Large magnetoresistance and thermoelectric properties of a quasi-skutterudite Ca3Pt4Sn13 single crystal

Yan Sun a, Chenguang Li a, Hui Liang a, Yiyan Wang a, Qingqing Luo d, Dingfu Shao d, Xiaoyu Yue b, Na Li a, Ying Zhou a, Qiuju Li c, Dandan Wu a, Xuan Luo d, Nan Zhou *d and Xuefeng Sun *a
aAnhui Provincial Key Laboratory of Magnetic Functional Materials and Devices, Institutes of Physical Science and Information Technology, Anhui University, Hefei 230601, China. E-mail: xfsun@ahu.edu.cn
bSchool of Optical and Electronic Information, Suzhou City University, Suzhou 215104, P. R. China
cSchool of Physics and Optoelectronic Engineering, Anhui University, Hefei 230601, P. R. China
dKey Laboratory of Materials Physics, Institute of Solid State Physics, HIPS, Chinese Academy of Sciences, Hefei 230031, China. E-mail: zhounan@issp.ac.cn

Received 13th November 2024 , Accepted 6th April 2025

First published on 7th April 2025


Abstract

We conducted a detailed study of the quasi-skutterudite Ca3Pt4Sn13 single crystal through measurements of electrical resistivity, Hall effect, specific heat, and thermoelectric properties. The resistivity exhibited metallic behavior from 300 to 2 K. The results of resistivity, the Wilson ratio, and low-temperature magnetic susceptibility indicated non-Fermi liquid behavior in the Ca3Pt4Sn13 crystal. Ca3Pt4Sn13 crystals showed a large longitudinal magnetoresistance (MR) of up to 200% at 5 T and 2 K for a magnetic field applied along the [110] direction. Combined with first-principles calculation results, we believe that carrier mobility is an important factor contributing to the large MR phenomenon. Meanwhile, we also performed angle-resolved MR (AMR) measurements on the Ca3Pt4Sn13 crystal. An eight-fold symmetry in the AMR was observed in the low-temperature region, indicating that Ca3Pt4Sn13 could be a potential candidate for valleytronics. Furthermore, the negative slope and linear dependence observed in the Hall resistance indicated electron conductivity, while the thermoelectric power results suggested the presence of hole-like carriers. The different signs of the Hall and Seebeck coefficients could be related to the unique structure of the Fermi surface boundaries in the Ca3Pt4Sn13 crystal. In conjunction with the above experimental observations, we found that the carrier mobility, carrier concentration, Hall coefficient, Seebeck coefficient, thermal conductivity, magnetic susceptibility, MR, and AMR all showed significant changes when the temperature dropped below T*, indicating that they might share a common origin closely related to the T* phase. The current findings could provide a non-superconducting system to investigate the intrinsic origin of T* in the Ca3T4X13 family.


Due to many potential applications in hard drives and magnetic sensors, new large magnetoresistance (MR) materials have attracted significant attention over the last few decades.1–4 The giant MR (GMR) in magnetic multilayers and colossal MR (CMR) in perovskite-manganites have been widely exploited.5–7 Recently, an extremely large MR (XMR), up to 106%, without magnetic ions was found in Dirac semimetals, Weyl semimetals, and transition metal dipnictides such as TPn2 (T = W, Mo, Nb, and Ta; Pn = P, As, Sb, and Bi),8–13 as well as in rock salt rare-earth compounds like LaBi/Sb, and others.14 In recent years, the mechanism of XMR has been explored through both theoretical and experimental approaches.13–16 A few different mechanisms have been proposed for the origin of XMR in materials lacking magnetic ions.17 They can be summarized as follows: (i) the origin of XMR in DSMs (Cd3As2, Na3Bi, etc.) is related to the unique band structure, which exhibits linear dispersion in momentum space;18–20 (ii) for LaBi/Sb and the WSMs of WTe2 and Td-MoTe2, electron–hole (e–h) compensation also plays an important role in the formation of XMR;14 (iii) for TaTe2 and TaNiTe5, which have a trivial band structure, the carriers show high mobility at low temperatures.21,22

