† Corresponding author. E-mail:
Project supported by the National Natural Science Foundation of China (Grant Nos. 11674377 and 11634015), the National Key R&D Program of China (Grant Nos. 2017YFA0302904 and 2016YFA0300502), and the Strategic Priority Research Program (B) of the Chinese Academy of Sciences (Grant No. XDB07020200). J. Y. is supported by the Youth Innovation Promotion Association of CAS.
The interplay between superconductivity and structural phase transition has attracted enormous interest in recent years. For example, in Fe-pnictide high temperature superconductors, quantum fluctuations in association with structural phase transition have been proposed to lead to many novel physical properties and even the superconductivity itself. Here we report a finding that the quasi-skutterudite superconductors (Sr1−xCax)3Ir4Sn13 (x = 0, 0.5, 1) and Ca3Rh4Sn13 show some unusual properties similar to the Fe-pnictides, through 119Sn nuclear magnetic resonance (NMR) measurements. In (Sr1−xCax)3Ir4Sn13, the NMR linewidth increases below a temperature T* that is higher than the structural phase transition temperature Ts. The spin-lattice relaxation rate (1/T1) divided by temperature (T), 1/T1T and the Knight shift K increase with decreasing T down to T*, but start to decrease below T*, and followed by more distinct changes at Ts. In contrast, none of the anomalies is observed in Ca3Rh4Sn13 that does not undergo a structural phase transition. The precursory phenomenon above the structural phase transition resembles that occurring in Fe-pnictides. In the superconducting state of Ca3Ir4Sn13, 1/T1 decays as exp(−Δ/kBT) with a large gap Δ = 2.21kBTc, yet without a Hebel–Slichter coherence peak, which indicates strong-coupling superconductivity. Our results provide new insight into the relationship between superconductivity and the electronic-structure change associated with structural phase transition.
Transition-metal compounds show diverse properties such as magnetism, superconductivity, and charge density wave, and often accompany a structural transition.[1–3] In these materials, the interplay between superconductivity and other orders is of great interest. For example, in the copper oxides,[4] heavy fermions,[5] and iron-based superconductors[6] that contain transition metal elements, superconductivity is found in the vicinity of a quantum critical point (QCP) at which other orders are completely suppressed at absolute zero temperature. In particular, in iron-based superconductors, not only a magnetic (spin density wave) QCP, but also another QCP associated with the structural phase transition exists.[7] In this case, quantum fluctuations of the electronic nematic order associated with the structural phase transition may lead to many novel physical properties such as T-linear electrical resistivity.[7–10]
Materials with the general stoichiometry R3M4X13 are a large family usually adopting a common quasi-skutterudite structure, where R is an alkaline-earth or rare-earth element, M is a transition metal, and X is a group-IV element.[11,12] Superconductivity with a fairly high transition temperature Tc ∼ 7 K was found in R3M4Sn13 more than 30 years ago,[13,14] but the physical properties were poorly understood. Recently, this class of materials received new attention because of a possible interplay between the superconductivity and the structure instability.
The electrical resistivity, susceptibility, Hall coefficient, and heat transport measurements on Ca3Ir4Sn13 found that an anomaly occurs at a temperature of 35 K, above the superconducting transition temperature Tc = 7 K.[15–17] The anomaly was ascribed to ferromagnetic spin fluctuation in early works.[15] Resistivity and susceptibility measurements on Sr3Ir4Sn13 also showed anomalies at 147 K. Subsequent x-ray diffraction and pressure effect measurements on Sr3Ir4Sn13 showed that a structural phase transition from a cubic I phase (
The nature of the electronic structure change due to the structural phase transition is not well understood. Neutron scattering and specific heat measurements revealed a second-order nature of the structural phase transition in (Sr1−xCax)3Ir4Sn13.[21,22] The Hall coefficient changed from a negative to a positive value and the optical measurement indicated that the Drude spectral weight is transferred to the high energy region across Ts in Sr3Ir4Sn13.[23,24] Based on these results, a reconstruction of the Fermi surface below Ts due to a charge density wave (CDW) formation was suggested.[23,24]
In this work, we grow single crystals of (Sr1−xCax)3Ir4Sn13 (x = 0, 0.5, 1) and Ca3Rh4Sn13, and perform electrical resistivity and 119Sn NMR measurements to elucidate the electronic properties change associated with the structural phase transition. (Sr1−xCax)3Ir4Sn13 (x = 0, 0.5, 1) undergo a structural phase transition at Ts = 147 K, 85 K, and 35 K, respectively, while Ca3RH4Sn13 does not. By NMR measurements, we find that an anomaly occurs already above Ts in (Sr1−xCax)3Ir4Sn13 (x = 0, 0.5, 1). Such an electronic anomaly prior to the structural transition resembles an actively-investigated phenomenon in some of the Fe-based superconductors where the physical properties show an in-plane anisotropy (nematicity) above Ts below which the C4 symmetry is lowered to the C2 symmetry. However, in Ca3RH4Sn13 that does not undergo a structural transition, the Korringa relation is satisfied down to T ∼ 20 K. The electronic state properties below T* are discussed by analyzing the change in the Korringa ratio. We also measure the superconducting state property of Ca3Ir4Sn13, and find that it is a strong-coupling s-wave superconductor. We will discuss the relationship between the superconductivity, the electronic state change associated with the structural transition, and electron correlations.
