Next Article in Journal
Two-Time Correlation Functions in Dissipative and Interacting Bose–Hubbard Chains
Next Article in Special Issue
Layered Superconductors
Previous Article in Journal
Nanoscale Phase Separation and Lattice Complexity in VO2: The Metal–Insulator Transition Investigated by XANES via Auger Electron Yield at the Vanadium L23-Edge and Resonant Photoemission
Previous Article in Special Issue
Crystal Growth Techniques for Layered Superconductors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Single Crystal Growth and Superconducting Properties of Antimony-Substituted NdO0.7F0.3BiS2

1
Department of Physics, Tokyo University of Science, 1-3 Kagurazaka, Shinjuku-ku, Tokyo 162-8601, Japan
2
The World Premier International Center for Materials Nanoarchitectonics (WPI-MANA), National Institute for Materials Science (NIMS), 1-2-1 Sengen, Tsukuba, Ibaraki 305-0047, Japan
*
Author to whom correspondence should be addressed.
Condens. Matter 2018, 3(1), 1; https://doi.org/10.3390/condmat3010001
Submission received: 25 October 2017 / Revised: 18 December 2017 / Accepted: 19 December 2017 / Published: 27 December 2017
(This article belongs to the Special Issue Layered Superconductors)

Abstract

:
Antimony (Sb) substitution of less than 8% was examined on a single crystal of a layered superconductor NdO0.7F0.3BiS2. The superconducting transition temperature of the substituted samples decreased as Sb concentration increased. A lattice constant along the c-axis showed a large decrease compared with that along the a-axis. Since in-plane chemical pressure monotonically decreased as Sb concentration increased, the suppression of the superconductivity is attributed to the decrease in the in-plane chemical pressure.

1. Introduction

New layered BiS2-based superconductors Bi4O4S3 and Ln(O,F)BiS2 (Ln = La, Ce, Pr, Nd, Sm, Yb, and Bi) have been reported [1,2,3,4,5,6,7,8,9]. Although a superconducting transition temperature (Tc) of Ln(O,F)BiS2 is around 4 K, Tc is increased when a lattice strain is introduced. One of the ways to introduce the lattice strain is by partial element substitution with a different ionic radius into Ln, Bi, or S site [10,11,12,13,14,15,16]. The strain introduced by partial element substitution in these crystals is the chemical pressure. For instance, when selenium (Se) is partially substituted with sulfur (S) in the superconducting layer in La(O,F)BiS2, Tc and a superconducting volume fraction increase [10]. The substitution for a smaller lanthanide ion into the Ln site in the block layer also increases Tc [11,12,13,14]. The increase in Tc by element substitution is correlated with the in-plane chemical pressure defined by Mizuguchi: Tc increases as in-plane chemical pressure increases [17]. The change in Tc by the element substitution of Ln or S sites is attributed to the increase in in-plane chemical pressure. On the other hand, it is not known whether the change in Tc is understood in terms of in-plane chemical pressure when an element is partially substituted into the Bi site.
Here, we report on an examination of a substitution of antimony (Sb) ions of less than 8% into NdO0.7F0.3BiS2. Since the Sb ion has the same valence as that of the Bi ion, the effect of the chemical pressure can be investigated independently of the effect of the carrier concentration. We found that the lattice constant along the c-axis of Sb-substituted samples decreased. Furthermore, superconductivity was suppressed by the Sb substitution. This suppression by Sb substitution can be explained by the change in the in-plane chemical pressure.