R3T4X13 series compounds, where R is a rare earth element, Sc, Y, or the alkaline earth metals, Ca or Sr; T is a Group VIII d-electron element; and X is either In, Ge, or Sn, are caged-structure compounds.23–33 These compounds typically crystallize in the same primitive cubic structure with the space group Pm[3 with combining macron]n and exhibit remarkably different properties depending on the combination of the three elements. Hence, R3T4X13 compounds with caged structures have attracted significant attention, as many interesting new phenomena, such as superconductivity, the Kondo effect, and good thermoelectric properties, have been observed. Among these quasi-skutterudite materials, Ca3Ir4Sn13 displays an anomaly in the resistivity at T* ∼ 38 K and a superconducting transition at Tc ∼ 7 K.23–25 Some studies attributed the anomaly at T* to ferromagnetic spin fluctuations,23 while another study proposed a charge-density wave (CDW) scenario at T*.24 To date, the microscopic origin of T* and its relationship with superconductivity remains an open issue. Meanwhile, superconductivity around 5 K, accompanied by clear anomalies around T* ∼ 147 K, has also been observed in a similar compound, Sr3Ir4Sn13.26–28 Both compounds provide a platform to study the relationship between quantum phase transitions and superconductivity. Furthermore, the quasi-skutterudite compound, Y3Ir4Ge13, with a large Seebeck coefficient (S), is a promising candidate for thermoelectric devices.29 Owing to the strong hybridization of the localized f-electrons of R ions with the p-electrons of X, Yb3Os4Ge13 shows a Kondo effect at low temperatures.30 The R3T4X13 series compounds provide a platform to study novel physical phenomena in materials with a cage structure. Therefore, exploring new quasi-skutterudite R3T4X13 compounds is of great interest.

Among the R3T4X13 family, the alkaline earth metals, Ca and Sr intermetallics, are of particular interest because of their novel properties, such as superconductivity and quantum critical behavior.23–28 Compared with the rare earth metals, alkaline earth metals, lacking the collective hybridization of localized f-electrons with the conduction electrons, may lead to systems that are easier to understand. For example, the detailed phase diagram of Ca3Ir4Sn13 has been studied under applied pressure and chemical changes induced by Sr ion substitution for Ca ions.31–33 Additionally, materials with large spin–orbital (SO) coupling effects have always attracted significant attention in condensed matter physics.34–39 Usually, the SO coupling is proportional to the atomic number Z2,37 and exploring materials with larger atomic numbers could lead to stronger SO coupling effects. The controllable SO coupling effect could provide a useful way to study topological insulators, such as Bi2(Se/Te)3,36 the quasi-one-dimensional superconductor Nb2PdS5,37 the Weyl semimetal WTe2,38,39 and others. Therefore, studying and comparing the electrical and thermal transport properties in new quasi-skutterudite compounds with large SO coupling effects will be highly valuable.

In order to further explore the physical properties of quasi-skutterudite Ca3T4Sn13-based materials, studied the isovalent substitution effect by introducing Pt onto the Ir site in the Ca3Ir4Sn13 system. Compared with the Ir element, the Pt element has a larger atomic number, indicating a stronger SO coupling effect. Subsequently, we performed a detailed investigation of the Ca3Pt4Sn13 crystal through measurements of electrical resistivity, specific heat, and thermoelectric properties. It exhibited metallic behavior from 300 K to 2 K. The results of the resistivity, Wilson ratio, and low-temperature magnetic susceptibility indicated that Ca3Pt4Sn13 was a non-Fermi liquid system. The MR at T = 2 K with an applied magnetic field of H = 5 T was 200% for H along the [110] direction. Based on the Hall resistance behavior, the origin of the large MR was closely related to the high mobility of the carriers. An eight-fold symmetry in the AMR was observed in the low-temperature region. Both the results of the first-principles calculations and the different signs of the Hall and Seebeck coefficients indicated the presence of a special structure in the Fermi surface boundaries of the Ca3Pt4Sn13 system. Furthermore, we found that the mobility of the carriers, carrier concentration, Hall coefficient, Seebeck coefficient, thermal conductivity, magnetic susceptibility, MR, and AMR all showed significant changes when the temperature dropped below T*.

Ca3Pt4Sn13 single crystals were grown using the flux method with Sn element as the self-flux. The starting materials of Ca (99.99%), Pt (99.99%), and Sn (99.999%) were mixed in a molar ratio of 3[thin space (1/6-em)]:[thin space (1/6-em)]4[thin space (1/6-em)]:[thin space (1/6-em)]93 and placed in alumina ampoules. All operations were performed in a glove box filled with high-purity argon. The alumina ampoules were then seeded into evacuated quartz tubes and heated in a furnace at 1050 °C for 10 h. Subsequently, the tubes were cooled to 600 °C at a rate of 2 °C h−1. The quartz tubes were then inverted and spun rapidly in a centrifuge to separate the Sn flux. Crystals with approximate dimensions of 3 × 2.5 × 2.5 mm3 (shown in Fig. 1(a)) were left in the alumina ampoule. The crystals typically exhibit a larger surface with a clean (110) face. The chemical composition of the crystals was analyzed by energy dispersive X-ray (EDX) spectroscopy using a scanning electron microscope (SEM). X-ray diffraction (XRD) experiments were performed on a PANalytical X’pert diffractometer using Cu Kα1 radiation (λ = 0.15406 nm) at room temperature. A large single crystal with a (110) surface was used for the measurements. Electrical resistivity was measured with the electric current along the (110) surface using a physical property measurement system (Quantum Design, PPMS-9 T, 2 K ≤ T ≤ 400 K, 0 T ≤ H ≤ 9 T). Resistivity was measured in cooling mode using a standard four-probe technique. Thermal conductivity and thermoelectric power measurements were also performed in the PPMS. The specific heat was measured using the thermal relaxation method in the PPMS. First-principles calculations were performed using the Vienna ab initio simulation package (VASP).40,41 The exchange–correlation potential was described using the Perdew–Burke–Ernerhof (PBE) formula within the generalized gradient approximation (GGA).42 The kinetic energy cutoff was set to 500 eV, and the K-point grid was set to 9 × 9 × 9 to sample the irreducible Brillouin zone (BZ). Spin–orbit coupling was included in the calculation.