Single crystals of (Sr1−xCax)3Ir4Sn13 and Ca3RH4Sn13 were grown by the self-flux method, as previously reported in Ref. [13]. The composition shown in this paper is the nominal one. Excessive Sn flux was removed in concentrated HCl acid. Crystals with the proper size were picked up and polished, then the temperature dependence of the resistivity was measured by the standard four-probe method using a physical properties measurement system (PPMS, Quantum Design). The Tc was determined by both DC susceptibility using a magnetic properties measurement system (MPMS, Quantum Design) with an applied magnetic field of 10 Oe, and AC susceptibility using an in-situ NMR coil. For 119Sn NMR measurements, since the sample shows a good electrical conductivity so that the skin depth is short, we crushed the single crystals into fine powders to gain the surface area. The 119Sn nucleus has a nuclear spin I = 1/2 and gyromagnetic ratio γn/2π = 15.867 MHz/T. The 119Sn NMR spectra were obtained by scanning the rf frequency and integrating the spin echo intensity at a fixed magnetic field H0. The spin-lattice relaxation time T1 was measured by using the saturation-recovery method, and obtained by a good fitting of the nuclear magnetization M(t) to 1 − M(t)/M0 = exp(−t/T1), where M(t) is the nuclear magnetization at time t after the single saturation pulse and M0 is the nuclear magnetization at thermal equilibrium.
Figure
Since 119Sn (I = 1/2) has no quadrupole moment, the nuclear spin Hamiltonian is simply given by the Zeeman interaction
Below we discuss the normal-state properties inferred from the NMR measurements. Figure
Figure
The total Knight shift consists of three parts, K = Kdia + Korb + Ks, where Kdia arises from the diamagnetic susceptibility χdia, Korb from the orbital (Van-Vleck) susceptibility χorb, and Ks from the spin susceptibility χs. The χs is estimated to be 8.1 × 10−4 emu/mol according to
In general, 1/T1T probes the transverse imaginary part of the dynamic susceptibility (
The anomaly seen in 1/T1T and K at T* resembles a puzzling phenomenon in the Fe-based superconductors such as BaFe2−xCoxAs2, BaFe2(As1−xPx)2, or NaFe1−xCoxAs,[27–29] where nematic properties already appear at a temperature far above Ts. In this class of Fe-based materials, a C4 to C2 structural phase transition place at Ts. It also shares some similarities with the pseudogap behaviors in underdoped copper oxide superconductors, where the DOS starts to decrease before the superconducting phase transition.[30] In any event, the precursory electronic anomaly above the Ts suggests that the structural phase transition is electronically-driven, rather than lattice-driven.
Next, we examine if the electronic-state change below T* can be totally ascribed to a loss of the DOS. In Fig.
For (Sr1−xCax)3Ir4Sn13 (x = 0, 0.5, 1), however, the temperature dependence of (T1T)−1/2 and K is weak at high temperatures so that a similar plot does not yield meaningful information. Instead we plot the so-called Korringa ratio S as a function of temperature in Fig.
Firstly, a second-order phase transition often accompanies the development of a short-range correlation just above the transition temperature, thus the structural instability may be responsible for the anomaly seen in our NMR data. Theoretical calculations of phonon dispersion suggest that imaginary phonon modes exist in Sr3Ir4Sn13 and the lattice instabilities lie at some wave vectors.[26] Neutron scattering data have shown the softening of phonon mode towards Ts.[21] Specific heat measurements on Sr3Ir4Sn13 also show that ΔC/T starts to increase at 160 K (Ts = 147 K) and the critical fluctuation model can fit the specific heat data well, which leads to the proposal of short-range correlation above Ts.[22] Therefore, the NMR quantities may also be affected by such structural short-range correlation through magneto–elastic coupling, resulting in the deviation from the Korringa relation below T*. On the other hand, we note that, for a CDW case, the quantity 1/T1T will increase with decreasing temperature towards the transition temperature,[31,32] in contrast to a decrease of 1/T1T observed in the present case.
