}_{\text{c}}\text{,}\end{array}$$
(1)

where g = 2 is a g-factor for free electron and μB ≈ 9.27×10–24 J T–1 is a Bohr magneton. The Δ(0) is a superconducting gap energy at 0 K described as Δ(0) = 1.76kBTc (kB ≈ 1.38×10–23 J K–1 is a Boltzmann constant) in the single gap Bardeen−Cooper−Schrieffer (BCS) model17. The large μ0Hc2(0) overwhelming the μ0HP can arise from special electronic states and structural properties such as multi-band effect18,19,20, spin-triplet cooper pairing21, Fulde-Ferrell-Larkin-Ovchinnikov (FFLO) superconducting state22,23, global or local inversion symmetry breaking24,25, and strong spin-orbit coupling (SOC)26,27. Particularly, spin-orbit scattering originating from SOC suppresses a cooper pair breaking by the Pauli paramagnetic effect because SOC destroys spin as a good quantum number, and makes spin susceptibility of the superconducting state close to that of a normal state, described in Werthamer–Helfand–Hohenberg (WHH) theory28,29. Therefore, the strong SOC has the potential to achieve a large μ0Hc2(0) superconducting state, and the strength can be controlled by chemical elemental substitution30. The strength of SOC, ξ can be approximately calculated using a hydrogen-like atom model31:

$$\begin{array}{c}\xi \propto \, \frac{{Z}^{4}}{{n}^{3}{l}\left({l}\,+\,\frac{1}{{2}}\right)\left({l}\,+\,1\right)}\text{,}\end{array}$$
(2)

where Z, n, and l are atomic number, principal quantum number, and orbital angular momentum, respectively. From the expression, we can understand the strength of SOC proportions to Z4 within the same electronic orbital. In the case of the Ti4Ir2O superconductor, we can expect the Ir-5d orbital hosting enhanced SOC should play an important role for the large μ0Hc2(0), and it was found that Ti-3d and Ir-5d orbitals hybridize near its Fermi level and the violation of the Pauli limit is a result of a combination of strong-coupled superconductivity, SOC, and strong electron correlation32. Furthermore, the large SOC splitting a band structure along Γ-K lines due to the Ir-5d electrons was weakened by applying pressures, and large μ0Hc2(0) undergoes a crossover at 35.6 GPa from well beyond to less than the μ0HP33. As mentioned above, the η-carbide-type superconductors have been studied from the points of view of the large μ0Hc2(0) and SOC effect based on d-block transition metals. Table 1 shows a list of the η-carbide-type superconductors with Tc, μ0HP, and μ0Hc2(0). Some kinds of η-carbide-type oxide superconductors have been reported; however, it is not sufficient as of now, and develo** new examples of them is important for a deeper understanding of the η-carbide-type oxide superconducting properties.

Table 1 A list of η-carbide-type superconductors.

Herein, we focus on a Zr4B2O system because superconductivity was solely confirmed in Zr4Rh2O in the system to the best of our knowledge. We report the discovery of an unrevealed superconducting nature of Zr4Pd2O known as a hydrogen storage material34,35. Polycrystalline samples of Zr4Pd2O were obtained by arc melting followed by annealing in an evacuated quartz tube. We performed synchrotron X-ray diffraction (SXRD) measurement at the beamline BL13XU in SPring-8 and checked chemical composition by means of the energy dispersive X-ray spectroscopy (EDX) method to characterize obtained samples. The bulk superconductivity was confirmed through magnetic susceptibility, electrical resistivity, and specific heat measurement, resulting in Tc = 2.8 K, 2.73 K, and 2.6 K, respectively. We discuss the μ0Hc2(0) for Zr4Pd2O and Zr4Rh2O using the electrical resistivity and specific heat data measured at several magnetic fields. We find that Zr4Rh2O shows the μ0Hc2(0) = 6.16 T, lower than the μ0HP = 7.59 T; however, Zr4Pd2O shows the large μ0Hc2(0) = 6.88 T, violating the μ0HP = 5.29 T different from Zr4Rh2O. The violation of the Pauli limit for Zr4Pd2O can be attributed to the larger strength of SOC derived from Pd-4d electrons.