image file: d4tc04812a-f1.tif
Fig. 1 (a) Crystal structure of Ca3Pt4Sn13; (b) photograph of the Ca3Pt4Sn13 single crystal used in this study. The large surfaces are indexed based on XRD measurements. The size is approximately 3 × 2.5 × 2.5 mm3; (c) XRD patterns of the crystal measured on the (100) and (110) surfaces.

The crystal structure of Ca3Pt4Sn13 is shown in Fig. 1(a). This structure features a cage-like network, in which one 20-coordinate cage surrounds the Ca atom and another surrounds the Pt atom, while a third cage is centered on an empty position. This configuration provides multiple available positions for the Sn atoms. Fig. 1(b) shows the studied crystal with the large (110) surface. Fig. 1(c) shows the powder XRD pattern of Ca3Pt4Sn13, obtained from crushed single crystals. All Bragg reflections could be appropriately fitted based on the Pr3Rh4Sn13-type cubic structure (space group: Pm[3 with combining macron]n, No. 223). Fig. 1(d) shows the XRD pattern of the large (110) surface.

Fig. 2(a) presents the temperature dependence of resistivity, ρ(T) for a Ca3Pt4Sn13 single crystal at zero magnetic field, measured between 2 K and 320 K with the electrical current I along the (110) surface. It can be seen that the value of ρ(T) increases monotonically with increasing temperature, indicating metallic behavior. The residual resistance ratio (RRR), defined as ρ(300 K)(2 K), was approximately 21, indicating the high quality of the crystal studied. As shown in the inset of Fig. 2(a), at low temperatures, the ρ(T) followed the equation:

 
ρ(T) = ρ0 + AT2(1)
where ρ0 and A represent the residual resistivity and the coefficient of the T2-term, respectively. Overall, the fitting line matched well with Fermi liquid theory at high temperatures, while it showed a significant deviation from the resistivity data in the low-temperature region, as shown in the inset of Fig. 2(a). To obtain more reliable results, we considered some additional evidence, such as the Wilson coefficient and the low-temperature magnetic susceptibility behavior, to comprehensively determine whether Ca3Pt4Sn13 conformed to a Fermi liquid system. The Wilson ratio, RW, is an effective parameter for evaluating Fermi-liquid behavior. Usually, the values of the RW ratio for correlated metals with strong spin–orbit coupling lie between 1 and 6.43 This notably high RW value is an outcome of the strong non-Fermi liquid nature of the compound. For example, high RW values have been observed in non-Fermi liquid compounds such as SrIrO3 (RW ∼ 55)44 and Pr2Ir2O7 (RW ∼ 21.7).45 In our case, the Wilson ratio of Ca3Pt4Sn13 was evaluated as RW = π2kB2χ0/3μB2γ ∼ 9.1, which was larger than that of Ca3Ir4Sn13 crystal (RW ∼ 1.7),23 indicating that Ca3Pt4Sn13 was a possible non-Fermi liquid system. In addition to the Wilson coefficient, the behavior of low-temperature magnetic susceptibility also serves as evidence for determining whether the system is Fermi or non-Fermi liquid. Usually, the magnetic susceptibility, χ, is independent of temperature T in a Fermi liquid system, while in a non-Fermi liquid system, it exhibits characteristics such as χ ∼ −In[thin space (1/6-em)]T or Tα (α < 1).46 For example, the divergence of magnetic susceptibility below 10 K follows a power-law form of χTα, with α ∼ 0.21 for Pb2Ir2O7−δ45 and α ∼ 0.28 for CeRhBi.46 Both of these are typical characteristics of a non-Fermi liquid. In our case, the low-temperature magnetic susceptibility fitted well with a power-law form of χTα, as shown in Fig. S1 (ESI).47 The fitting coefficient α ∼ 0.5 indicated a non-Fermi liquid system for Ca3Pt4Sn13. Therefore, by combining the results from resistivity, the Wilson ratio, and low-temperature magnetic susceptibility, we considered that Ca3Pt4Sn13 crystal behaved as a non-Fermi liquid system.