Secondly, we discuss the possibility of magnetic correlations. To explore this issue in more detail, we turn to the data at temperatures below T = 20 K. In Fig.
In Fig.
As the pressure increases, Ts and T* are suppressed while Tc increases slowly, which means that there exists a competition between the structural phase transition and superconductivity. A similar phase diagram has been seen in other systems such as LaPt2−xGe2+x,[37] where Tc increases from 0.41 K to 1.95 K and Ts decreases from 394 K to 50 K. Note that the Knight shift and thus the electronic DOS remain T-independent at low temperatures, while its absolute value increases from Sr3Ir4Sn13 to Ca3RH4Sn13. Therefore, the increase of DOS may be partly responsible for the increase of Tc.
Ca3RH4Sn13 is located near the end point of the T* curve, and its superconducting transition temperature Tc = 8 K is close to the highest value of this class of materials under chemical or physical pressures. In cuprates and Fe-based superconductors, Tc has a close connection to magnetic fluctuations or structural/orbital fluctuations. Although it is not clear at the moment how the antiferromagnetic spin fluctuations found in this work are related to the structural phase transition, the antiferromagnetic spin fluctuations may also contribute to the increase of Tc. In fact, a systematic change of the Korringa ration S is found as the chemical pressure is increased, as seen in Fig.
In this section, we discuss the properties of Ca3Ir4Sn13 in the superconducting state. In order to minimize the effect of the external field, we have performed NMR measurements for the Sn(2) site under a low field of H0 = 0.4 T that is much smaller than Hc2 ≈ 7 T. The Tc is 6.4 K at H0 = 0.4 T. The 119Sn Knight shift below Tc(H) is shown in Fig.
Figure
We have grown single crystals of (Sr1−xCax)3Ir4Sn13 (x = 0, 0.5, 1) and Ca3RH4Sn13, and performed electrical resistivity and 119Sn NMR measurements. In the normal state, we found an anomaly at T* above the structural phase transition temperature Ts in (Sr1−xCax)3Ir4Sn13 (x = 0, 0.5, 1). The NMR line width increases below T* and 1/T1T and K begin to decrease, followed by more distinct changes at Ts. None of these anomalies was observed in Ca3RH4Sn13 that does not undergo a structural phase transition. Our detailed analysis of (Sr1−xCax)3Ir4Sn13 suggests antiferromagnetic spin fluctuations developing below T* and becoming more visual below T ∼ 20 K as a possible cause. The increase of Tc from Sr3Ir4Sn13 to Ca3RH4Sn13 can be partly ascribed to the increase of the electronic DOS, but the antiferromagnetic spin fluctuations may also make a contribution. Remarkably, the precursory electronic anomaly shares a similarity with a phenomenon under active investigation in the Fe-based high-Tc superconductors where a change in the electronic properties expected at T* occurs already above Ts. Therefore, our work sheds light on other correlated electron systems in a broad context.
In the superconducting state of Ca3Ir4Sn13, the spin susceptibility vanishes at low temperature, indicating a spin-singlet electron pairing. The spin-lattice relaxation rate decays exponentially with decreasing temperature as exp(−Δ/kBT), which indicates a fully opened energy gap. The large superconducting gap 2Δ = 4.42kBTc accompanied by a lack of the coherence peak indicates that Ca3Ir4Sn13 is a strong-coupling superconductor.
[1] | |
[2] | |
[3] | |
[4] | |
[5] | |
[6] | |
[7] | |
[8] | |
[9] | |
[10] | |
[11] | |
[12] | |
[13] | |
[14] | |
[15] | |
[16] | |
[17] | |
[18] | |
[19] | |
[20] | |
[21] | |
[22] | |
[23] | |
[24] | |
[25] | |
[26] | |
[27] | |
[28] | |
[29] | |
[30] | |
[31] | |
[32] | |
[33] | |
[34] | |
[35] | |
[36] | |
[37] | |
[38] | |
[39] | |
[40] | |
[41] | |
[42] | |
[43] | |
[44] | |
[45] |