Next Article in Journal
Improvement of ZrC/Zr Coating on the Interface Combination and Physical Properties of Diamond-Copper Composites Fabricated by Spark Plasma Sintering
Previous Article in Journal
An Overview of Parameters Controlling the Decomposition and Degradation of Ti-Based Mn+1AXn Phases
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Electrical and Thermal Transport Properties of Layered Superconducting Ca10(Pt4As8)((Fe0.86Pt0.14)2As2)5 Single Crystal

1
College of Science, Liaoning Shihua University, Fushun 113001, China
2
College of Computer Science and Engineering, Northeastern University, Shenyang 110004, China
3
Institute of Materials Physics and Chemistry, School of Materials Science and Engineering, Northeastern University, Shenyang 110004, China
*
Author to whom correspondence should be addressed.
Materials 2019, 12(3), 474; https://doi.org/10.3390/ma12030474
Submission received: 20 December 2018 / Revised: 31 January 2019 / Accepted: 31 January 2019 / Published: 4 February 2019

Abstract

:
We have synthesized single crystals of iron-based superconducting Ca10(Pt4As8)((Fe0.86Pt0.14)2As2)5 and performed extensive measurements on their transport properties. A remarkable difference in the behavior and a large anisotropy between in-plane and out-of-plane resistivity was observed. Disorder could explain the in-plane square-root temperature dependence resistivity, and interlayer incoherent scattering may contribute to the out-of-plane transport property. Along the ab plane, the estimated value of the coherence length is 15.5 Å. From measurements of the upper critical magnetic field Hc2 (T ≥ 20 K), we estimate Hc2(0) = 313 T. Thermal conductivity for Ca10(Pt4As8)((Fe0.86Pt0.14)2As2)5 is relatively small, which can be accounted for by the disorder in the crystal and the low-charge carrier density as verified by the Hall effect.

1. Introduction

The Ca10(PtnAs8)(Fe2As2)5 (n = 3 or 4) system was first reported to exhibit superconductivity with a wide range of transition temperatures in 2011 [1,2,3]. After that, many studies have appeared in the past seven years [4,5,6,7,8]. Ca10(Pt4As8)(Fe2As2)5 has a layered structure consisting of superconducting FeAs layers separated by the spacer layers arranged as Ca–Pt4As8–Ca, which is shown in Figure 1a. This superconductor has been reported to crystallize in possible space groups including P4/n (tetragonal) [2,3], P21/n (monoclinic) [7], and P 1 ¯ (triclinic) [1,3]. The structure of the Pt4As8 spacer layer is similar to a square lattice of As atoms, one fifth of the As atoms are replaced by substitutional Pt1 atoms [2], and the same amount of Pt2 atoms are interstitial, which leads to the displacement of As atoms from their ideal positions to form As dimers [2]. Because of the constraint from the As–As dimers and the FeAs sublattices, substitutional Pt1 atoms can sit in plane while the interstitial Pt2 atoms sit on the site either above or below the plane against Ca ions.
Most FeAs-based superconductors need doping to induce the superconductivity; Ca10(Pt4As8)(Fe2As2)5 allows doping on the Ca site or on the Fe site. In particular, Pt doping on the Fe site could bring faults and disorder into the crystal and also adjust the Tc [2,5]. By studying its structure and physical properties, the origin of superconductivity may be elucidated. The interlayer distance of Ca10(Pt4As8)(Fe2As2)5 is reported to be ~10.5 Å [2,3], which is larger than many other FeAs-based superconductors, and thus its properties are expected to be anisotropic.
The electronic structure for Ca10(Pt4As8)(Fe2As2)5 has been studied via angle resolved photoemission spectroscopy (ARPES) [9]. Generally, there is a hole pocket and an electron pocket around the zone center. There are several electron pockets around the zone corner, among which two are suggested to be contributed by the FeAs layer, and several are contributed by Pt4As8 layers. Thus, the Pt4As8 layer is suggested to be metallic. Furthermore, the dxz and dyz bands are not degenerate at the Brillouin zone center (Γ point) and there is only hole-like Fermi surface at the Γ point originated from dxy orbitals, it may be caused by the interaction between the Pt4As8 and FeAs layers [9].
In this paper, we studied the electrical properties of Ca10(Pt4As8)(Fe2As2)5 along its c-axis in order to understand how the interaction between Pt4As8 and FeAs layers influences the physical properties and anisotropy in this system. In addition, thermal conductivity has not yet been reported as another approach in measuring transport properties, and thus we also provide more extensive measurements and discussion on this approach.