image file: d4tc04812a-f2.tif
Fig. 2 (a) Temperature dependence of resistivity at H = 0 T and 5 T with the electrical current along the (110) surface; inset: fitting result of the low-temperature resistance based on Fermi-liquid theory, ρ = ρ0 + AT2; (b) temperature-dependent MR under an applied magnetic field of H = 5 T; inset: temperature dependence of d(MR)/dT. T* corresponds to the temperature where the sign of d(MR)/dT changes from positive to negative as temperature decreases; (c) magnetic field (H)-dependent ρxx at different temperatures; (d) temperature dependence of MR defined as MR = (ρH(9 T) − ρH(0 T))/ρH(0 T). The lines represent fitting results according to MR ∼ H2 and MR ∼ H.

As shown in Fig. 2(b), with an applied magnetic field H = 5 T, at low temperatures, d(MR)/dT was positive. However, as the temperature increased, the sign of d(MR)/dT changed from positive to negative around T* ∼ 33 K for I along the [110] direction. The origin of MR in Ca3Pt4Sn13 differed from that in Gd3It4Sn13. In Gd3It4Sn13, the MR was related to the magnetic ordering of Gd ions under the applied magnetic field; however, no magnetic order was observed in Ca3Pt4Sn13. The origin of MR in Ca3Pt4Sn13 is still not clear. Furthermore, the angle-resolved magnetoresistance (AMR) of Ca3Pt4Sn13 was investigated, with the relative spatial position relations shown schematically in Fig. S3(a) and S4(a) (ESI).47 When the magnetic field was always perpendicular to the excitation current flow, as shown in Fig. S3 (ESI), the AMR data exhibited an isotropic symmetry at high temperatures. As the temperature decreased, the curves gradually evolved into an eight-fold symmetry. These features were similar to the multivalley system observed in Ag3Sn,48 which strongly depended on the orientation of the magnetic field, thus indicating that Ca3Pt4Sn13 could be a possible candidate for valleytronics. In contrast, the AMR maintained an isotropic symmetry over the entire temperature range when the magnetic field rotated within the bc plane, as shown in Fig. S4 (ESI).47 Some previous studies have proposed magnetic or Lifshitz transition scenarios to explain the changing symmetry.48 However, the microscopic origin of these changes remains an open issue.

As shown in Fig. 3, we systematically investigated the temperature-dependent thermal conductivity, thermoelectric power, resistivity, and ZT for the Ca3Pt4Sn13 single crystal under 0 T and 5 T. Characteristic transitions were clearly observed in the thermal conductivity, thermoelectric power, and ZT curves. However, interestingly, in addition to the large MR at low temperatures, no abnormal behavior was observed in the temperature-dependent resistivity data. To our knowledge, the possible reason is that resistance measurements typically select the optimal path in the system, which may not fully reflect the overall behavior of the sample. Correspondingly, measurements of thermal conductivity and thermoelectric power could be considered thermodynamic measurements, providing more bulk information about the selected crystal. Therefore, it is reliable to conclude that these characteristic transitions were determined from the thermoelectric measurements.


image file: d4tc04812a-f3.tif
Fig. 3 (a)–(c) Temperature dependence of the thermal conductivity (κ), Seebeck coefficient (S), and resistivity (ρ) at H = 0 T and 5 T, respectively; (d) temperature-dependent calculated ZT under applied magnetic fields of H = 0 T and 5 T.

To obtain more detailed information, we carefully extracted these critical temperature points. Two significant transitions were observed in the thermal conductivity (Tdrag = 24 K, Tpeak = 140 K) and thermoelectric power (T* = 35 K, Tpeak = 140 K) curves, while only one transition (T = 160 K) was observed in the ZT curve. After careful comparison, we found that although the thermal conductivity and thermoelectric power exhibited completely different evolutionary behaviors, their critical temperature points were almost identical. The maximum value of the merit factor ZT was 0.015 when the temperature was around 160 K. Combining the above results and analysis, we confirm the critical temperature points for these observed transitions, which will provide useful insights into understanding the interesting physical phenomena in the Ca3Pt4Sn13 single crystal.