2. Experimental Section

NdO0.7F0.3Bi1−xSbxS2(x = 0.01–0.08) single crystals were synthesized using a CsCl/KCl flux method in an evacuated quartz tubes [18,19]. Powders of Nd2S3, Bi2O3, Bi2S3, Sb2S3, and BiF3 with Bi grains were used as a starting material. The Bi2S3 powders were prepared by reacting Bi and S grains in an evacuated quartz tube at 500 °C for 10 h. A mixture of 0.8 g of starting material and 8 g of CsCl/KCl powder was sealed in an evacuated quartz tube. The tube was heated at 800 °C for 10 h, kept at 800 °C for 10 h, and cooled down to 630 °C at the rate of 0.3 °C/h or 1 °C/h. After this thermal process, the obtained material was washed by distilled water to remove the flux. X-ray diffraction (XRD) patterns using single crystal and powder samples were collected by a Rigaku X-ray diffractometer (Rigaku, Tokyo, Japan) with Cu Kα radiation using a θ–2θ method. These powder samples used in XRD measurements were prepared by grinding single crystals. Lattice constants along the a- and c-axes were determined using the θ–2θ technique. The surface condition of the single crystals was observed by a scanning electron microscope (SEM) (JEOL, Tokyo, Japan). The actual chemical composition was determined by an energy-dispersive X-ray (EDX) measurement (JEOL, Tokyo, Japan). Temperature dependence of the magnetic susceptibility down to 2 K was measured with the MPMS (magnetic property measurement system) (Quantum Design, San Diego, CA, USA). Temperature dependence of the electrical resistivity was measured down to 2.5 K with the four terminals method.