2. Material Preparation and Characterization Methods

The preparation process is similar to descriptions provided elsewhere [2,4]. To grow single crystals of Ca10(Pt4As8)(Fe2−xPtxAs2)5, stoichiometric amounts of high purity Ca shot (99.999%, Alfa Aesar, Haverhill, MA, USA), Fe powder (99.95%, Alfa Aesar, Haverhill, MA, USA), Pt powder (99.95%, Alfa Aesar, Haverhill, MA, USA), and As powder (99.999%, Alfa Aesar, Haverhill, MA, USA) were mixed in the ratio of 17:14:9:31. The mixture was placed in an Al2O3 crucible (Gongtao Ceramics, Shanghai, China) and sealed in a quartz tube (Gongtao Ceramics, Shanghai, China) under vacuum. The whole ampule was heated to 700 °C in a box furnace (Henan Sante, Luoyang, China) at a rate of 140 °C/h and kept at this temperature for 5 h. It was further heated to 1100 °C at a rate of 80 °C/h where it was held for 50 h and then cooled to 1050 °C at a rate of 1.25 °C/h. It was further cooled to 500 °C in the next 100 h and finally cooled down to room temperature naturally by switching the power off. The cooling process is crucial because the mixture melts when it heats up and forms in a stable phase when it cools down. Shiny plate-like single crystals were obtained with a typical size of 3 mm × 3 mm × 0.8 mm, and a photo of the as prepared sample is shown in Figure 1b.
Crystal structure and phase purity were checked by both single-crystal and powder X-ray diffraction, which were carried out on a Rigaku-D/max X-ray diffractometer (XRD) (Rigaku, Tokyo, Japan) using Cu-Kα radiation (λ = 1.54056 Å). Powder X-ray diffraction employed the powder grounded from the as-grown single crystals. The crystal used for this experiment was selected from the same batch as that used for composition and physical properties measurements. The chemical composition of the sample was identified by Energy Dispersive X-Ray Spectroscopy (EDX) (Hitachi, Tokyo, Japan). Electrical resistivity, magnetoresistance, Hall effect, thermal conductivity, and Seebeck coefficient measurements were performed in a Quantum Design PPMS System. Both ab plane and c-axis resistivities were measured using the standard four-probe technique.