Fig. 4 shows the specific heat, Cp, of the Ca3Pt4Sn13 single crystal. At higher temperatures, the value of Cp approached the limiting value of 500 J mol−1 K−1, which was close to the Dulong–Petit limit, Cv = 3nR0 = 498.8 J mol−1 K−1, where n = 20 is the total number of atoms per formula unit and R0 is the molar gas constant. Meanwhile, the magnitude of Cp decreased monotonically with decreasing temperature and showed no visible signatures of a structural transition. The inset of Fig. 4 shows the low-temperature specific heat data, Cp/T, plotted as a function of T2. It can be fitted by the following formula:

 
Cp/T = γ + βT2(2)
where γ is the Sommerfeld coefficient and βT2 represents the lattice contribution. The fitted values of γ and β were 4.76 mJ mol−1 K−12 and 3.16 mJ mol−1 K−14, respectively. According to the Sommerfeld theory of metals, the coefficient γ can be expressed as a function of the density of states (DOS) at the Fermi energy level as follows:
 
image file: d4tc04812a-t1.tif(3)
where N(EF) is the DOS at the Fermi energy level and λ is the electron–phonon coupling constant. Assuming λ = 0, the obtained value of N(EF) was 2.9 states per eV per f.u. The Debye temperature ΘD was obtained from the following formula:
 
image file: d4tc04812a-t2.tif(4)
For Ca3Pt4Sn13, n was equal to 20 and β was 3.16 mJ mol−1 K−1; therefore, the calculated ΘD was approximately 220 K.


image file: d4tc04812a-f4.tif
Fig. 4 Temperature dependence of the heat capacity Cp. The red line is the fitting result according to eqn (5). The inset shows the T2-dependence of Cp/T. The blue line is the fitting result based on the formula Cp/T = γ + βT2, where γ is the Sommerfeld coefficient and β is the phonon heat capacity coefficient.

To understand the behavior of the specific heat, Cp, we performed a detailed analysis of the conduction electron and phonon contributions using:

 
CpT = γT + nCDebyeV(T)(5)
and
 
image file: d4tc04812a-t3.tif(6)
where R is the molar gas constant, ΘD is the Debye temperature, and n = 20 is the number of atoms per formula unit. Using the obtained values of γ = 4.76 mJ mol−1 K−1 and ΘD = 220 K, the red line represented the calculated result, which matched the experimental data very well and confirmed that the obtained values of γ and ΘD were reasonable.

As shown in Fig. 5(b), we systematically performed the temperature-dependent Hall coefficient measurements at various fixed temperature points. The negative slope observed in the Hall resistance indicated that electron-type carriers were dominant in the Ca3PtSn13 crystal. Interestingly, the thermoelectric power results suggested the presence of hole-like carriers. The difference in the signs of the Hall and Seebeck coefficients has been observed in pure thallium,49 bismuth thin film,50 Ag2Se crystal,51 and others. Some theoretical studies suggest that the inconsistency in the signs of Hall and Seebeck coefficients may be related to the special structure of the Fermi surface boundaries. Because the boundary lines for the sign change of the Hall and Seebeck coefficients did not coincide, it was useful to define a region between them in which the Hall and Seebeck coefficients had different signs. This anomalous region always corresponded to a positive Seebeck coefficient and a negative Hall mobility, as found experimentally.52 In our case, to understand their intrinsic origin, we performed first-principles calculations for the Ca3Pt4Sn13 crystal, as shown in Fig. S2 (ESI).47 We found a “neck-shaped” behavior in the three-dimensional Fermi surface, as shown in Fig. S2(d) (ESI).47 However, at this stage, whether these “neck-shaped” Fermi surfaces are the microscopic origin of the different signs of the Hall and Seebeck effects remains an open issue. Further in-depth theoretical analysis in the future will clarify this issue.


image file: d4tc04812a-f5.tif
Fig. 5 (a) Temperature dependence of the Hall coefficient under an applied magnetic field of H = 5 T; (b) field dependence of the Hall resistivity ρxy at representative temperatures; (c) and (d) temperature dependence of the carrier density and mobility, respectively, obtained from the fitting of data in (b).

In order to obtain more information, we also extracted the temperature-dependent mobility of the carriers, the carrier concentration, and the hall coefficient RH, as shown in Fig. 5(a), (c) and (d). All of these extracted data exhibited a rapid increase below a characteristic temperature of ∼35 K, which was almost identical to the low-temperature transition of the Seebeck coefficient. Meanwhile, we found that the MR and AMR also showed significant changes when the temperature was below ∼35 K. On the other hand, the thermal conductivity exhibited a relatively poorer match with the above critical temperatures, showing a phase transition near 24 K. The magnetic susceptibility also showed a kink around 40 K in the zero-field cooling curve, as shown in Fig. S1(b) (ESI).47 Although the starting temperature points may vary due to different measurement methods, their evolutionary behaviors imply that they may share a common origin closely related to the T* phase. Some previous studies have suggested that the T* in the Ca3Ir4Sn13 superconductor may originate from CDW order or ferromagnetic spin fluctuations,23 which remains a controversial issue. Based on the current results, we seem to have also encountered a similar situation in the Ca3Pt4Sn13 system. Nevertheless, the current findings could provide a non-superconducting system to investigate the intrinsic origin of T*.