Results

Crystal structures of Zr4Pd2O and Zr4Rh2O

Schematic images of the crystal structure for Zr4Tr2O (Tr = Pd or Rh) are shown in Fig. 1a. These compounds crystalline a cubic η-carbide crystal structure with a space group Fd \(\stackrel{\mathrm{-}}{3}\) m (No. 227). The metal atoms Zr occupy Wyckoff positions 48f (labeled as Zr1), 16d (labeled as Zr2) and Tr occupies Wyckoff position 32e position. The O atom occupies Wyckoff position 16c, and the occupying can be regarded as void filling in a Ti2Ni-type structure. The complicated η-carbide crystal structure consists of Zr1 octahedra centered by O (Fig. 1b) and a geometrically frustrated stella quadrangula lattice (Fig. 1c)36. The Zr1 octahedra caging the O atom at the center are arranged in the unit cell sharing the corner as seen in a pyrochlore structure. The stella quadrangula lattice can be formed by inserting a small tetrahedron into each tetrahedron making the pyrochlore lattice, and the unit consists of nested Tr and Zr2 tetrahedra with the same center of gravity37,38. The SXRD patterns and Rietveld refinement results of Zr4Pd2O and Zr4Rh2O at Room temperature are shown in Fig. 1d and e, respectively. The SXRD patterns were well fit to cubic η-carbide crystal structure and the lattice constants of Zr4Pd2O and Zr4Rh2O were determined to a = 12.4617(1) Å and 12.3977(3) Å, respectively. The values of a for Zr4Pd2O and Zr4Rh2O were confirmed to be close to the previous report12,34,35. We found a small amount of impurity phases such as ZrO2 and Zr in both of them, and reliability factors Rwp were 8.438% for Zr4Pd2O and 16.550% for Zr4Rh2O. We also refined the atomic coordinates and isotropic atomic displacement parameter Uiso for each atom, as summarized in Table 2. The Uiso of the O atom was fixed to be 0.004 because the value approximately close to be zero within the errors on the fitting quality. The chemical compositions of Zr4Pd2O and Zr4Rh2O confirmed through EDX were to be 2.10(2) for Zr:Pd and 2.2(2) for Zr:Rh.

Figure 1
figure 1

Schematic images of the cubic η-carbide crystal structure. (a) Zr4Tr2O (Tr = Pd or Rh) unit cell. (b) Zr1 octahedron centered by O atom. (c) Unit of stella quadrangla lattice. Rietveld refined Room temperature SXRD patterns of (d) Zr4Pd2O and (e) Zr4Rh2O. The red points and cyan lines represent obtained SXRD data and calculated data, respectively. Lower solid lines show the Bragg peak positions. The lower blue lines are differences between obtained SXRD data and calculated data. The impurity contents are shown as mass fraction.

Table 2 Crystalline parameters obtained from Rietveld refinement using SXRD patterns at Room temperature for Zr4Pd2O and Zr4Rh2O.

Magnetization, electrical resistivity, and specific heat

We measured temperature- and magnetic field-dependent magnetizations for Zr4Pd2O and Zr4Rh2O with polished rectangular cuboid samples. The value of a demagnetizing factor N with a rectangular cuboid sample applied a vertical magnetic field can be calculated using dimensional information of the sample: length l, width w, and thickness t39:

$$\begin{array}{c}{N} \, \text{=} \, \frac{4lw}{4lw\,+\,3t\left(l\,+\,w\right)}\text{.}\end{array}$$
(3)

The calculated values of N were 0.66 for Zr4Pd2O (l = 1.40 mm, w = 1.50 mm, t = 0.51 mm) and 0.81 for Zr4Rh2O (l = 0.96 mm, w = 1.37 mm, t = 0.18 mm). An actual magnetic field in samples should be modified to an effective inner magnetic field as described in Heff = H – 4πMN, where H is an applied magnetic field and M is a magnetization. Thus, magnetic susceptibility χ taken to account for the demagnetizing effect is defined as follows:

$$\begin{array}{c}\chi \, \text{=} \, \frac{M}{H_{\text{eff}}} \, \text{=} \, \frac{M}{H\,-\,4\pi MN}\text{.}\end{array}$$
(4)