3. Results and Discussion

Figure 1c shows the X-ray diffraction spectrum for the as-grown thin, plate-like single crystal along the c-axis. Note that only (00l) peaks are observed and the pattern matches very well with that previously reported in [4]. A powder XRD spectrum is shown in Figure 1d. The result exhibits the peaks from diffractions not only (00l) planes but also (hkl) planes with nonzero h, k, l. X-ray structure determination confirms that the lattice parameters and angles for the sample are a = 8.7548(12) Å, b = 8.7642(10) Å, c = 10.69005(8) Å, and α = 94.674(9)°, β = 104.396(8)°, γ = 90.037(10)°, respectively, indicating our sample crystallizes in a triclinic structure with space group symmetry P 1 ¯ at room temperature. No other impurities such as FeAs and PtAs2 were observed in any X-ray spectrum.
The chemical composition was identified by the EDX measurement. A scanning electron microscope (SEM) image measured on the surface of the single crystal is shown in the Figure 2 inset. For each sample, five locations scattered on the surface were chosen for EDX measurement. An average of these scans was calculated. The results of the different samples are consistent with each other, implying homogenous growth was acquired throughout the batch. The average deviation is 3% for Ca, 5% for Pt, and 1% for Fe. The measured composition of the single crystals is Ca10Pt5.4Fe8.6As18, corresponding to a formula of Ca10(Pt4As8)((Fe0.86Pt0.14)2As2)5 if considering the Pt substitution effect.
Figure 3 displays the temperature dependence of the in-plane (ρab) and out-of-plane resistivity (ρc) of Ca10(Pt4As8)((Fe0.86Pt0.14)2As2)5 single crystals between 2 K and 300 K. The magnitude of ρab and its overall features are similar to the previous reports [2]. The normal-state ρab decreases with decreasing T ( d ρ ab d T > 0 ), showing a metallic behavior. It drops sharply at Tc_onset = ~34 K, reaching zero-resistivity state at 31.2 K. The transition width in temperature is less than 3 K, indicating that our sample has high quality and spatial homogeneity. The residual resistivity ratio (RRR) (300K)/ρ(Tc_onset) is 2.4. Relatively large Tc_onset and small RRR reflects the presence of Pt doping on the Fe site in FeAs layers as suggested in [1,2,3]. However, the out-of-plane resistivity ρc behaves strikingly different with ρab; it exhibits nearly independent T at higher temperatures and starts to increase with decreasing T ( d ρ c d T < 0 ) below 200K, showing a nonmetallic behavior. It then goes through a sharp peak around 38 K before dropping to zero. The anisotropic property between ρab and ρc has not been noted in previous reports [1,2,4]. We acknowledge that this ρcT profile is reproducible and it should be an intrinsic property of Ca10(Pt4As8)(Fe2−xPtxAs2)5.
Due to the layered crystal structure of Ca10(Pt4As8)((Fe0.86Pt0.14)2As2)5, the in-plane resistivity may be understood as a net resistivity of parallel connected resistors. A schematic of such a system is shown in Figure 4a. Here, we do not consider the resistance from the Ca layer because the FeAs and Pt4As8 layers are the main contributors of the electronic structure around the Fermi surface [9]. The net resistance (R) can be written as:
R = 1 / [ n ( 1 R F e A s + 1 R P t 4 A s 8 ) ]
in which n is the number of Pt4As8–FeAs layers, and RFeAs and RPt4As8 are the resistance of the FeAs layer and Pt4As8 layer, respectively. If the Pt4As8 layer is semiconducting or insulating, it should have a much larger resistance than the FeAs layer. Then, the net resistance R should be dominated by the resistance of the Pt4As8 layer and exhibit nonmetallic behavior. Yet, according to our results, the normal state ρab shows metallic behavior, indicating the spacer layer Pt4As8 is metallic, which agrees with previous reports [2,4,9].
The normal state ρab of some traditional FeAs-based superconductors shows metallic behavior and may have linear dependence on temperature due to the inelastic scattering originating from electron-phonon interactions. In some cuprates, ρab has T1.5 dependence, implying the Fermi-liquid behavior. Yet, the temperature dependence of the normal state ρab of Ca10(Pt4As8)((Fe0.86Pt0.14)2As2)5 is neither linear, quadratic temperature dependent, nor T1.5 dependent, as observed in cuprates or other FeAs-based compounds [10]. The normal state ρab from 50 K to 300 K can be well fitted by ρ a b = A + B T with A = 0.079 ± 0.008 mΩ cm and B = 0.031 ± 0.003 mΩ cm K−1/2. The red solid line in Figure 3a is the fit to the experimental data of the normal state ρab. All reported ρab for Ca10Pt4As8(Fe2As2)5 has the similar square-root temperature dependence regardless of Tc [1,2,4], and thus, it may be an intrinsic property to Ca10Pt4As8(Fe2As2)5.
The square-root temperature dependence of electrical resistivity is normally expected at low temperatures in disordered metals and degenerate semiconductors because of interference with scattering by impurities [11,12]. The studies on the Fermi surface of Ca10Pt4As8(Fe2As2)5 imply that the Pt4As8 layer contributes several electron pockets. Thus, there may be competition between the Pt4As8 layer and the negatively-charged FeAs layer for electrons, leading to possible Pt deficiency in the Pt4As8 layer and partial substitution of Fe by Pt in the Fe1−xPtxAs layer [3,5,13,14]. As a result, charge carriers in both the Pt4As8 and Fe1−xPtxAs layers could experience the effect of disorder, which may be the reason why the in-plane electrical resistivity has the square-root temperature dependence.
As shown in Figure 3b, ρc is much larger than ρab in the normal state. It can be understood with a schematic plot of the net resistance R = n (RFeAs + RPt4As8) along the out-of-plane direction shown in Figure 4b. Different from the normal state ρc of some cuprates which can be described using the equation ρc = A/T + B × T exhibiting the non-Fermi liquid behavior, the nonmetallic normal state ρc of Ca10(Pt4As8)((Fe0.86Pt0.14)2As2)5 can be appropriately fitted using the equation ρ c = A + B T + C T 1 . The solid line in Figure 3b shows the fit of ρc from 90 K to 300 K. The fitting parameters are calculated as follows: A′ = 2.642 ± 0.009 mΩ cm, B′ = 0.308 ± 0.006 mΩ cm K−1/2, and C′ = 667.7 ± 15 mΩ cm K.
The square-root temperature dependent component in ρc is due to the in-plane scattering, which may be intrinsic as Anderson proposed for high-Tc cuprates [15] or extrinsic due to a possible stacking fault [2]. The T−1 term can be attributed to interlayer incoherent scattering, similar to the properties of high-Tc cuprates [15] and (Sr4V2O6)Fe2As2 [16]. It implies the spacer layer Ca10Pt4As8 is not superconducting in nature, so Josephson junctions are formed along the c-axis of Ca10Pt4As8(Fe2As2)5 by stacking of the FeAs superconducting layer, the Ca10Pt4As8 semiconducting layer, and the FeAs superconducting layer.
The resistivity anisotropy γ is calculated by γ = ρ c / ρ a b and it increases from ~4.1 at 300 K to ~7.9 around Tc as displayed in Figure 5. The γ of Ca10Pt4As8(Fe2As2)5 is larger than that of LaFeAsO0.9F0.1 [17]. Relatively large anisotropy and a conspicuous difference in the behavior of ρc and ρab suggests rgw strong 2D nature of electronic structures in Ca10Pt4As8(Fe2As2)5, which is rarely encountered in traditional Fe-based superconductors.
The resistivity anisotropy (γ) in the superconducting state of Ca10(Pt4As8)((Fe0.86Pt0.14)2As2)5 was evaluated through resistivity measurements with an applied magnetic field H. Figure 6 shows the temperature dependence of ρab and ρc under H along both the c-axis and the ab plane.
In the case of I//ab and H//c-axis, it can be seen that the zero resistivity transition temperature (Tc0) decreases with the rising applied magnetic field accompanied by an increase in the transition width. The Tc0 is 18 K and the transition width Tc is 16 K for H = 14 T, considering Tc_onset is 34 K. This type of superconducting transition broadening with increasing magnetic field (H) is rarely seen in Fe-based superconductors but very common in cuprates due to the presence of strong thermal vortices fluctuations [18].