Unlike some Dirac semimetals, the Fermi surface is solely contributed by Dirac cones. We have not found any special electronic structure (near the Fermi surface) or significant compensation mechanisms to support physical mechanisms that could explain the large MR effect, such as the Dirac system, Weyl semimetal system, electric-hole compensation, and so on. In such cases, more fundamental and intrinsic factors of origin will need to be considered. Some studies have pointed out that high carrier mobility, μ, is an essential prerequisite for the observation of large MR.53 In our case, the mobility was almost consistent with the evolutionary behavior of the large MR. Therefore, at the current stage of research, we tend to believe that the high mobility behavior in the low-temperature region is an important factor contributing to the large MR phenomenon in Ca3Pt4Sn13. Certainly, more detailed microscopic mechanisms related to the T* scenarios (associated with the CDW order) need to be verified by further experimental and theoretical studies in the future.

In summary, we have carried out a detailed study of the quasi-skutterudite Ca3Pt4Sn13 single crystal through measurements of electrical resistivity, Hall effect, specific heat, and thermoelectric properties. The resistivity exhibited metallic behavior from 300 K to 2 K. The results of the resistivity, Wilson ratio, and low-temperature magnetic susceptibility indicated that Ca3Pt4Sn13 was a non-Fermi liquid system. The MR with an applied magnetic field H = 5 T at T = 2 K was 200% for H along the [110] direction. Combining the results of first-principles calculations and the Hall resistance, we tend to consider that the origin of the large MR is closely related to the high mobility of the carriers. Meanwhile, the AMR was also systematically investigated. An eight-fold symmetry AMR observed in the low-temperature region indicated that the Ca3Pt4Sn13 crystal could be a possible candidate for valleytronics. The results of the first-principles calculations and the different signs of the Hall and Seebeck coefficients indicated that there was a special structure of the Fermi surface boundaries in the Ca3Pt4Sn13 crystal. Furthermore, we found that the mobility of carriers, carrier concentration, Hall coefficient, Seebeck coefficient, thermal conductivity, magnetic susceptibility, MR, and AMR all showed significant changes when the temperature was below T*, indicating that they might share a common origin. The current findings could provide a non-superconducting Ca3Pt4Sn13 system to study the intrinsic origin of T*.

This work is supported by the National Key Research and Development Program of China (Grant No. 2023YFA1406500), the National Natural Science Foundation of China (No. 11904003, No. 12204487, 12474098, No. 12104011, No. 12274388, No. 12174361, No. 52102333, No. 12104010, and No. 12204004), and the Natural Science Foundation of Anhui Province under contract 1908085MA09.

Data availability

The data supporting the findings of this study can be obtained from the corresponding author upon request.

Conflicts of interest

There are no conflicts to declare.