3. Results

Figure 1a shows X-ray diffraction patterns for NdO0.7F0.3Bi1−xSbxS2 (x = 0.01–0.08). All peaks correspond to the (00l) peaks of the CeOBiS2 type structure with the space group P4/nmm symmetry. The (004) peaks of these samples are magnified in Figure 1b. The peaks are gradually shifted to the high angle side as the Sb concentration increased. This is indicative of success in Sb substitution into NdO0.7F0.3BiS2 up to x = 0.08. Figure 1c shows an Sb concentration dependence on the lattice constant along the c-axis. The lattice constant is almost constant up to x = 0.04 and decreases when x > 0.04. A lattice constant along the a-axis shows almost constant until x = 0.08, as depicted in Figure 1d. The lattice shrinkage in the c-axis by Sb substitution was caused by the difference in ionic radius between the Bi and Sb ions, as we expected.
To evaluate sample quality, SEM and EDX measurements were performed in single crystals where x = 0.01 and 0.08, as shown in Figure 2. Figure 2a,b show SEM images of these obtained single crystals where x = 0.01 and 0.08, respectively. A square shape reflected from the tetragonal crystal structure can be seen. The shape did not change as the Sb concentration increased. In addition, the distribution of all elements included in the crystal was evaluated for the sample where x = 0.08 in Figure 2c–h. All elements were homogeneously distributed in the sample. Therefore, the Sb ion was homogeneously substituted into the single crystal of NdO0.7F0.3BiS2 until x = 0.08. Furthermore, an actual Sb concentration was estimated for samples where x = 0.01 and 0.08. The actual Sb concentrations of the samples with a nominal composition of 0.01 and 0.08 were around 0.001 and 0.021, respectively. The actual composition was quite lower than the nominal composition. Therefore, the actual Sb concentration increased until around 0.02 as nominal Sb concentration increased until 0.08. Next, the superconducting properties of these samples were evaluated to investigate the effect of Sb substitution on superconductivity.
Figure 3a,b show the magnetic susceptibility as a function of temperature for NdO0.7F0.3Bi1−xSbxS2 (x = 0.01–0.08) at a magnetic field of 10 Oe. All samples show diamagnetic signal due to an appearance of superconductivity. A magnitude of 4πχ decreases from 1 at x = 0.01 to 0.2 at x = 0.08. The superconducting transition temperature (Tc) monotonously decreases with increasing Sb concentration as shown in Figure 3b. In the zero field cooling process, all samples show a broad superconducting transition. This broad transition is often observed in the other Ln(O,F)BiCh2 (Ln = La, Ce, Pr, Nd, Ch = S, Se) [20,21,22,23,24]. Although the reason why the transition is broadened is not yet understood, the broad transition seems to be a common feature in these materials.
Figure 4b shows the electrical resistivity near Tc as a function of temperature. A steep drop due to superconductivity can be observed at around 5.2 K in the sample where x = 0.01. The transition temperature decreased as the Sb concentration increased. Resistivity at room temperature (RT) continuously increased as the Sb substitution increased, as shown in Figure 4a. This is because the Sb ion was substituted for the Bi ion located in the conduction layers. This increase in resistivity at RT has also been observed in Pb-substituted Ln(O,F)BiS2 (Ln = La, Nd) [15,20], on the other hand Pb substitution induced an enhancement of Tc. Therefore, the effect of substitution on superconductivity does not correlate to the strength of the scattering by the substituted atoms.
Figure 5a shows a summary of the observed Tc. Here, Tcmag is defined as the temperature at which magnetic susceptibility begins to decrease. Tconset and Tczero are defined as the temperature at which resistivity begins to decrease and the temperature where resistivity becomes zero, respectively. All Tc values monotonically decrease with an increase in Sb concentration.