Interestingly, the superconducting transition width Tc does not change with an applied field when I//c and H//ab plane (Figure 6b) and H pushes both Tc_onset and Tc0 to the lower temperature by the same amount. This shift is also much smaller than that of Tc for cases I//ab and H//c.
As shown in Figure 7, the Hc2 (T) of Ca10(Pt4As8)((Fe0.86Pt0.14)2As2)5 was extracted by identifying the Tc(H) at which resistivity drops to 90% ρn, 50% ρn, and 10% ρn. ρn is the normal resistivity before superconducting transition, as indicated by dashed lines in Figure 6.
The temperature dependence of the Hc2 anisotropy parameter Γ H = H c 2 a b H c 2 c obtained from Figure 7 is presented in Figure 8. ΓH taken at 90% resistivity drop reaches 8 near Tc, which is very close to the normal-state γ we obtained previously in the resistivity measurements. ΓH varies from 4 to 6 using a 50% criteria as compared to 122 type. For example, with ΓH = ~2 for BaFe2As2 [19,20], the Hc2 anisotropy of Ca10(Pt4As8)((Fe0.86Pt0.14)2As2)5 is much larger.
It also can be noted from Figure 7b that H c 2 a b ( T ) increases almost linearly with decreasing temperature, while H c 2 c ( T ) exhibits an upward shape with a steep increase at low temperatures. This phenomena quite possibly originates from the multi-band effect such as in the NdFeAsO0.7F0.3 and MgB2 systems [21,22]. Thus, the positive curvature of H c 2 c ( T ) reflects a multi-band nature in electrical structure for Ca10(Pt4As8)(Fe2As2)5.
Upper critical field Hc2 is an important parameter for all superconductors especially in their practical applications. We estimated Hc2 at 0 K using the Werthamer–Helfand–Hohenberg (WHH) approximation [23]:
H c 2 ( 0 ) = 0.69 T c × | d H c 2 d T | T c
The corresponding coherence lengths are calculated via the Ginburg Landau (GL) formula:
{ ξ ab ( 0 ) = φ 0 2 π H c 2 c ( 0 ) ξ c ( 0 ) = φ 0 2 π ξ ab H c 2 a b ( 0 )
where φ 0 = 2.07 × 10 15 Wb. The obtained results are listed in Table 1. Although all derived values depend on how Hc2(T) is extracted, the large difference of H c 2 a b and H c 2 c implies Ca10(Pt4As8)((Fe0.86Pt0.14)2As2)5 exhibits a large anisotropy in its superconducting state. Furthermore, the values of ξc(0) derived from different criteria are all less than 10.69005(8) Å, i.e., the length of c-axis of the unit cell, suggesting that Ca10(Pt4As8)((Fe0.86Pt0.14)2As2)5 meets the dimensional requirement of fabricating the intrinsic Josephson junctions (iJJs).
For a normal metal with Fermi liquid behavior, the Hall coefficient (RH) is independent of temperature. The situation is more complex if the material has multi-band or non-Fermi liquid behavior such as, for example, cuprates or heavy fermions. Their Hall coefficient exhibits strong temperature and doping dependencies. The H c 2 c ( T ) curves show the multi-band nature of Ca10(Pt4As8)((Fe0.86Pt0.14)2As2)5, so the Hall effect was measured by applying a magnetic field HI//ab, and the data is presented in Figure 9.
The transverse RH remains negative at all temperatures above Tc, indicating that the charge carrier is dominated by electrons. The magnitude of RH increases with decreasing temperature, suggesting a multi-band and non-Fermi liquid behavior. Even though the doping in the FeAs layers is considered to be isovalent (i.e., Pt2+ replaces Fe2+) [2,5], it may still influence RH. Our RH(T) is similar with previous reports [1,2], but it exhibits stronger temperature dependence, which suggests that the carrier concentration cannot be solely determined by the Pt concentration in this material.
If a single band model is adopted, the corresponding carrier concentration (n) could be calculated with RH via n = −1/eRH. The temperature dependence of n is shown as the red line in Figure 9. Generally, n decreases monotonically with decreasing T. The calculated n is about 2.5 × 1022 cm−3 at 300 K. It is comparable to other FeAs-based superconductors; for example, n of SrFe2As2 is about 1.52 × 1022 cm−3 at 300 K [24]. The carrier concentration indicates that the normal state of Ca10(Pt4As8)((Fe0.86Pt0.14)2As2)5 behaves as a good metal. Before superconducting transition, the carrier concentration drops to 0.5 × 1022 cm−3; however, it is still an order of magnitude larger than that of Ca10-3-8 (~1021 between 2 K and 300 K) [2]. This is because the one extra Pt atom in the Pt4As8 intermediary layer exceeds the Zintl valence satisfaction requirement and introduces redundant electrons into the system, leading to enhanced metallicity and a relatively higher Tc.
A different method of probing conduction mechanism is provided by thermal conductivity. We measured the temperature dependence of thermal conductivity and the Seebeck coefficient for Ca10(Pt4As8)((Fe0.86Pt0.14)2As2)5 from 2 K to 350 K, and the results are shown in Figure 10a.
The Seebeck coefficient (S) is negative at the measured temperature range above Tc, with a value of −25.5604 μV/K at 300 K and a minimum value of −29.1343 μV/K near 154 K resulting from a phonon-drag contribution. This confirms that the electron-type charge carrier dominates in Ca10(Pt4As8)((Fe0.86Pt0.14)2As2)5, which is consistent with the Hall effect measurement. Then, S starts to increase with decreasing temperature and reaches zero steeply at the superconducting transition temperature. In addition, the Seebeck curve shows no anomalous enhancements associated with the crystal structure or spin density wave (SDW) transitions widely observed in undoped compounds such as SmFeAsO and BaFe2As2 [25,26], suggesting that there were no corresponding transitions in our sample at the temperatures measured.
With regard to the thermal conductivity (κ), it drops monotonously above Tc when temperature is lower than Tc, κ and decreases sharply with the opening of the superconducting gap. Before the drastic drop, κ first displays an abrupt increase with a bump feature upon entering the superconducting (SC) state, and a similar behavior is commonly observed in the cuprate superconductors, e.g., YBa2Cu3O7 [27] and Bi2Sr2CaCu2O8 [28]. The enhancement of κ below Tc reflects the increase of the phonon mean free path by the condensation of charge carriers. Phonons then cease to dissipate their momentum in collisions with such a condensate.
The electronic contribution to thermal conductivity ( κ e ) above Tc could be evaluated by the Wiedemann–Franz law κ e l = σ L T , where σ is the electrical conductivity of a metal and T is the temperature, L, known as the Lorenz number, which is equal to:
L = π 2 3 ( k B e ) 2 = 2.44 × 10 8   W Ω K 2
in which e is elementary charge and kB is the Boltzmann constant. The calculated results are presented in Figure 10b. κ e l is smaller by about 5 orders of magnitude than κ as shown in Figure 10a. Thus, heat in Ca10(Pt4As8)((Fe0.86Pt0.14)2As2)5 is mainly carried by phonons, and the electron contribution can be negligible. This is wholly different from the copper oxide superconductors. Ca10(Pt4As8)((Fe0.86Pt0.14)2As2)5 is a relatively low-charge carrier density system.
It is worth noting that the κ value for Ca10(Pt4As8)((Fe0.86Pt0.14)2As2)5 is smaller than that for LaFeAsO0.89F0.11 [29]. We know that the main scattering mechanisms for phonons in crystal are carriers and structural defects, and intrinsic phonon–phonon scattering only exists in clean materials. While in the Ca10(Pt4As8)(Fe2As2)5 system, both the off-centered Pt atoms in the Pt4As8 plane and the substitutional Pt atoms in the FeAs plane introduce disorder into the crystal, which causes phonons to be strongly scattered and the crystal lattice vibration to localize, resulting in the rather smaller thermal conductivity of Ca10(Pt4As8)((Fe0.86Pt0.14)2As2)5.