References

  1. C. Shekhar, A. K. Nayak, Y. Sun, M. Schmidt, M. Nicklas, I. Leermakers, U. Zeitler, Y. Skourski, J. Wosnitza and Z. Liu, et al. , Nat. Phys., 2015, 11, 645 Search PubMed.
  2. J. Hu, J. Liu, D. Graf, S. Radmanesh, D. Adams, A. Chuang, Y. Wang, I. Chiorescu, J. Wei and L. Spinu, et al. , Sci. Rep., 2016, 6, 18674 Search PubMed.
  3. T. Liang, Q. Gibson, M. N. Ali, M. Liu, R. Cava and N. Ong, Nat. Mater., 2015, 14, 280 CrossRef CAS PubMed.
  4. N. Zhou, Y. Sun, C. Q. Xu, C. Y. Xi, Z. S. Wang, B. Li and J. J. Feng, et al. , Phys. Rev. B, 2020, 101, 245102 CrossRef CAS.
  5. H. Asano, J. Hayakawa and M. Matsui, et al. , Appl. Phys. Lett., 1997, 71, 844 CrossRef CAS.
  6. Y. Tomioka, H. Kuwahara, A. Asamitsu and M. Kasai, Appl. Phys. Lett., 1997, 70, 3609 CrossRef CAS.
  7. H. Y. Hwang, S. W. Cheong and B. Batlogg, Appl. Phys. Lett., 1996, 68, 3494 CrossRef CAS.
  8. L. X. Zhao, J. B. He, D. Chen, S. Zhang, Z. A. Ren and G. F. Chen, Phys. Rev. B, 2019, 99, 205116 CrossRef CAS.
  9. R. Singha, A. Pariari, G. K. Gupta, T. Das and P. Mandal, Phys. Rev. B, 2018, 97, 155120 CrossRef CAS.
  10. Z. J. Yuan, H. Lu, Y. J. Liu, J. F. Wang and S. Jia, Phys. Rev. B, 2016, 93, 184405 CrossRef.
  11. D. S. Wu, J. Liao, W. Yi, X. Wang, P. G. Li, H. M. Weng, Y. G. Shi, Y. Q. Li, J. L. Luo, X. Dai and Z. Fang, Appl. Phys. Lett., 2016, 108, 042105 CrossRef.
  12. C. L. Zhang, C. Guo, H. Lu, X. Zhang, Z. J. Yuan, Z. Q. Lin, J. F. Wang and S. Jia, Phys. Rev. B, 2015, 92, 041203 CrossRef.
  13. Y. Y. Lv, X. Li, J. L. Zhang, B. Pang, S. S. Chen, L. Cao, B. B. Zhang, D. J. Lin, Y. B. Chen, S. H. Yao, J. Zhou, S. T. Zhang, M. H. Lu, M. L. Tian and Y. F. Chen, Phys. Rev. B, 2018, 97, 245151 CrossRef CAS.
  14. S. S. Sun, Q. Wang, P. J. Guo, K. Liu and H. C. Lei, New J. Phys., 2016, 18, 082002 CrossRef.
  15. Y. L. Wang, L. R. Thoutam, Z. L. Xiao, J. Hu, S. Das, Z. Q. Mao, J. Wei, R. Divan, A. Luican-Mayer, G. W. Crabtree and W. K. Kwok, Phys. Rev. B, 2015, 92, 180402 CrossRef.
  16. Q. Chen, Z. F. Lou, S. N. Zhang, Y. X. Zhou, B. J. Xu, H. C. Chen, S. J. Chen, J. H. Du, H. D. Wang, J. H. Yang, Q. S. Wu, O. V. Yazyev and M. H. Fang, Phys. Rev. B, 2021, 104, 115104 CrossRef CAS.
  17. J. Diaz, K. Wang, J. Straquadine, C. Putzke, Q. Yang, B. Yan, S. L. Budko, P. C. Canfield and P. J. W. Moll, Nature Commun, 2024, 15, 4585 CrossRef CAS PubMed.
  18. D. S. Wu, X. Wang, X. Zhang, C. L. Yang, P. Zheng, P. G. Li and Y. G. Shi, Sci. China: Phys., Mech. Astron., 2015, 58, 1 CrossRef CAS.
  19. I. Di Bernardo, J. Collins, W. Wu, J. Zhou, S. A. Yang, S. Ju, M. T. Edmonds and M. S. Fuhrer, Phys. Rev. B, 2020, 102, 045124 CrossRef CAS.
  20. Z. K. Liu, B. Zhou, Y. Zhang, Z. J. Wang, H. M. Weng, D. Prabhakaran, S.-K. Mo, Z. X. Shen, Z. Fang, X. Dai, Z. Hussain and Y. L. Chen, Science, 2014, 343, 864 CrossRef CAS PubMed.
  21. H. X. Chen, Z. L. Li, L. W. Guo and X. L. Chen, EPL, 2017, 117, 27009 CrossRef.
  22. J. Y. Hua, Z. X. Dai, X. C. Kana, G. H. Zheng, Z. Chena and Y. Q. Ma, J. Alloys Compd, 2022, 895, 162563 CrossRef.
  23. J. H. Yang, B. Chen, C. Michioka and K. Yoshimura, J. Phys. Soc. Jpn., 2010, 79, 113705 CrossRef.
  24. K. F. Wang and C. Petrovic, Phys. Rev. B, 2012, 86, 024522 CrossRef.
  25. S. Y. Zhou, H. Zhang, X. C. Hong, B. Y. Pan, X. Qiu, W. N. Dong, X. L. Li and S. Y. Li, Phys. Rev. B, 2012, 86, 064504 CrossRef.
  26. A. F. Fang, X. B. Wang, P. Zheng and N. L. Wang, Phys. Rev. B, 2014, 90, 035115 CrossRef CAS.
  27. C. N. Kuo, H. F. Liu, C. S. Lue, L. M. Wang, C. C. Chen and Y. K. Kuo, Phys. Rev. B, 2014, 89, 094520 CrossRef.