Mizuguchi et al. reported that the Tc of BiCh2 superconductors is correlated with in-plane chemical pressure [17]. To compare the result in Sb-substituted samples with the previous report, we estimated the in-plane chemical pressure for NdO0.7F0.3Bi1−xSbxS2. In-plane chemical pressure is defined by the ionic radius of the Bi and chalcogenide ions, as well as the distance between the Bi and chalcogenide ions located in the BiCh2 plane. In the case of Sb substitution, this in-plane chemical pressure is defined as follows:
in-plane chemical pressure = ((1 − x)RBi + xRSb + RCh)/Bi-Ch1 (in-plane) distance
where RBi is the ionic radius of the Bi ion. In this study, the ionic radius of the Bi ion was obtained from the structural data of the LaO0.77F0.23BiS2 single crystal [25]. Thus, the value of RBi was 104.75 pm, which is almost the same value as that in the previous report [17]. RSb and RCh are the ionic radii of Sb3+ and Ch2− ions, respectively. These values were determined as 76 and 184 pm, based on [26]. The composition of the Sb ion was used as the nominal value because this value in [17] was also calculated by using the nominal composition of the substituted ions. The Bi–Ch1 distance in this study was estimated by the lattice constant along the a-axis. Since the lattice constant along the a-axis corresponds to the length of a side of the Bi square lattice, we used 1/ 2 times the lattice constant along the a-axis as the Bi–Ch1 distance. Figure 5b shows the estimated in-plane chemical pressure dependence of Tconset, Tcmag, and Tczero, shown in open circles, filled circles, and squares, respectively. The Tc of all samples decreases with the decrease in in-plane chemical pressure. Based on this result, Tc in the Sb-substituted materials is positively correlated to the in-plane chemical pressure. To compare this result with the results with respect to CeyNd1−yO0.5F0.5BiS2 and NdySm1−yO0.5F0.5BiS2 in [17], the Tc of CeyNd1−yO0.5F0.5BiS2 and NdySm1−yO0.5F0.5BiS2 are superimposed in Figure 5b, which are shown in filled triangles. The Tc in these materials shows a same change as our results against the in-plane chemical pressure. Thus, our result is qualitatively consistent with the previous results of [17]. In addition, these results suggest that the concept of in-plane chemical pressure is valid in the case of substitution not only at the Ln site in the blocking layer but also at the Bi site in the conduction layer.
In this study, the in-plane chemical pressure was estimated by the lattice constant along the a-axis. On the other hand, the in-plane chemical pressure in [17] was estimated by the Bi–Ch1 distance, which was obtained from single crystal analysis. To compare the difference between these estimated in-plane chemical pressure values, both values were estimated using data from the single crystal analysis in [23]. In Figure 5b, the in-plane chemical pressure estimated by the Bi–Ch1 distance is plotted as a triangle outline, while that estimated by the a-axis is plotted as an upside-down triangle. The latter, compared with the former, has a slightly higher value. However, since this value is located at the extrapolated line of the Tc of the Sb-substituted samples, this estimation can be used to estimate the in-plane chemical pressure qualitatively.
Finally, we compared this effect of Sb substitution with the effect of Pb substitution in NdO0.7F0.3BiS2 because this Pb ion is expected to be substituted into the Bi site. Sb substitution suppresses superconductivity. This suppression is explained by the decrease in in-plane chemical pressure described in this paper. On the other hand, Pb substitution increased Tc to up to 6 K. The in-plane chemical pressure of Pb-substituted samples is almost constant because an ionic radius of the Pb2+ ion is almost the same as that of the Bi2.7+ ion calculated by the single crystal analysis. The ionic radius of the Pb2+ ion is 100 pm [26]. If the Pb ion is substituted until x = 0.10, RBi + RPb is around 103.8 pm, which is almost the same value of the ionic radius of the Bi2.7+ ion of 104.75 pm. This indicates that the Tc of Pb-substituted samples is independent of the value of the in-plane chemical pressure.