4. Conclusions

In conclusion, X-ray diffraction, resistivity, Hall effect, Seebeck coefficient, and thermal conductivity measurements were performed on high quality Ca10(Pt4As8)((Fe0.86Pt0.14)2As2)5 single crystals. We observed metallic in-plane resistivity but non-metallic out-of-plane resistivity for Ca10Pt4As8(Fe2As2)5. The anisotropic property is unusual, and its normal state resistivity exhibits a large anisotropy (~8) near Tc, making it one of the most anisotropic FeAs-based superconductors. The normal state in-plane resistivity has a square-root temperature dependence which is intrinsic to Ca10Pt4As8(Fe2As2)5. The interlayer incoherent scattering contributes to the out-of-plane transport property. A large coherence length along the ab plane and upper critical field were observed. Disorder and low-charge carrier density in the crystal may account for the relatively small thermal conductivity. The layered structure and the relatively higher transition temperature with the large electrical transport anisotropy of Ca10Pt4As8(Fe2As2)5 implies it may be a new good candidate for and have potential application in the fabrication of high frequency microelectronic devices such as next generation intrinsic Josephson junctions (IJJs).

Author Contributions

Y.Q. and D.W. conceived and designed the experiments; X.M. performed the experiments; Y.Z. and H.Y analyzed the data; D.W. and J.Y. wrote the paper.