Figure 2a and b show temperature dependences of the χ for Zr4Pd2O and Zr4Rh2O, respectively. We observed a clear superconducting transition at Tc = 2.8 K for Zr4Pd2O and Tc = 3.5 K for Zr4Rh2O. The temperature width in the superconducting transition for Zr4Rh2O was broader than Zr4Pd2O. Moreover, the Tc of that was lower than the previous report (Tc = 4.3 K) and it would be based on the vacancy of oxygen12. The superconducting state at 1.8 K reached perfect diamagnetism in a zero-field cooling (ZFC) process different from a case for field cooling (FC) process. The hysteresis of temperature-dependent χ with the cooling process reflects the nature of type-II superconductor. Temperature dependence of a lower critical field μ0Hc1 can be obtained through magnetic field dependence of M as shown in Fig. 2c and d. The horizontal axis is displayed in μ0Heff instead of H for precise estimation of the lower critical field. In Zr4Pd2O, the measurement was carried out in a range of 1.8 K < T < 2.8 K with increments of 0.1 K. In the case of Zr4Rh2O, the data was taken at 1.8 K, 1.9 K, and 2.0 K, then taken with increments of 0.2 K up to 3.6 K. We observed the convex downward curve of M as functions of μ0Heff at each temperature, and the minimum points gradually shifted lower filed as increasing temperature. In a low-filed region, the linear responses of M corresponding to the Meissner state were observed, and the behavior can be described as Mfit = a*Heff + b*, where a* and b* are numerical constants. The fitting was carried out in a range of 0 mT < μ0Heff < 5 mT. Figure 2e and f are differences between M and Mfit for Zr4Pd2O and Zr4Rh2O, respectively. The dashed lines represent the value of μ0Heff which deviates from the linear behavior of M. Temperature dependences of lower critical field μ0Hc1(T) collected from the M-Mfit are shown in Fig. 2g and h for Zr4Pd2O and Zr4Rh2O, respectively. The lower critical field at 0 K, μ0Hc1(0) can be obtained from the following empirical formula:

$$\begin{array}{c}{\mu }_{0}{{H}}_{\text{c1}}\text{(}{T}\text{)} \, \text{=} \, {\mu}_{0}{{H}}_{\text{c1}}\left({0}\right)\left[1\,-\,{\left(\frac{T}{T_{\text{c}}}\right)}^{2}\right],\end{array}$$
(5)

resulting μ0Hc1(0) = 11.3 mT and 8.2 mT for Zr4Pd2O and Zr4Rh2O, respectively. Zr4Pd2O showed a higher μ0Hc1(0) than that of Zr4Rh2O even lower Tc, suggesting that Zr4Pd2O tends to be more robust against magnetic field rather than Zr4Rh2O.

Figure 2
figure 2

(a,b) ZFC and FC temperature-dependent magnetic susceptibility under μ0H = 1 mT for (a) Zr4Pd2O and (b) Zr4Rh2O. The insets are enlarged view near Tc. (c,d) Effective inner magnetic field dependence of magnetic susceptibility for (c) Zr4Pd2O and (d) Zr4Rh2O. The solid lines are fit of Mfit = a*Heff + b* in a range of 0 mT < μ0Heff < 5 mT. (e,f) A difference between M and Mfit for (e) Zr4Pd2O and (f) Zr4Rh2O. The dashed lines correspond to the field where M begins to deviate from the linear behavior. The dashed lines are used to determine temperature dependence of the lower critical field. (g,h) Temperature dependence of lower critical field for (g) Zr4Pd2O and (h) Zr4Rh2O. The solid lines are the fit of μ0Hc1(T) = μ0Hc1(0)[1-(T/Tc)2]. The values of μ0Hc1(0) were calculated to be 11.3 mT and 8.2 mT for Zr4Pd2O and Zr4Rh2O, respectively.

Temperature dependences of electrical resistivity ρ(T) for Zr4Pd2O and Zr4Rh2O at zero field are shown in Fig. 3a and b, respectively. In a low-temperature region, we observed a drop of ρ(T) to zero, suggesting a superconducting transition. Zr4Pd2O showed a sharp transition, and the zero resistivity was observed at Tczero = 2.73 K. Zr4Rh2O, however, showed a broad transition, and the zero resistivity was observed at Tczero = 3.73 K. The Tc obtained from the zero resistivity agreed with the result from the magnetic susceptibility measurement. The ρ(T) exhibits a metallic behavior in a normal state for both Zr4Pd2O and Zr4Rh2O. In the low-temperature normal state, the ρ(T) curve can be fitted using the power-law model:

$$\begin{array}{c}\rho \left({T}\right) \, \text{=} \, {\rho }_{0} \, + \, {A}{T}^{n_{\text{PL}}}\text{,}\end{array}$$
(6)