4. Conclusions

An examination of Sb substitution of less than 8% was performed with NdO0.7F0.3BiS2. The Sb substitution caused a decrease in lattice constants along the c-axis, as well as a decrease in Tc and the superconducting volume fraction. The deterioration by these superconducting properties by the partial Sb substitution into the Bi site is explained in terms of the decrease in in-plane chemical pressure, indicating the utility of in-plane chemical pressure in BiCh2 superconductors.

Acknowledgments

This work was partly supported by a Grant-in-Aid for Young Scientists (B) (No. 15K17710) and the Nanotech Career-up Alliance (Nanotech CUPAL).

Author Contributions

Satoshi Otsuki, Satoshi Demura and Hideaki Sakata conceived and designed the experiments; Satoshi Otsuki mainly performed the experiments; Satoshi Otsuki, Satoshi Demura and Yuita Fujisawa contributed the data analysis; Yoshihiko Takano contributed analysis tools; Satoshi Demura wrote the paper. All authors have read and approved the final manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mizuguchi, Y.; Fujihisa, H.; Gotoh, Y.; Suzuki, K.; Usui, H.; Kuroki, K.; Demura, S.; Takano, Y.; Izawa, H.; Miura, O. BiS2-based layered superconductor Bi4O4S3. Phys. Rev. B 2012, 86, 220510. [Google Scholar] [CrossRef]
  2. Singh, S.K.; Kumar, A.; Gahtri, B.; Sharma, G.; Patnaik, S.; Awana, V.P.S. Bulk Superconductivity in Bismuth Oxysulfide Bi4O4S3. J. Am. Chem. Soc. 2012, 134, 16504–16507. [Google Scholar] [CrossRef] [PubMed]
  3. Mizuguchi, Y.; Demura, S.; Deguchi, K.; Takano, Y.; Fujihisa, H.; Gotoh, Y.; Izawa, H.; Miura, O. Superconductivity in Novel BiS2-Based Layered Superconductor LaO1−xFxBiS2. J. Phys. Soc. Jpn. 2012, 81, 114725. [Google Scholar] [CrossRef]
  4. Awana, V.P.S.; Kumar, A.; Jha, R.; Kumar Singh, S.; Pal, A.; Saha, S.J.; Patnaik, S. Appearance of superconductivity in layered LaO0.5F0.5BiS2. Solid State Commun. 2013, 157, 21–23. [Google Scholar] [CrossRef]
  5. Demura, S.; Mizuguchi, Y.; Deguchi, K.; Okazaki, H.; Hara, H.; Watanabe, T.; Denholme, S.J.; Fujioka, M.; Ozaki, T.; Fujihisa, H.; et al. New Member of BiS2-Based Superconductor NdO1−xFxBiS2. J. Phys. Soc. Jpn. 2013, 82, 033708. [Google Scholar] [CrossRef]
  6. Jha, R.; Kumar, A.; Singh, S.K.; Awana, V.P.S. Superconductivity at 5 K in NdO0.5F0.5BiS2. J. Appl. Phys. 2013, 113, 056102. [Google Scholar] [CrossRef]
  7. Xing, J.; Li, S.; Ding, X.; Yang, H.; Wen, H.-H. Superconductivity appears in the vicinity of semiconducting-like behavior in CeO1−xFxBiS2. Phys. Rev. B 2012, 86, 214518. [Google Scholar] [CrossRef]
  8. Jha, R.; Kumar, A.; Sngh, S.K.; Awana, V.P.S. Synthesis and Superconductivity of New BiS2 Based Superconductor PrO0.5F0.5BiS2. J. Supercond. Nov. Magn. 2013, 26, 499–502. [Google Scholar] [CrossRef]
  9. Yazici, D.; Huang, K.; White, B.D.; Jeon, I.; Burnett, V.W.; Friedman, A.J.; Lum, I.K.; Nallaiyan, M.; Spagna, S.; Maple, M.B. Superconductivity induced by electron doping in La1−xMxOBiS2 (M = Ti, Zr, Hf, Th). Phys. Rev. B 2013, 87, 174512. [Google Scholar] [CrossRef]
  10. Hiroi, T.; Kajitani, J.; Omachi, A.; Miura, O.; Mizuguchi, Y. Evolution of Superconductivity in BiS2-Based Superconductor LaO0.5F0.5Bi(S1−xSex)2. J. Phys. Soc. Jpn. 2015, 84, 024723. [Google Scholar] [CrossRef]
  11. Kajitani, J.; Omachi, A.; Hiroi, T.; Miura, O.; Mizuguchi, Y. Chemical pressure effect on Tc in BiS2-based Ce1−xNdxO0.5F0.5BiS2. Phys. C Supercond. Appl. 2014, 504, 33–35. [Google Scholar] [CrossRef]
  12. Fang, Y.; Yazici, D.; White, B.D.; Maple, M.B. Enhancement of superconductivity in La1−xSmxO0.5F0.5BiS2. Phys. Rev. B 2015, 91, 064510. [Google Scholar] [CrossRef]
  13. Yazici, D.; Jeon, I.; White, B.D.; Maple, M.B. Superconductivity in layered BiS2-based compounds. Phys. C Supercond. Appl. 2015, 514, 218–236. [Google Scholar] [CrossRef]
  14. Fang, Y.; Wolowiec, C.T.; Yazici, D.; Maple, M.B. Chemical Substitution and High Pressure Effects on Superconductivity in the LnOBiS2(Ln = La-Nd) System. Nov. Supercond. Mater. 2015, 1, 79–94. [Google Scholar] [CrossRef]
  15. Demura, S.; Fujisawa, Y.; Otuski, S.; Ishio, R.; Takano, Y.; Sakata, H. Enhancement of Tc in BiS2-based superconductors NdO0.7F0.3BiS2 by substitution of Pb for Bi. Solid State Commun. 2015, 223, 40–44. [Google Scholar] [CrossRef]
  16. Demura, S. Evolution of superconductivity and magnetism in BiS2-based layered compounds. Nov. Supercond. Mater. 2016, 2, 1–15. [Google Scholar] [CrossRef]
  17. Mizuguchi, Y.; Miura, A.; Kajitani, J.; Hiroi, T.; Miura, O.; Tadanaga, K.; Kumada, N.; Magome, E.; Miriyoshi, C.; Kuroiwa, Y. In-plane chemical pressure essential for superconductivity in BiCh2-based (Ch: S, Se) layered structure. Sci. Rep. 2015, 5, 14968. [Google Scholar] [CrossRef] [PubMed]
  18. Nagao, M.; Miura, A.; Demura, S.; Deguchi, K.; Watauchi, S.; Takei, T.; Takano, Y.; Kumada, N.; Tanaka, I. Growth and superconducting properties of F-substituted ROBiS2 (R = La, Ce, Nd) single crystals. Solid State Commun. 2014, 178, 33–36. [Google Scholar] [CrossRef]
  19. Nagao, M. Growth and characterization of R(O,F)BiS2 (R = La,Ce, Pr, Nd) superconducting single crystals. Nov. Supercond. Mater. 2015, 1, 64. [Google Scholar] [CrossRef]
  20. Otski, S.; Demura, S.; Sakai, Y.; Fujisawa, Y.; Sakata, H. Effect of Lead substitution on LaO0.5F0.5BiS2. Solid State Commun. 2017, 270, 17–21. [Google Scholar] [CrossRef]
  21. Miura, A.; Nagao, M.; Takei, T.; Watauchi, S.; Mizuguchi, Y.; Takano, Y.; Tanaka, I.; Kumada, N. Structure, Superconductivity, and Magnetism of Ce(O,F)BiS2 Single Crystals. Cryst. Growth Des. 2015, 15, 39–44. [Google Scholar] [CrossRef]
  22. Nagao, M.; Miura, A.; Watauchi, S.; Takano, Y.; Tanaka, I. C-axis electrical resistivity of PrO1−aFaBiS2 single crystals. Jpn. J. Appl. Phys. 2015, 54, 083101. [Google Scholar] [CrossRef]
  23. Nagao, M.; Demura, S.; Deguchi, K.; Miura, A.; Watauchi, S.; Takei, T.; Takano, Y.; Kumada, N.; Tanaka, I. Structural Analysis and Superconducting Properties of F-Substituted NdOBiS2 Single Crystals. J. Phys. Soc. Jpn. 2013, 82, 113701. [Google Scholar] [CrossRef]
  24. Tanaka, M.; Nagao, M.; Matsushita, Y.; Fujioka, M.; Denholme, S.J.; Yamaguchi, T.; Takeya, H.; Takano, Y. First single crystal growth and structural analysis of superconducting layered bismuth oxyselenide; La(O,F)BiSe2. J. Solid State Chem. 2014, 219, 168–172. [Google Scholar] [CrossRef]
  25. Miura, A.; Nagao, M.; Takei, T.; Watauchi, S.; Tanaka, I. Crystal structures of LaO1−xFxBiS2 (x~0.23, 0.46): Effect of F doping on distortion of Bi–S plane. J. Solid State Chem. 2014, 212, 213–217. [Google Scholar] [CrossRef]
  26. Jia, Y.Q. Crystal radii and effective ionic radii of the rare earth ions. J. Solid State Chem. 1991, 95, 184–187. [Google Scholar] [CrossRef]
Figure 1. XRD patterns and lattice parameters of NdO0.7F0.3Bi1−xSbxS2 single crystals. (a) XRD patterns from x = 0.01 to 0.08 on the NdO0.7F0.3Bi1−xSbxS2. (b) Magnified figure near the (004) peak in Figure 1a. (c,d) Lattice constants along the c- and a-axes for NdO0.7F0.3Bi1−xSbxS2.