Funding

This research was funded by the National Key Research Program during the 13st Five-Year Plan Period of China (Grant No. 2017YFB0304200), the Liaoning Shihua University Foundation, China (Grant No. 2018XJJ-001) and the Northeastern University Foundation, China (Grant No. N171603015).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kakiya, S.; Kudo, K.; Nishikubo, Y.; Oku, K.; Nishibori, E.; Sawa, H.; Yamamoto, T.; Nozaka, T.; Nohara, M. Superconductivity at 38 K in Iron-Based Compound with Platinum–Arsenide Layers Ca10(Pt4As8)(Fe2−xPtxAs2)5. J. Phys. Soc. Jpn. 2011, 80, 093704. [Google Scholar] [CrossRef]
  2. Ni, N.; Allred, J.M.; Chan, B.C.; Cava, R.J. High Tc electron doped Ca10(Pt3As8)(Fe2As2)5 and Ca10(Pt4As8)(Fe2As2)5 superconductors with skutterudite intermediary layers. Proc. Natl. Acad. Sci. USA 2011, 108, E1019–E1026. [Google Scholar] [CrossRef] [PubMed]
  3. Lohnert, C.; Sturzer, T.; Tegel, M.; Frankovsky, R.; Friederichs, G.; Johrendt, D. Superconductivity up to 35 K in the iron-platinum arsenides (CaFe1−xPtxAs)10 (Pt4−yAs8) with layered structures. Angew. Chem. Int. Ed. 2011, 50, 9195–9199. [Google Scholar] [CrossRef] [PubMed]
  4. Ding, Q.P.; Tsuchiya, Y.; Mohan, S.; Taen, T.; Nakajima, Y.; Tamegai, T. Magnetic and transport properties of iron-platinum arsenide Ca10(Pt4−δAs8)(Fe2−xPtxAs2)5 single crystal. Phys. Rev. B 2012, 85, 104512. [Google Scholar] [CrossRef]
  5. Stürzer, T.; Derondeau, G.; Johrendt, D. Role of different negatively charged layers in Ca10(FeAs)10(Pt4As8) and superconductivity at 30 K in electron-doped (Ca0.8La0.2)10(FeAs)10(Pt3As8). Phys. Rev. B 2012, 86, 060516(R). [Google Scholar] [CrossRef]
  6. Ni, N.; Straszheim, W.E.; Williams, D.J.; Tanatar, M.A.; Prozorov, R.; Bauer, E.D.; Ronning, F.; Thompson, J.D.; Cava, R.J. Transport and thermodynamic properties of (Ca1−xLax)10 (Pt3As8)(Fe2As2)5 superconductors. Phys. Rev. B 2013, 87, 060507(R). [Google Scholar] [CrossRef]
  7. Sturzer, T.; Derondeau, G.; Bertschler, E.M.; Johrendt, D. Superconductivity by rare earth doping in the 1038-type compounds (Ca1−xREx)10(FeAs)10(Pt3As8) with RE = Y, La–Nd, Sm–Lu. Solid State Commun. 2015, 201, 36–39. [Google Scholar] [CrossRef]
  8. Yun, E.G.; Ahmad, D.; Kim, G.C.; Kwon, Y.S.; Kim, Y.C. Fishtail Effect and the Upper Critical Field in a Superconducting Ca10(Pt4As8)(Fe2−xPtxAs2)5. J. Supercond. Novel Magn. 2017, 30, 2963–2969. [Google Scholar] [CrossRef]
  9. Shen, X.P.; Chen, S.D.; Ge, Q.Q.; Ye, Z.R.; Chen, F.; Xu, H.C.; Tan, S.Y.; Niu, X.H.; Fan, Q.; Xie, B.P.; et al. Electronic structure of Ca10(Pt4As8)(Fe2−xPtxAs2)5 with metallic Pt4As8 layers: An angle-resolved photoemission spectroscopy study. Phys. Rev. B 2013, 88, 115124. [Google Scholar] [CrossRef]
  10. Luo, X.G.; Chen, X.H. Crystal structure and phase diagrams of iron-based superconductors. Sci. China Mater. 2015, 58, 77–89. [Google Scholar] [CrossRef]
  11. Al’tshuler, B.L.; Aronov, A.G. Influence of electron-electron correlations on the resistivity of dirty metals. JETP Lett. 1978, 27, 662–664. [Google Scholar]
  12. Al’tshuler, B.L. Temperature dependence of impurity conductivity of metals at low temperatures. Sov. Phys. JETP 1978, 48, 670–675. [Google Scholar]
  13. Shein, I.R.; Ivanovshii, A.L. Ab initio study of the nature of the chemical bond and electronic structure of the layered phase Ca10(Pt4As8)(Fe2As2)5 as a parent system in the search for new superconducting iron-containing materials. Theor. Exp. Chem. 2011, 47, 292–295. [Google Scholar] [CrossRef]
  14. Nakamura, H.; Machida, M. First-principles study of Ca–Fe–Pt–As-type iron-based superconductors. Phys. C Supercond. 2013, 484, 39–42. [Google Scholar] [CrossRef]
  15. Anderson, P.W. The Theory of Superconductivity in the High-Tc Cuprates; Princeton University Press: Princeton, NJ, USA, 1997; pp. 57–110. [Google Scholar]
  16. Moll, P.J.W.; Zhu, X.Y.; Cheng, P.; Wen, H.H.; Batlogg, B. Intrinsic Josephson junctions in the iron-based multi-band superconductor (V2Sr4O6)Fe2As2. Nat. Phys. 2014, 10, 644–647. [Google Scholar] [CrossRef]