  28. C. W. Luo, P. C. Cheng, C. M. Tu, C. N. Kuo, C. M. Wang and C. S. Lue, New J. Phys., 2016, 18, 073045 CrossRef.
  29. A. M. Strydom, J. Phys.: Condens. Matter, 2007, 19, 386205 CrossRef.
  30. C. L. Yang, X. Wang, X. Zhang, D. S. Wu, M. Liu, P. Zheng, J. Y. Yao, Z. Z. Li, Y.-F. Yang, Y. G. Shi, J. L. Luo and N. L. Wang, Phys. Rev. B, 2015, 91, 075120 CrossRef.
  31. D. A. Tompsett, Phys. Rev. B, 2014, 89, 075117 CrossRef.
  32. L. S. I. Veiga, J. R. L. Mardegan, M. V. Zimmermann, D. T. Maimone, F. B. Carneiro, M. B. Fontes, J. Strempfer, E. Granado, P. G. Pagliuso and E. M. Bittar, Phys. Rev. B, 2020, 101, 104511 CrossRef CAS.
  33. J. Luo, J. Yang, S. Maeda, Z. Li and G. Q. Zheng, Chin. Phys. B, 2018, 27, 7 Search PubMed.
  34. M. Zeng, D. H. Xu, Z. M. Wang and L. H. Hu, Phys. Rev. B, 2023, 107, 094507 CrossRef CAS.
  35. C. Wang, B. Lian, X. M. Guo, J. H. Mao, Z. T. Zhang, D. Zhang, B. L. Gu, Y. Xu and W. H. Duan, Phys. Rev. Lett., 2019, 123, 126402 CrossRef CAS PubMed.
  36. K. Zollner and J. Fabian, Phys. Status Solidi B, 2021, 258, 2000081 CrossRef CAS.
  37. N. Zhou, X. F. Xu, J. R. Wang, J. H. Yang, Y. K. Li, Y. Guo, W. Z. Yang, C. Q. Niu, B. Chen, C. Cao and J. H. Dai, Phys. Rev. B, 2014, 90, 094520 CrossRef CAS.
  38. Q. Youa, X. C. Wang, G. F. Chen and Z. D. Zheng, Phys. E, 2017, 88, 87 CrossRef.
  39. J. Jiang, F. Tang, X. C. Pan, H. M. Liu, X. H. Niu, Y. X. Wang, D. F. Xu, H. F. Yang, B. P. Xie, F. Q. Song, P. Dudin, T. K. Kim, M. Hoesch, P. Kumar Das, I. Vobornik, X. G. Wan and D. L. Feng, Phys. Rev. Lett., 2015, 115, 166601 CrossRef CAS PubMed.
  40. J. Taylor, H. Guo and J. Wang, Phys. Rev. B, 2001, 63, 303 Search PubMed.
  41. G. Kresse and J. Furthmüller, Phys. Rev. B, 1996, 54, 11169 CrossRef CAS PubMed.
  42. J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865 CrossRef CAS PubMed.
  43. L. E. DeLong, R. P. Guertin, S. Hasanain and T. Fariss, Phys. Rev. B, 1985, 31, 7059 CrossRef CAS PubMed.
  44. G. Cao, V. Durairaj, S. Chikara, L. E. DeLong, S. Parkin and P. Schlottmann, Phys. Rev. B, 2007, 76, 100402 CrossRef.
  45. M. S. Khan, I. Carlomagno, C. Meneghini, P. K. Biswas, F. Bert, S. Majumdar and S. Ray, Phys. Rev. B, 2022, 105, 085137 CrossRef CAS.
  46. V. K. Anand, D. T. Adroja, A. D. Hillier, K. Shigetoh, T. Takabatake, J. G. Park, K. A. McEwen, J. H. Pixley and Q. Si, J. Phys. Soc. Jpn., 2018, 87, 064708 CrossRef.
  47. See ESI for the temperature dependent magnetic susceptibility, the first-principles calculations, the polar plot of field angle-resolved resistivity in the ac plane, the polar plot of field angle-resolved resistivity in the bc plane.
  48. N. Zhou, Y. Sun, C. Q. Xu, Z. S. Wang, B. Li, J. J. Feng, L. Zhang, X. Z. Xing, Y. F. Zhang, Y. Q. Pan, Y. Meng, X. L. Yi, L. Pi, X. F. Xu and Z. X. Shi, Phys. Rev. B, 2020, 101, 245102 CrossRef CAS.
  49. A. T. Burkov and P. P. Konstantinov, Tech. Phys., 2024, 69, 1510 CrossRef CAS.
  50. V. A. Komarov, V. M. Grabov, A. V. Suslov, N. S. Kablukova and M. V. Suslov, Semiconductors, 2019, 53, 593 CrossRef CAS.
  51. M. Ferhat and J. Nagao, J. Appl. Phys., 2000, 88, 813 CrossRef CAS.
  52. M. R. Boon, Phys. Status Solidi B, 1972, 49, K163 CrossRef CAS.
  53. N. Xin, J. Lourembam, P. Kumaravadivel, A. E. Kazantsev, Z. Wu, C. Mullan, J. Barrier, A. A. Geim, I. V. Grigorieva, A. Mishchenko, A. Principle, V. I. Falko, L. A. Ponomarenko, A. K. Geim and A. I. Berdyugin, Nature, 2013, 616, 270 CrossRef PubMed.

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4tc04812a

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.