where ρ0, A, and nPL are residual resistivity, temperature-independent coefficient, and power exponent respectively. The ρ(T) curve of Zr4Pd2O in a range of 4 K < T < 80 K was fitted using the model, providing ρ0 = 0.22 mΩ cm, A = 0.00011 mΩ K–2, and nPL = 1.2. For the ρ(T) curve of Zr4Rh2O, we fitted in a range of 6 K < T < 80 K, obtained ρ0 = 0.29 mΩ cm, A = 0.00017 mΩ K-2, and nPL = 1.3. The values of nPL close to 1 suggest that the ρ(T) show linear-temperature dependence in the low-temperature normal state for Zr4Pd2O and Zr4Rh2O, and the behavior was observed in some η-carbide-type superconductors12,13. In a high-temperature region (80 K < T < 300 K) where electron-phonon interaction is a dominant mechanism of electron scattering, the ρ(T) shows a convex upward curve with increasing temperature. A similar trend can be found in many superconductors consisting of d-block element40,41,42,43, and the convex upward curve of ρ(T) in the high-temperature region can be fitted using the following parallel resistor model, yielded by Wiesmann et al.44:

Figure 3
figure 3

(a,b) Temperature dependences of electrical resistivity under zero field for (a) Zr4Pd2O and (b) Zr4Rh2O. The insets are enlarged view near Tc. The solid and dashed lines are fit to parallel resistor model and power-law model, respectively. (c,d) Temperature dependences of electrical resistivity under several magnetic fields for (c) Zr4Pd2O and (d) Zr4Rh2O. The dashed lines represent the 10%, 50%, and 90% criteria to determine temperature dependence of the upper critical field. (e,f) Temperature dependences of total specific heat under several magnetic fields for (e) Zr4Pd2O and (f) Zr4Rh2O. The solid lines are fit to C(T)/T = γ + βT2 + δT4. (g,h) Temperature dependences of electronic specific heat zero field for (g) Zr4Pd2O and (h) Zr4Rh2O. The solid lines are used to estimate Tc and dashed lines represent γ value.

$$\begin{array}{c}\rho \left({T}\right) \, \text{=} \, {\left[\frac{1}{{\rho }_{\text{sat}} \, }+\frac{1}{{\rho }_{\text{ideal}}\left({T}\right)}\right]}^{-1}\text{.}\end{array}$$
(7)

The temperature-independent term ρsat corresponds to the saturation of ρ(T) in high temperature. Fisk and Webb found the saturation of resistivity in A-15 superconductors such as Nb3Sn45, and the saturation can be realized when a mean free path becomes comparable to interatomic separations of the material, called as Ioffe-Regel condition46. The temperature-dependent component ρideal is described with the Bloch-Grüneisen model47:

$$\begin{array}{c}{\rho }_{\text{ideal}}\left({T}\right) \, \text{=} \, {\rho }_{\text{ideal,0}} \, + \, {B}{\left(\frac{T}{{\Theta }_{\text{D}}}\right)}^{n_{\text{BG}}}{\int }_{0}^{\frac{{\Theta }_{\text{D}}}{T}}\frac{{t}^{n_{\text{BG}}} \, }{\left(e^{t}- 1\right)\left(1 - e^{-t}\right)}{\text{dt}}\text{,}\end{array}$$
(8)

where ρideal,0, B, ΘD, and nBG are ideal temperature-independent residual resistivity, temperature-independent coefficient, Debye temperature, and power exponent with the Bloch-Grüneisen model, respectively. The nBG usually takes 2, 3, or 5 depending on the scattering nature. The best fit was obtained when nBG = 5 for both Zr4Pd2O and Zr4Rh2O, and the calculation provided ρsat = 0.52 mΩ cm, ρideal,0 = 0.41 mΩ cm, B = 1.49 mΩ cm, and ΘD = 261 K for Zr4Pd2O, and ρsat = 0.82 mΩ cm, ρideal,0 = 0.49 mΩ cm, B = 1.68 mΩ cm, and ΘD = 198 K for Zr4Rh2O. The ρ0 can be calculated using ρsat and ρideal,0 as given in ρ0 = ρsatρideal,0/(ρsat + ρideal,0), and the obtained values of ρ0 are 0.23 mΩ cm and 0.31 mΩ cm, for Zr4Pd2O and Zr4Rh2O, respectively. These ρ0 values are consistent with those obtained by the power-law model. The fitted curves using the power-law model and parallel resistor model are displayed as dashed lines and solid lines, respectively, in Fig. 3a and b. Resistivity at 300 K, ρ300 K, was found to be 0.32 mΩ cm for Zr4Pd2O and 0.47 mΩ cm for Zr4Rh2O. Residual resistivity ratio, RRR = ρ300 K/ρ0 was calculated to be 1.42 and 1.61 for Zr4Pd2O and Zr4Rh2O, respectively. The small RRR value is also seen in other η-carbide-type superconductors12,13,14, and the poor metallic behavior is a common feature of polycrystalline metallic oxide compounds whose grain-boundary scattering is significant48. Figure 3c and d show the ρ(T) curves at several magnetic fields for Zr4Pd2O and Zr4Rh2O, respectively. The magnetic fields are applied with an increment of 0.2 T for Zr4Pd2O up to μ0H = 3.8 T. For Zr4Rh2O, the magnetic fields are increased by 0.2 T up to μ0H = 4.0 T and then increased by 0.5 T up to μ0H = 6.5 T. The Tczero shifted lower temperature with increasing magnetic field as we expected. We used typical 10%, 50%, and 90% criteria defined with ρ0 to determine temperature dependences of the upper critical field for Zr4Pd2O and Zr4Rh2O (discussed later).

Figure 3e and f show temperature dependences of total specific heat C(T) at several magnetic fields for Zr4Pd2O and Zr4Rh2O, respectively. Magnetic fields were applied with increments of 0.2 T up to μ0H = 2 T and also measured at μ0H = 9 T. We observed clear specific heat jumps, suggesting superconducting transition, up to μ0H = 2 T, and the temperature at which the jumps observed shifted to a lower temperature, consistent with the results from electrical resistivity. Zr4Rh2O showed broader transitions than that of Zr4Pd2O, as seen in magnetic susceptibility and electrical resistivity measurements. To calculate Sommerfeld coefficient γ and ΘD, we fitted C(T) using the following formula:

$$\begin{array}{c}{C}\left({T}\right) \, \text{=} \, \gamma {T} \, + \, \beta {T}^{ 3}+\delta {T}^{5}\text{,}\end{array}$$
(9)

where β and δ are the coefficients of the phonon contributions for the harmonic and anharmonic terms, respectively. As a result of the fitting, we obtained the values of γ to be γ = 32.5 mJ K-2 mol–1 for Zr4Pd2O and γ = 18.1 mJ K–2 mol–1 for Zr4Rh2O. For the phonon contribution coefficients, we obtained β = 1.87 mJ K–4 mol–1 and δ = 0.0016 mJ K-6 mol-1 for Zr4Pd2O, and β = 1.12 mJ K–4 mol–1 and δ = 0.014 mJ K–6 mol–1 for Zr4Rh2O. The fitting curves are shown as solid lines in Fig. 3e and f. The values of ΘD can be calculated using the β as the following formula:

$$\begin{array}{c} \, {\Theta }_{\text{D}} \, \text{=} \, {\left(\frac{{12}{\pi}^{4}{NR}}{{5}\beta}\right)}^{\frac{1}{{3}}}\text{,}\end{array}$$
(10)

where N = 7 is the number of atoms per formula unit and R ≈ 8.31 J K–1 mol–1 is an ideal gas constant. The calculated ΘD was ΘD = 194 K and 230 K for Zr4Pd2O and Zr4Rh2O, respectively, and the values were close to the calculation result obtained by the parallel resistor model in electrical resistivity measurement. Temperature dependences of the electron contribution of the specific heat Cel(T) estimated by subtracting phonon contributions βT3 + δT5 from C(T) are shown in Fig. 3g and h for Zr4Pd2O and Zr4Rh2O, respectively. Tc determined from Cel(T) at zero field was 2.6 K for Zr4Pd2O and 3.3 K for Zr4Rh2O. The normalized jumps of Cel(T), ΔCelTc, were estimated to be 1.58 and 1.57 for Zr4Pd2O and Zr4Rh2O, respectively. The values of the jump were similar and slightly higher than 1.43, which is the expected value by the weak-coupling BCS theory17. This result suggests that Zr4Pd2O and Zr4Rh2O are electron-phonon coupling superconductors with a little strong-coupling nature. We can calculate an electron-phonon coupling constant λel-ph using the McMillan formula49:

$$\begin{array}{c}{\lambda }_{\text{el-ph}} \, \text{=} \, \frac{1.04\,+\,\mu^{\text{*}}{\text{ln}}\left(\frac{{\Theta }_{\text{D}}}{1.45{T}_{\text{c}}}\right)}{\left(1\,-\,0.62\mu^{\text{*}}\right){\text{ln}}\left(\frac{{\Theta }_{\text{D}}}{1.45{T}_{\text{c}}}\right)\,-\,1.04}\text{,}\end{array}$$
(11)

where μ* = 0.13 is a Coulomb coupling constant and the value is used empirically for similar materials containing transition metals. We obtained the values of λel-ph to be 0.60 for Zr4Pd2O and 0.61 for Zr4Rh2O. An electronic density of states at the Fermi energy D(EF) is proportional to a term (1 + λel-ph) when we consider the electron–phonon coupling. Therefore, D(EF) with spin degeneracy can be expressed in the following:

$$\begin{array}{c}{D}\left({E}_{\text{F}}\right) \, \text{=} \, \frac{3\gamma }{{\uppi }^{2}{k}_{\text{B}}^{2}\left(1\,+\,\lambda_{\text{el-ph}}\right)}.\end{array}$$
(12)

The measured γ and calculated λel-ph provide D(EF) = 8.61 states eV–1 per formula unit (f.u.) and 4.76 states eV–1 per f.u. for Zr4Pd2O and Zr4Rh2O, respectively. The higher Tc of Zr4Rh2O than that of Zr4Pd2O may be based on higher ΘD as explained in BCS theory17.

Discussion

Here, we discuss the upper critical fields and other superconducting parameters of Zr4Pd2O and Zr4Rh2O. Figure 4a and b are temperature dependences of upper critical field μ0Hc2(T) for Zr4Pd2O and Zr4Rh2O, respectively. The data points were taken from temperature dependences of ρ(T) with 10%, 50%, and 90% criteria, and C(T) under several magnetic fields. The upper critical field at 0 K, μ0Hc2(0) can be calculated by fitting the data using the Ginzburg-Landau (GL) model:

Figure 4
figure 4

(a,b) Temperature dependence of upper critical field for (a) Zr4Pd2O and (b) Zr4Rh2O. The solid lines are fit to the GL model. The value of μ0HP was calculated using μ0HP = 1.86Tc with ρ(T) 50% criteria data.

$$\begin{array}{c}{\mu }_{0}{{H}}_{\text{c2}}\left({T}\right) \, \text{=} \, {\mu }_{0}{{H}}_{\text{c2}}\left({0}\right)\left[\frac{\text{1}\,-\,{\left(\frac{T}{T_{\text{c}}}\right)}^{2}}{\text{1}\,+\,{\left(\frac{T}{T_{\text{c}}}\right)}^{2}}\right]\text{.}\end{array}$$
(13)

We obtained the values of μ0Hc2(0) for Zr4Pd2O to be 7.18 T for ρ(T) 10% criterion, 6.88 T for ρ(T) 50% criterion, 6.72 T for ρ(T) 90% criterion, and 9.17 T for C(T). For Zr4Rh2O, the obtained μ0Hc2(0) to be 6.27 T for ρ(T) 10% criterion, 6.16 T for ρ(T) 50% criterion, 5.91 T for ρ(T) 90% criterion, and 7.74 T for C(T). We found that the whole values of μ0Hc2(0) for Zr4Pd2O were higher than that of μ0HP = 5.29 T calculated with Tc of ρ(T) 50% criterion. On the other hand, for Zr4Rh2O, the values of μ0Hc2(0) derived from ρ(T) criterion were lower than that of μ0HP = 7.59 T calculated with Tc of ρ(T) 50% criterion. The value of μ0Hc2(0) derived from C(T) was close to the μ0HP. The absence of violation of the Pauli limit for Zr4Rh2O is consistent with the previous study12. The violation of the Pauli limit observed in Zr4Pd2O is an unreported superconducting nature, and a similar violation was reported in other η-carbide-type superconductors as mentioned in the Introduction part. A quasiparticle mean free path l at a normal state near the superconducting state can be estimated using the following formula derived from Singh et al.50:

$$\begin{array}{c}{l} \, \text{=} \, 2.372 \, \times {10}^{-14}\frac{{\left(\frac{{m}^{\text{*}}}{m_{\text{e}}}\right)}^{2}{V_{\text{M}}^{2}}}{{D}{\left({E}_{\text{F}}\right)}^{2}{\rho}_{0}}\text{,}\end{array}$$
(14)

where me, m*, and VM are free-electron mass, effective mass of the individual quasiparticles, and molar volume. The l is in cm unit when we take VM, D(EF), and ρ0 are in cm3 mol–1, states eV–1 per f.u., and Ω cm, respectively. If we assume m*/m = 1, we obtain l = 0.76 Å for Zr4Pd2O using VM = 72.8 cm3 mol–1, D(EF) = 8.61 states eV-1 per f.u. and ρ0 = 0.22 mΩ cm. Likewise for Zr4Rh2O, we obtain l = 1.84 Å using VM = 71.7 cm3 mol–1, D(EF) = 4.76 states eV-1 per f.u., and ρ0 = 0.29 mΩ cm. A GL coherence length ξGL can be calculated using the GL model with μ0Hc2(0) as the following:

$$\begin{array}{c}{\mu }_{0}{H}_{\text{c2}}\left({0}\right)\text{=}\frac{{\Phi }_{0}}{{2}\pi {\xi }_{\text{GL}}^{2}}\text{,}\end{array}$$
(15)

where Φ0 ≈ 2.07 ×10-15 Wb is a magnetic flux quantum. The values of ξGL for Zr4Pd2O and Zr4Rh2O were calculated to be 69 Å and 73 Å, respectively, using the μ0Hc2(0) obtained from ρ(T) 50% criterion. The values of ξGL were found to be much longer than that of l for both Zr4Pd2O and Zr4Rh2O. Therefore, both Zr4Pd2O and Zr4Rh2O are supposed to be in the dirty limit. The orbital limit μ0Horb can be estimated with WHH theory without considering spin-orbit scattering28,29. In the dirty limit, μ0Horb is expressed in the following formula:

$$\begin{array}{c}{\mu }_{0}{{H}}_{\text{orb}} \, \text{=} \, -0.693{T}_{\text{c}}{\left.\frac{{d}{\mu }_{0}{H}_{\text{c2}}\left({T}\right)}{dT}\right|}_{{T}\,=\,{T}_{\text{c}}}\text{.}\end{array}$$
(16)

The slope of μ0Hc2(T) at Tc was estimated to be − 2.72 TK–1 and − 1.81 TK-1 for Zr4Pd2O and Zr4Rh2O, respectively, when using the μ0Hc2(T) data of the ρ(T) 50% criterion. These obtained values yielded μ0Horb = 5.36 T for Zr4Pd2O and 5.11 T for Zr4Rh2O. For Zr4Pd2O, the value of μ0Hc2(0) determined by ρ(T) 50% criterion was found to be larger than that of both μ0Hp and μ0Horb. On the other hand, for Zr4Rh2O, the value of μ0Hc2(0) determined by ρ(T) 50% criterion was higher than that of μ0Horb but lower than that of μ0Hp. The enhanced μ0Hc2(0) larger than both μ0Hp and μ0Horb for Zr4Pd2O implies the importance of spin-orbit scattering caused by strong SOC because the strong SOC can suppress the Pauli paramagnetic pair-breaking effect28,29 and calculation of μ0Horb in Eq. (16) does not consider the SOC. The importance of SOC was pointed out by Ruan et al.32 and Shi et al.33 in Ti4Ir2O, exhibiting the violation of the Pauli limit. Similarly, we can expect that the enhanced μ0Hc2(0) of Zr4Pd2O would be raised from strong SOC. The strength of SOC proportions to Z4 within the same electronic orbital as expressed in Eq. (2), therefore the absence of enhanced μ0Hc2(0) for Zr4Rh2O may be explained by the lower strength of SOC because of the number of d electron configuration: Rh consists of 4d8, but Pd consists of 4d10. For deeper understanding the observed enhanced μ0Hc2(0) in Zr4Pd2O, further investigations such as density functional theory (DFT) calculation considered SOC effect and measurement of electronic state are needed. A GL penetration depth λGL can be obtained using μ0Hc1 and ξGL in the following formula:

$$\begin{array}{c}{\mu }_{0}{{H}}_{\text{c1}}\text{(}{0}\text{)} \, = \, \frac{{\Phi }_{0}}{{4}\pi {\lambda }_{\text{GL}}^{2}}{\text{l}}{\text{n}}\left(\frac{{\lambda }_{\text{GL}}}{{\xi }_{\text{GL}}}\right)\text{.}\end{array}$$
(17)

We obtained the values of λGL to be 2250 Å for Zr4Pd2O and 2697 Å for Zr4Rh2O. GL parameters κGL = λGLGL were estimated to be 33 and 37 for Zr4Pd2O and Zr4Rh2O, respectively. The calculation results agree with the nature of the type-II superconductor, shown in Fig. 2a and b. A thermodynamic critical field μ0Hc can be estimated using the following expression:

$$\begin{array}{*{20}c} {\mu _{0} {H}_{{{\text{c1}}}} {\text{(}}0{\text{)}} \cdot {\mkern 1mu} \mu _{0} {H}_{{{\text{c2}}}} {\text{(}}0{\text{)}}{\mkern 1mu} = {\mkern 1mu} \left[ {\mu _{0} {H}_{{\text{c}}} {\text{(0)}}} \right]^{2} {\text{ln}}\kappa _{{{\text{GL}}}} {\text{.}}} \\ \end{array}$$
(18)

The calculation provided the values of μ0Hc(0) to be 150 mT and 118 mT for Zr4Pd2O and Zr4Rh2O, respectively. Finally, we summarized the whole obtained superconducting properties in Table 3.

Table 3 The measured superconducting properties of Zr4Pd2O and Zr4Rh2O.

In summary, we have discovered the bulk superconductivity in Zr4Pd2O. The crystal structure was found to be the η-carbide-type structure with a space group Fd \(\stackrel{\mathrm{-}}{3}\) m (No. 227) through SXRD measurement. The bulk superconductivity was measured by magnetic susceptibility, electrical resistivity, and specific heat measurement, resulting in Tc = 2.8 K, 2.73 K, and 2.6 K, respectively. Zr4Pd2O was found to belong to the type-II superconductor by magnetic susceptibility measurement in ZFC and FC processes. The upper critical field was determined from electrical resistivity and specific heat data under several magnetic fields. We found that Zr4Pd2O exhibited an enhanced upper critical field μ0Hc2(0) = 6.72 T violating the Pauli limit μ0HP = 5.29 T, whereas the absence of the property in isostructural η-carbide-type oxide superconductor Zr4Rh2O. The enhanced upper critical field can be raised from strong SOC. DFT calculation and measurement of electronic state are future works for further understanding the enhanced upper critical field.

Methods

Sample preparation

Polycrystalline samples of Zr4Pd2O and Zr4Rh2O were prepared by reaction of the Zr plate (99.2%, Nilaco Corporation), ZrO2 powder (98.0%, Wako Special Grade), Rh powder (99.9%, Kojundo Chemical), and Pd powder (99.9%, Kojundo Chemical). These starting materials were weighed to a stoichiometric ratio, and the powders of that were pressed into a pellet. At first, the obtained pellet and Zr plate were melted together by means of an arc melting method on a water-cooled copper stage. Gas inside the arc furnace was replaced by pure argon gas 3 times and then filled with pure argon gas. Before melting the sample, a titanium ingot was melted to reduce residual oxygen gas in the furnace. The sample was melted at least 6 times and turned over at each melting for homogeneity. We observed a negligible 1–2% mass loss after the melting. Second, we crushed the as-cast sample into fine powder and pressed it into a pellet. Subsequently, we sealed the pellet into an evacuated quartz tube and treated an annealing process for 10 days at 800 ℃. A mass loss was not observed after the annealing, implying oxygen in the sample was maintained.

Crystal structure and composition

The phase purity and crystal structure of Zr4Pd2O and Zr4Rh2O were checked by XRD with Cu-Kα radiation using θ-2θ method. The XRD measurement was performed on a Miniflex 600 (Rigaku) equipped with a high-resolution semiconductor detector D/tex-Ultra. For further investigation, we also performed SXRD measurement at the beamline BL13XU in SPring-8 (proposal no. 2023B1669) with a wavelength of λ = 0.354367 Å. The obtained SXRD patterns were refined by means of the Rietveld method using RIETAN-FP51. Schematic images of the crystal structure were depicted using VESTA52. The chemical compositions of Zr and Tr (Pd or Rh) were examined by EDX on a scanning electron microscope TM-3030plus (Hitachi High-Tech) equipped with computer software SwiftED (Oxford). The chemical composition of oxygen was not considered because of the difficulty of detecting light elements with X-ray spectroscopy.

Measurement of superconducting properties

Temperature and magnetic field dependence of magnetization were measured using a superconducting quantum interference device (SQUID) on a Magnetic Property Measurement System 3 (MPMS3, Quantum Design) equipped with a 7 T superconducting magnet. The measurement was performed using a vibrating sample magnetometry (VSM) mode with polished rectangular cuboid samples to estimate precise demagnetizing factors. The samples were placed in a vertically applied magnetic field. Temperature dependence was measured under μ0H = 1 mT in both zero-field cooling (ZFC) and field cooling (FC) processes. The magnetic field dependence was measured up to μ0H = 30 mT at several temperatures. Temperature and magnetic field dependence of Electrical resistivity and specific heat measurements were performed using a physical property measurement system (PPMS Dynacool, Quantum Design) equipped with a 9 T superconducting magnet. Electrical resistivity was measured by a four-probe DC method using silver paste and gold wires for the contact between a polished rectangular cuboid sample and sample puck. The measurement was performed using an excitation current of 1 mA. The specific heat measurement was carried out by means of a thermal relaxation method. The sample was mounted on a stage with N-grease for good thermal connection.