  17. Ji, H.S.; Lee, G.; Shim, J.H. Small anisotropy in iron-based superconductors induced by electron correlation. Phys. Rev. B 2011, 84, 054542. [Google Scholar] [CrossRef]
  18. Blatter, G.; Feigel’man, M.V.; Geshkenbein, V.B.; Larkin, A.I.; Vinokur, V.M. Vortices in high-temperature superconductors. Rev. Mod. Phys. 1994, 66, 1125–1388. [Google Scholar] [CrossRef]
  19. Yamamoto, A.; Jaroszynski, J.; Tarantini, C.; Balicas, L.; Jiang, J.; Gurevich, A.; Larbalestier, D.C.; Jin, R.; Sefat, A.S.; McGuire, M.A.; et al. Small anisotropy, weak thermal fluctuations, and high field superconductivity in Co-doped iron pnictide Ba(Fe1−xCox)2As2. Appl. Phys. Lett. 2009, 94, 062511. [Google Scholar] [CrossRef]
  20. Kano, M.; Kohama, Y.; Graf, D.; Balakirev, F.; Sefat, A.S.; Mcguire, M.A.; Sales, B.C.; Mandrus, D.; Tozer, S.W. Anisotropy of the Upper Critical Field in a Co-Doped BaFe2As2 Single Crystal. J. Phys. Soc. Jpn. 2009, 78, 084719. [Google Scholar] [CrossRef]
  21. Jaroszynski, J.; Hunte, F.; Balicas, L.; Jo, Y.J.; Raicevic, I.; Gurevich, A.; Larbalestier, D.C.; Balakirev, F.F.; Fang, L.; Cheng, P.; et al. Upper critical fields and thermally-activated transport of NdFeAsO0.7F0.3 single crystal. Phys. Rev. B 2008, 78, 174523. [Google Scholar] [CrossRef]
  22. Gurevich, A. Enhancement of the upper critical field by nonmagnetic impurities in dirty two-gap superconductors. Phys. Rev. B 2003, 67, 184515. [Google Scholar] [CrossRef]
  23. Werthamer, N.R.; Helfand, E.; Hohenberg, P.C. Temperature and Purity Dependence of the Superconducting Critical Field, Hc2. III. Electron Spin and Spin-Orbit Effects. Phys. Rev. 1966, 147, 295–302. [Google Scholar] [CrossRef]
  24. Zhao, J.; Yao, D.X.; Li, S.L.; Hong, T.; Chen, Y.; Chang, S.; Ratcliff, W., II; Lynn, J.W.; Mook, H.A.; Chen, G.F.; Luo, J.L.; et al. Low Energy Spin Waves and Magnetic Interactions in SrFe2As2. Phys. Rev. Lett. 2008, 101, 167203. [Google Scholar] [CrossRef] [PubMed]
  25. Wang, C.; Li, Y.K.; Zhu, Z.W.; Jiang, S.; Lin, X.; Luo, Y.K.; Chi, S.; Li, L.J.; Ren, Z.; He, M.; et al. Effects of cobalt doping and phase diagrams of LFe1−xCoxAsO (L = La and Sm). Phys. Rev. B 2009, 79, 054521. [Google Scholar] [CrossRef]
  26. Li, L.J.; Wang, Q.B.; Luo, Y.K.; Chen, H.; Tao, Q.; Li, Y.K.; Lin, X.; He, M.; Zhu, Z.W.; Cao, G.H.; et al. Superconductivity induced by Ni doping in BaFe2As2. New J. Phys. 2009, 11, 025008. [Google Scholar]
  27. Uher, C.; Huang, W.N. Thermoelectric power and thermal conductivity of neutron-irradiated YBa2Cu3O7−δ. Phys. Rev. B 1989, 40, 2694–2697. [Google Scholar] [CrossRef]
  28. Krishana, K.; Ong, N.P.; Li, Q.; Gu, G.D.; Koshizuka, N. Plateaus Observed in the Field Profile of Thermal Conductivity in the Superconductor Bi2Sr2CaCu2O8. Science 1997, 277, 83–85. [Google Scholar] [CrossRef]
  29. Sefat, A.S.; Mcguire, M.A.; Sales, B.C.; Jin, R.; Howe, J.Y.; Mandrus, D. Electronic correlations in the superconductor LaFeAsO0.89F0.11 with low carrier density. Phys. Rev. B 2008, 77, 174503. [Google Scholar] [CrossRef]
Figure 1. (a) Crystal structure of Ca10Pt4As8(Fe2As2)5; (b) as prepared Ca10(Pt4As8)(Fe2−xPtxAs2)5 single crystals; (c) XRD spectrum with index peaks of Ca10(Pt4As8)(Fe2−xPtxAs2)5 measured on single crystals along the c-axis at room temperature; (d) XRD spectrum of powder Ca10(Pt4As8)(Fe2−xPtxAs2)5 at room temperature.
Figure 1. (a) Crystal structure of Ca10Pt4As8(Fe2As2)5; (b) as prepared Ca10(Pt4As8)(Fe2−xPtxAs2)5 single crystals; (c) XRD spectrum with index peaks of Ca10(Pt4As8)(Fe2−xPtxAs2)5 measured on single crystals along the c-axis at room temperature; (d) XRD spectrum of powder Ca10(Pt4As8)(Fe2−xPtxAs2)5 at room temperature.
Materials 12 00474 g001
Figure 2. EDX measurement spectrum of Ca10Pt5.4Fe8.6As18; the inset figure is a surface SEM image of the measured sample.
Figure 2. EDX measurement spectrum of Ca10Pt5.4Fe8.6As18; the inset figure is a surface SEM image of the measured sample.
Materials 12 00474 g002
Figure 3. (a) The temperature dependence of the in-plane resistivity ρab and (b) out-of-plane resistivity ρc.
Figure 3. (a) The temperature dependence of the in-plane resistivity ρab and (b) out-of-plane resistivity ρc.