Figure 1. XRD patterns and lattice parameters of NdO0.7F0.3Bi1−xSbxS2 single crystals. (a) XRD patterns from x = 0.01 to 0.08 on the NdO0.7F0.3Bi1−xSbxS2. (b) Magnified figure near the (004) peak in Figure 1a. (c,d) Lattice constants along the c- and a-axes for NdO0.7F0.3Bi1−xSbxS2.
Condensedmatter 03 00001 g001aCondensedmatter 03 00001 g001b
Figure 2. (a,b) SEM images of NdO0.7F0.3Bi1−xSbxS2 single crystals (x = 0.01, 0.08). (ch) Mapping of the chemical composition of Nd, Bi, Sb, S, O, and F for the single crystal of NdO0.7F0.3Bi1−xSbxS2 (x = 0.08).
Figure 2. (a,b) SEM images of NdO0.7F0.3Bi1−xSbxS2 single crystals (x = 0.01, 0.08). (ch) Mapping of the chemical composition of Nd, Bi, Sb, S, O, and F for the single crystal of NdO0.7F0.3Bi1−xSbxS2 (x = 0.08).
Condensedmatter 03 00001 g002
Figure 3. Magnetic susceptibility for NdO0.7F0.3Bi1−xSbxS2 as a function of temperature. (a) Temperature dependence of magnetic susceptibility for NdO0.7F0.3Bi1−xSbxS2 from 1 to 6 K. (b) The magnetic susceptibility near the superconducting transition for NdO0.7F0.3Bi1−xSbxS2.
Figure 3. Magnetic susceptibility for NdO0.7F0.3Bi1−xSbxS2 as a function of temperature. (a) Temperature dependence of magnetic susceptibility for NdO0.7F0.3Bi1−xSbxS2 from 1 to 6 K. (b) The magnetic susceptibility near the superconducting transition for NdO0.7F0.3Bi1−xSbxS2.
Condensedmatter 03 00001 g003
Figure 4. Temperature dependence of electrical resistivity for NdO0.7F0.3Bi1−xSbxS2. (a) Temperature dependence of electrical resistivity for NdO0.7F0.3Bi1−xSbxS2 (x = 0.01–0.08) from 2 to 300 K. (b) Temperature dependence of electrical resistivity for NdO0.7F0.3Bi1−xSbxS2 (x = 0.01–0.08) from 2 to 10 K.
Figure 4. Temperature dependence of electrical resistivity for NdO0.7F0.3Bi1−xSbxS2. (a) Temperature dependence of electrical resistivity for NdO0.7F0.3Bi1−xSbxS2 (x = 0.01–0.08) from 2 to 300 K. (b) Temperature dependence of electrical resistivity for NdO0.7F0.3Bi1−xSbxS2 (x = 0.01–0.08) from 2 to 10 K.
Condensedmatter 03 00001 g004
Figure 5. (a) Sb concentration x dependence of Tc for NdO0.7F0.3Bi1−xSbxS2. (b) Estimated in-plane chemical pressure dependence of Tc for NdO0.7F0.3Bi1−xSbxS2. The Tc values of (Ce,Nd)O0.5F0.5BiS2 and (Nd,Sm)O0.5F0.5BiS2 are plotted as triangles, which were fitted by hand on the dotted line [18].
Figure 5. (a) Sb concentration x dependence of Tc for NdO0.7F0.3Bi1−xSbxS2. (b) Estimated in-plane chemical pressure dependence of Tc for NdO0.7F0.3Bi1−xSbxS2. The Tc values of (Ce,Nd)O0.5F0.5BiS2 and (Nd,Sm)O0.5F0.5BiS2 are plotted as triangles, which were fitted by hand on the dotted line [18].
Condensedmatter 03 00001 g005

Share and Cite

MDPI and ACS Style

Demura, S.; Otsuki, S.; Fujisawa, Y.; Takano, Y.; Sakata, H. Single Crystal Growth and Superconducting Properties of Antimony-Substituted NdO0.7F0.3BiS2. Condens. Matter 2018, 3, 1. https://doi.org/10.3390/condmat3010001

AMA Style

Demura S, Otsuki S, Fujisawa Y, Takano Y, Sakata H. Single Crystal Growth and Superconducting Properties of Antimony-Substituted NdO0.7F0.3BiS2. Condensed Matter. 2018; 3(1):1. https://doi.org/10.3390/condmat3010001

Chicago/Turabian Style

Demura, Satoshi, Satoshi Otsuki, Yuita Fujisawa, Yoshihiko Takano, and Hideaki Sakata. 2018. "Single Crystal Growth and Superconducting Properties of Antimony-Substituted NdO0.7F0.3BiS2" Condensed Matter 3, no. 1: 1. https://doi.org/10.3390/condmat3010001

Article Metrics

Back to TopTop