Materials 12 00474 g003
Figure 4. (a) A schematic plot of resistors made of FeAs layers and Pt4As8 layers connected in parallel; (b) a schematic plot of resistors made of FeAs layers and Pt4As8 layers connected in series.
Figure 4. (a) A schematic plot of resistors made of FeAs layers and Pt4As8 layers connected in parallel; (b) a schematic plot of resistors made of FeAs layers and Pt4As8 layers connected in series.
Materials 12 00474 g004
Figure 5. The temperature dependence of normal state anisotropy γ = ρ c / ρ a b .
Figure 5. The temperature dependence of normal state anisotropy γ = ρ c / ρ a b .
Materials 12 00474 g005
Figure 6. (a) The temperature dependence of ρab with applied magnetic field H = 0, 1, 3, 5, 8, 10, and 14 T along the c direction in the superconducting state; (b) the temperature dependence of ρc with applied magnetic field H = 0, 1, 3, 5, 8, 10, and 14 T along the ab plane in the superconducting state.
Figure 6. (a) The temperature dependence of ρab with applied magnetic field H = 0, 1, 3, 5, 8, 10, and 14 T along the c direction in the superconducting state; (b) the temperature dependence of ρc with applied magnetic field H = 0, 1, 3, 5, 8, 10, and 14 T along the ab plane in the superconducting state.
Materials 12 00474 g006
Figure 7. Temperature dependence of the upper critical field Hc2(T) with (a) H//c and(b) H//ab.
Figure 7. Temperature dependence of the upper critical field Hc2(T) with (a) H//c and(b) H//ab.
Materials 12 00474 g007
Figure 8. The temperature dependence of the Hc2 anisotropy parameter Γ H = H c 2 a b H c 2 c .
Figure 8. The temperature dependence of the Hc2 anisotropy parameter Γ H = H c 2 a b H c 2 c .
Materials 12 00474 g008
Figure 9. Hall coefficients RH(T) (blue dots) and carrier concentration (red circles) between 40 K and 300 K for Ca10(Pt4As8)((Fe0.86Pt0.14)2As2)5. The solid lines are guides for the eye.
Figure 9. Hall coefficients RH(T) (blue dots) and carrier concentration (red circles) between 40 K and 300 K for Ca10(Pt4As8)((Fe0.86Pt0.14)2As2)5. The solid lines are guides for the eye.
Materials 12 00474 g009
Figure 10. (a) Temperature dependence of thermal conductivity and the Seebeck coefficient for Ca10(Pt4As8)((Fe0.86Pt0.14)2As2)5; (b) electronic contribution to thermal conductivity above transition temperature Tc evaluated by the Wiedemann–Franz law.
Figure 10. (a) Temperature dependence of thermal conductivity and the Seebeck coefficient for Ca10(Pt4As8)((Fe0.86Pt0.14)2As2)5; (b) electronic contribution to thermal conductivity above transition temperature Tc evaluated by the Wiedemann–Franz law.
Materials 12 00474 g010
Table 1. The upper critical field and coherence length of Ca10(Pt4As8)((Fe0.86Pt0.14)2As2)5.
Table 1. The upper critical field and coherence length of Ca10(Pt4As8)((Fe0.86Pt0.14)2As2)5.
Criteria in Determining Hc2 d H c 2 c d T T C ( T / K ) H c 2 c ( 0 ) (Tesla) (WHH Approach) d H c 2 a b d T T C ( T / K ) H c 2 a b ( 0 ) (Tesla) (WHH Approach)ξab(0) (Å)ξc(0) (Å)
0.9 ρn−6.5 ± 0.2138 ± 5−14.1 ± 0.1313 ± 615.5 ± 0.26.8 ± 0.3
0.5 ρn−0.89 ± 0.119.7 ± 2−4.4 ± 0.199 ± 240.9 ± 0.58.1 ± 0.3
0.1 ρn−0.74 ± 0.116.4 ± 1−3.8 ± 0.184 ± 344.8 ± 0.48.7 ± 0.5
WHH: Werthamer–Helfand–Hohenberg.

Share and Cite

MDPI and ACS Style

Wu, D.; Meng, X.; Zhai, Y.; Yu, H.; Yu, J.; Qi, Y. Electrical and Thermal Transport Properties of Layered Superconducting Ca10(Pt4As8)((Fe0.86Pt0.14)2As2)5 Single Crystal. Materials 2019, 12, 474. https://doi.org/10.3390/ma12030474

AMA Style

Wu D, Meng X, Zhai Y, Yu H, Yu J, Qi Y. Electrical and Thermal Transport Properties of Layered Superconducting Ca10(Pt4As8)((Fe0.86Pt0.14)2As2)5 Single Crystal. Materials. 2019; 12(3):474. https://doi.org/10.3390/ma12030474

Chicago/Turabian Style

Wu, Dapeng, Xiaodong Meng, Yingying Zhai, Huaming Yu, Jiao Yu, and Yang Qi. 2019. "Electrical and Thermal Transport Properties of Layered Superconducting Ca10(Pt4As8)((Fe0.86Pt0.14)2As2)5 Single Crystal" Materials 12, no. 3: 474. https://doi.org/10.3390/ma12030474

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop