Next Article in Journal / Special Issue
Effect of Pressure Treatment on the Specific Surface Area in Kaolin Group Minerals
Previous Article in Journal
Structure of Cubic Al73.8Pd13.6Fe12.6 Phase with High Al Content
Previous Article in Special Issue
A New High-Pressure Phase Transition in Natural Gedrite
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

New Tetragonal ReGa5(M) (M = Sn, Pb, Bi) Single Crystals Grown from Delicate Electrons Changing

by
Madalynn Marshall
1,
Karolina Górnicka
2,
Ranuri S. Dissanayaka Mudiyanselage
1,
Tomasz Klimczuk
2 and
Weiwei Xie
1,*
1
Department of Chemistry, Louisiana State University, Baton Rouge, LA 70803, USA
2
Faculty of Applied Physics and Mathematics, Gdansk University of Technology, Narutowicza 11/12, 80-233 Gdansk, Poland
*
Author to whom correspondence should be addressed.
Submission received: 27 August 2019 / Revised: 29 September 2019 / Accepted: 1 October 2019 / Published: 14 October 2019
(This article belongs to the Special Issue High Pressure Synthesis in Crystalline Materials)

Abstract

:
Single crystals of the new Ga-rich phases ReGa~5(Sn), ReGa~5(Pb) and ReGa~5(Bi) were successfully obtained from the flux method. The new tetragonal phases crystallize in the space group P4/mnc (No. 128) with vertex-sharing capped Re2@Ga14 oblong chains. Vacancies were discovered on the Ga4 and Ga5 sites, which can be understood as the direct inclusion of elemental Sn, Pb and Bi into the structure. Heat capacity measurements were performed on all three compounds resulting in a small anomaly which resembles the superconductivity transition temperature from the impurity ReGa5 phase. The three compounds were not superconducting above 1.85 K. Subsequently, electronic structure calculations revealed a high density of states around the Fermi level, as well as non-bonding interactions that likely indicate the stability of these new phases.

1. Introduction

Endohedral gallium cluster phases are a chemical family that has been sought after for their various electronic properties arising from the interesting cluster chemistry associated with gallium’s moderate electronegativity [1,2]. However, because of overly large anionicity, isolated Ga clusters will very rarely form. This leads to a significant interplay between Ga cluster-cluster interactions, i.e., exo-bond formations and the stability of the electronic structures. While moving across the periodic table, a spectrum of diverse Ga cluster phases can be observed. Beginning with electropositive alkali metals (A), AmGan compounds most readily form electron precise Zintl phases as a result of the large electronegativity difference, making a promising candidate for thermoelectric materials [3,4,5]. In this sense, electrons will be transferred from the alkali metals to Gan clusters to satisfy its valence electron requirement. Many of these compounds will form deltahedral clusters, as in borane chemistry, and have skeletal electron counts which will typically follow Wade’s rules [6].
Decreasing the electronegativity difference and transitioning towards actinides and lanthanides (R), a significant reduction in the band gap is observed and, in some cases is completely diminished, resulting in metallic behavior. In search of stability, RmGan clusters often distort from ideal deltahedral symmetries and form exo-bonds [7]. The addition of transition metals into RmGan clusters can reduce the cluster charge and has led to various intriguing materials such as the unconventional superconductor PuCoGa5 [8]. Finally, coming to the transition metal (T) gallide clusters, the ionic behavior becomes more obscure as a result of the similar electronegativities between Ga and transition metals. In this region, the relationship between electron counts and cluster formation with regards to the superconductivity transition temperature, unfolds. Currently, the only known TmGan superconductors are low-temperature superconductors, making them potential materials for producing high magnetic fields at low temperatures. This can be observed in the superconductors Mo8Ga41 and Mo6Ga31—as the electron counts decrease, the Tc increases and the clusters form vertex-sharing interactions rather than edge-sharing [9]. MnGa4.96 is another Ga-rich cluster which crystallizes into a tetragonal unit cell with capped face sharing Mn@Ga8 clusters and correspondingly exhibits no superconductivity [10]. Therefore, the stoichiometry and valence electrons from the transition metal of endohedral gallide clusters play a critical role in the exo-bond formations, i.e., electron-rich clusters prefer edge-sharing while electron-poor ones prefer vertex-sharing clusters, which as a result, directly affects the Tc. Considering the significant decrease in Tc from Rh2Ga9 to Ir2Ga9, and contradictorily the increase in Tc from Cr-Ga to Mo-Ga clusters, the transition of Tc across TmGan clusters can be further analyzed [11,12,13].
Narrowing the spectrum down to the 5d transition metal gallide clusters below Mn in group 7, we sought to further investigate the recently discovered ReGa5 superconductor with a Tc of ~2.3 K and vertex sharing Re@Ga9 clusters [14]. Proceeding to understand the structural characteristics and electron counts of endohedral Ga clusters on superconductivity and further exploring new Re-Ga phases, we successfully discovered three new compounds ReGa~5(Sn), ReGa~5(Pb) and ReGa~5(Bi). Reported here are the heat capacity measurements along with the crystal and electronic structure characterizations. No superconductivity was found for ReGa~5(Sn), ReGa~5(Pb) and ReGa~5(Bi) down to 1.85 K.

2. Experimental Section

Synthesis. The new compounds ReGa~5(Sn), ReGa~5(Pb) and ReGa~5(Bi) were synthesized via flux method using Ga as the self-flux. Elements used include tin granules (99.9%, BTC), lead shots (99.999%, BTC), bismuth chunks (99.999%, lump, Alfa Aesar), gallium ingot (99.99% (metals basis), Alfa Aesar) and rhenium powder (−325 mesh, 99.99% (metals basis), Alfa Aesar). The three reactions were prepared with sample sizes of ~1.5–2.0 g and loading compositions of ReSnGa48, RePb5Ga25 and ReBi5Ga25. Each sample was placed in an alumina crucible then inside a silica tube. Quartz glass pieces and quartz wool were packed on top of the crucible as the filter. The silica tube was subsequently evacuated (<10−5 Torr) and sealed. Samples were heated to 950 °C at a rate of 200 °C/hr and annealed there for 24hr then slow cooled at a rate of 10 °C/hr for ReGa~5(Sn) and 4 °C/hr for ReGa~5(Pb) and ReGa~5(Bi) to 600 °C at which the samples were centrifuged. Excess Ga flux was removed using ~2 M HCl. All products are found to be stable in air and moisture.
Phase Analysis. For each ReGa~5(Sn), ReGa~5(Pb) and ReGa~5(Bi) sample the phase was identified and purity verified through a Rigaku MiniFlex 600 powder X-ray diffractometer using Cu Kα radiation (λ = 1.5406 Å, Ge monochromator) [15]. A scan speed of 1.25°/min and step of 0.005° were used over a Bragg angle (2θ) ranging from 5 to 90° for ReGa~5(Pb) and ReGa~5(Bi). For ReGa~5(Sn) a scan speed of 0.6°/min and step of 0.005° were used over a Bragg angle (2θ) ranging from 10 to 90°. Full Proof software was used to analyze the phase identification and lattice parameters of the experimental and theoretical powder patterns for ReGa~5(Sn), ReGa~5(Pb) and ReGa~5(Bi) and the experimental powder patterns for the impurity phases obtained from ICSD [16].
Structure Determination. Single crystals from ReGa~5(Sn), ReGa~5(Pb) and ReGa~5(Bi) were picked to perform a structural analysis using a Bruker Apex II diffractometer equipped with Mo radiation (λ = 0.71073 Å). Scattering intensity data were collected at room temperature with 0.5° per scan in ω and an exposure time of 10s per frame. The crystal structure was solved using a SHELXTL package with direct methods and full-matrix least-squares on F2 model [17,18].
Scanning Electron Microscopy. A FEI Quanta 3D Field Emission Gun (FEG) Focused Ion Beam (FIB)/Scanning Electron Microscope (SEM) and Energy Dispersive X-ray Spectroscopy (EDX) were utilized with the analysis of chemical stoichiometry. For each sample, multiple areas were selected for spectrum collection with a 20kV accelerating voltage and 100 seconds of scanning time.
Physical Property Measurements. Heat capacity measurements were carried out using the two- τ time- relaxation method in a Physical Property Measurement System (PPMS). The data was collected between 1.85 and 300 K. The sample was mounted to the measuring stage using Apiezon N grease to ensure good thermal contact.
Tight-Binding, Linear Muffin-Tin Orbital-Atomic Sphere Approximation (TB-LMTO-ASA) [19]. The TB-LMTO-ASA program with Stuttgart code was utilized to calculate the density of states (DOS) and Orbital Hamiltonian Population (COHP) curves of a hypothetically ordered “ReGa4.5” [20,21]. The convergence criterion was set to 0.05 meV. The Muffin-Tin radius (RMT) for each element includes: 0.995 Å for Re1; 1.40 Å for Ga2; 1.88 Å for Ga3; 0.83 Å for Ga4. The band structure and DOS were both calculated with a 4 × 4 × 2 k-point in the Brillouin zone [22,23].

3. Results and Discussions

Phase Analysis. As a result of increasing the valence electron count from the orthorhombic ReGa5, three new tetragonal phases ReGa~5(Sn), ReGa~5(Pb) and ReGa~5(Bi) were revealed. Unreacted Re was present in all of the three compounds ranging from 5–15% as well as ReGa5 from 0.1–39%, based on the HighScore Plus software. Unreacted Pb and Bi were also found in both ReGa~5(Pb) and ReGa~5 (Bi) by approximately 3 and 7%, respectively. A comparison of the three powder patterns, as well as images of the single crystals, are shown in Figure 1. Individual refined powder patterns of the three phases can be found in Figure S1 (Supplementary Materials) of the Supporting Information. To confirm the chemical composition and stoichiometry SEM-EDS was utilized. The determined chemical formulas are Re1.0(3)Ga5.0(2)Sn0.1(8), Re1.0(2)Ga5.0(2)Pb0.1(5) and Re1.0(2)Ga5.0(2)Bi0.2(4). A complete table of the SEM-EDS data is shown in Table S2 in the Supporting Information.
Crystal Structure. Single crystal X-ray diffraction was utilized to further understand the effect of atomic size and electron counts on the stability of Ga-rich phases. As a result of the increase in valence electron counts induced from the elements Sn, Pb and Bi, a tetragonal structure with space group P4/mnc (No. 128) was formed. A table of the single crystal refinement data as well as atomic coordinates and equivalent isotropic displacement parameters are given in Table 1 and Table 2. The three new phases ReGa~5(Sn), ReGa~5(Pb) and ReGa~5(Bi) consist of two face-sharing square antiprismatic Re@Ga8 polyhedra capped by four Ga atoms, or five Ga atoms when considering ReGa~5(Sn), on the remaining free square faces. Consequently, these clusters form networks of vertex sharing capped Re2@Ga14 oblong chains, similar to MnGa4.96. The Re@Ga8 clusters in ReGa~5(Sn), ReGa~5(Pb) and ReGa~5(Bi) resemble the geometry of the Re@Ga9 endohedral clusters found in the ReGa5 orthorhombic structure, however, in this case, the polyhedra are more than singly capped. Conversely, the Re@Ga9 clusters in ReGa5 are vertex sharing, while ReGa~5(Sn), ReGa~5(Pb) and ReGa~5(Bi) not only have vertex sharing but also face sharing polyhedra, which resultantly produces the capped Re2@Ga14 oblong chains. As seen in the decrease in Tc from the superconductor Mo8Ga41 to Mo6Ga31 the exo-bond formation, as well as the additional electron counts, may be a key factor causing the loss of superconductivity in ReGa~5(Sn), ReGa~5(Pb) and ReGa~5(Bi). Initially, atomic vacancies were tested and revealed vacancies on the Ga4 and Ga5 sites. The Ga4 site vacancies were found to vary significantly depending on the electron count. After further inspection of the change in atomic distances with varying electron count, it was found that the Re-Ga4 distance experiences the most significant change. A table of atomic distances for each new phase is given in Table 3. This occurrence may have been understood by the Re@Ga8 polyhedra undergoing stretching and compression due to the various applied chemical pressures. However, the atomic distances in the Re@Ga8 polyhedra appear to remain relatively consistent. The changes in distance experienced by Re-Ga4 resulted from the vacancies on the Ga4 site. As the vacancies decrease, the Ga4 site merges two atoms into one and the Re-Ga4 distance increases. This occurs as the structure changes from Sn to Bi to Pb, where ReGa~5(Sn) and ReGa~5(Bi) have Ga4 on the 4e site while ReGa~5(Pb) has Ga4 on the 2a site. As a result, the Ga4 and Ga5 vacancies can be realized as Sn, Pb and Bi giving a mixture of Ga and Sn, Pb or Bi on the Ga4 and Ga5 sites, as shown in Figure 2d. Therefore, considering both the atomic size and electron count of Sn, Pb and Bi the changing Ga4 vacancies and Re-Ga4 distance can be well understood. Taking into account that the most significant changes are resulting from the Re-Ga4 and the Ga5-Ga5 distances (diagonally along the b axis) then both can be recognized as contributing factors to the change in structure between the three compounds. The four Ga5 atoms that sit on either side of the Ga4 site seem to open and compress, in sync with the Ga5 occupancies as the Ga4 atomic vacancy and site location changes, subsequently pushing Ga4 from the 4e to the 2a site. This then increases the Re-Ga4 distance as the structures transition from ReGa~5(Sn) to ReGa~5(Bi) to ReGa~5(Pb).
Physical Properties. To evaluate the impact of chemical pressure on superconductivity, specific heat measurements were carried out. Figure 3a,c,e show the specific heat data plotted as, CP/T versus T2, and Figure 3b,d,f display CP versus T for ReGa~5(Sn), ReGa~5(Pb) and ReGa~5(Bi), respectively. The low-temperature experimental data (obtained under field of 0.3 T) were fitted using Cp/T = γ + βT2, where the first and second terms are attributed to the electronic (Cel) and lattice contributions (Cph) to the specific heat, respectively. The fit, represented by the red solid line (see Figure 3a,c,e), gives the Sommerfeld coefficient,γ,4.0(1), 3.4(1), and 3.5(2) mJ mol−1 K−2 and β equals to 0.39(2), 0.55(1) and 0.84(3) mJ mol−1 K−4 for ReGa~5(Sn), ReGa~5(Pb) and ReGa~5(Bi), respectively. Furthermore, the Debye temperature ΘD can be determined using the simple Debye model: Θ D = ( 12 π 4 5 β n R ) 1 / 3 , where R = 8.31 J mol−1K−1. The calculated Debye temperature ΘD is 309(4), 275(1) and 239(1) K for compounds with elemental Sn, Pb and Bi, respectively. The obtained values of the Sommerfeld coefficient and Debye temperature are significantly lower when compared with ReGa5 (γ = 4.68(7) mJ mol−1 K−2 and ΘD = 314(2)K). A small anomaly at around 2.2 K, is observed for each phase in the specific heat data measured under zero field, that resembles the critical superconducting temperature of ReGa5 (not shown here). This is consistent with the high concentration of ReGa5 present in each compound. Thus, we conclude that no phase transition was observed for ReGa~5(Sn), ReGa~5(Pb) and ReGa~5(Bi), indicating that the effect of the valence electron counts on the ReGa5 system resulted not only in a change of the crystal structure but also a loss of superconductivity. The whole temperature range Cp(T), see Figure 3b,d,f, shows a typical behavior and at high temperature, Cp approaches the expected Dulong–Petit value (3nR ≈ 150 J mol−1 K−1), where n is the number of atoms per formula unit (n = 6) and R is the gas constant (R = 8.31 J mol−1 K−1).
Electronic Structure. To evaluate the influence of the electronic structure on ReGa~5(Sn), ReGa~5(Pb) and ReGa~5(Bi), TB-LMTO-ASA calculations were performed to analyze the density of states (DOS), Crystal Orbital Hamiltonian Population (-COHP) curves and band structures. Due to the Ga vacancies and close proximity with regards to other atoms, a hypothetical “ReGa4.5” was utilized to calculate the electronic structure, these plots are shown in Figure 4a–c. Based on the stoichiometry of Sn, Pb and Bi determined from the SEM/EDS data, the electron counts of ReGa4.96Sn0.1, ReGa5.08Pb0.1 and ReGa5.13Bi0.2 were calculated to be 22.28, 22.64 and 23.39 valence electrons (VE) per Re, respectively. The Fermi energy level is indicated for each compound in Figure 4. As seen with ReGa5 and other endohedral gallide cluster superconductors, the Fermi energy level is located in a pseudo gap in the DOS, which is thought to play a role in the structural stability [24]. However, no von Hove singularities are found in the electronic structure of ReGa4.96(Sn), ReGa5.08(Pb) and ReGa5.13(Bi) around the Fermi level [25]. Consequently, this could be a key factor in the loss of superconductivity in these gallide clusters. This is consistent with the band structure calculation, Figure 4c, which indicates metallic behavior and shows no sign of flat bands near the Fermi energy. The DOS at the Fermi energy increases from ReGa5.08(Pb) to ReGa5.13(Bi) to ReGa4.96(Sn) suggesting the stability of the structures may come from the degree of Ga4 vacancies. -COHP curves, shown in Figure 4b, were calculated to analyze the interactions between Re and Ga in ReGa~5(Sn), ReGa~5(Pb) and ReGa~5(Bi). The -COHP shows the Fermi levels tend to move from strong antibonding interactions to the non-bonding interactions. This could be strongly related to the structural stability of these distorted phases and a major influence on the lack of superconductivity.

4. Conclusions

The understanding of Ga-rich TmGan endohedral clusters was further assessed through three new phases: ReGa~5(Sn); ReGa~5(Pb) and ReGa~5(Bi) presented here. The crystal structure characterization reveals that the new phases crystallize in a tetragonal unit cell resembling that of MnGa4.96. The electron-rich new Re-Ga phases produced edge-sharing polyhedra, further confirming the exo-bond formations between electron-rich and poor gallide clusters. Heat capacity measurements were performed, and no superconductivity was found down to 1.85 K, only a small anomaly associated with the ReGa5 superconductivity transition temperature. The Sommerfeld and Debye temperatures are significantly lower than ReGa5. Electronic structure calculations offered additional insight into the structural stability. Conclusively, the non-bonding character and absence of a pseudo gap and at the Fermi energy level along with the capped face sharing square antiprismatic Re@Ga8 polyhedra likely have a direct impact on the structural stability and loss of superconductivity in these Re-Ga clusters. These non-superconducting compounds may be useful for other potential applications.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/2073-4352/9/10/527/s1, Figure S1: Powder X-ray diffraction patterns, Figure S2: Heat capacity measurements, Table S1: Anisotropic displacement parameters, Table S2: Scanning electron microscopy.

Author Contributions

Conceptualization, M.M. and W.X.; methodology, M.M. K.G., R.S.D.M.; writing—original draft preparation, M.M.; writing—review and editing, W.X., K.G., R.S.D.M., T.K.; supervision, W.X.

Funding

M.M. and W.X. deeply appreciate the support from Beckman Young Investigator Program and the U.S. Department of Energy, Office of Science, Basic Energy Sciences, under EPSCoR Grant No. DE-SC0012432 with additional support from the Louisiana Board of Regents. Work at GUT was supported by National Science Centre (Poland), grant number: UMO-2018/30/M/ST5/00773.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Poole, C.K.; Farach, H.A.; Creswick, R.J. Handbook of Superconductivity; Elsevier: Amsterdam, The Netherlands, 1999. [Google Scholar]
  2. Henning, R.W.; Corbett, J.D. Formation of Isolated Nickel-Centered Gallium Clusters in Na10Ga10Ni and a 2-D Network of Gallium Octahedra in K2Ga3. Inorg. Chem. 1999, 38, 3883–3888. [Google Scholar] [CrossRef]
  3. Belin, C.; Tillard-Charbonnel, M. Frameworks of clusters in alkali metal-gallium phases: Structure, bonding and properties. Prog. Solid State Chem. 1993, 22, 59–109. [Google Scholar] [CrossRef]
  4. Henning, R.W.; Corbett, J.D. Cs8Ga11, a New Isolated Cluster in a Binary Gallium Compound. A Family of Valence Analogues A8Tr11X:  A = Cs, Rb; Tr = Ga, In, Tl; X. = Cl, Br, I. Inorg. Chem. 1997, 36, 6045–6049. [Google Scholar] [CrossRef] [PubMed]
  5. Kauzlarich, S.M.; Brown, S.R.; Snyder, G.J. Zintl phases for thermoelectric devices. Dalton Trans. 2007, 2099–2107. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Wade, K. Structural and Bonding Patterns in Cluster Chemistry. In Advances in Inorganic Chemistry and Radiochemistry; Emeléus, H.J., Sharpe, A.G., Eds.; Academic Press: Cambridge, MA, USA, 1976; pp. 1–66. [Google Scholar]
  7. Ellinger, F.H.; Zachariasen, W.H. The crystal structures of PuGa4 and PuGa6. Acta Crystallogr. 1965, 19, 281–283. [Google Scholar] [CrossRef]
  8. Curro, N.J.; Caldwell, T.; Bauer, E.D.; Morales, L.A.; Graf, M.J.; Bang, Y.; Balatsky, A.V.; Thompson, J.D.; Sarrao, J.L. Unconventional superconductivity in PuCoGa5. Nature 2005, 434, 622–625. [Google Scholar] [CrossRef] [PubMed]
  9. Neha, P.; Sivaprakash, P.; Ishigaki, K.; Kalaiselvan, G.; Manikandan, K.; Dhaka, R.S.; Uwatoko, Y.; Arumugam, S.; Patnaik, S. Nuanced superconductivity in endohedral gallide Mo8Ga41. Mater. Res. Express 2018, 6, 016002. [Google Scholar] [CrossRef]
  10. Tillard, M.; Belin, C. Investigation in the Ga-rich side of the Mn–Ga system: Synthesis and crystal structure of MnGa4 and MnGa5−x (x ~ 0.15). Intermetallics 2012, 29, 147–154. [Google Scholar] [CrossRef]
  11. Shibayama, T.; Nohara, M.; Aruga Katori, H.; Okamoto, Y.; Hiroi, Z.; Takagi, H. Superconductivity in Rh2Ga9 and Ir2Ga9 without Inversion Symmetry. J. Phys. Soc. Jpn. 2007, 76, 073708. [Google Scholar] [CrossRef]
  12. Belgacem-Bouzida, A.; Djaballah, Y.; Notin, M. Calorimetric measurement of the intermetallic compounds Cr3Ga and CrGa4 and thermodynamic assessment of the (Cr–Ga) system. J. Alloys Compd. 2005, 397, 155–160. [Google Scholar] [CrossRef]
  13. Yannello, V.J.; Kilduff, B.J.; Fredrickson, D.C. Isolobal Analogies in Intermetallics: The Reversed Approximation MO Approach and Applications to CrGa4- and Ir3Ge7-Type Phases. Inorg. Chem. 2014, 53, 2730–2741. [Google Scholar] [CrossRef] [PubMed]
  14. Xie, W.; Luo, H.; Phelan, B.F.; Klimczuk, T.; Cevallos, F.A.; Cava, R.J. Endohedral gallide cluster superconductors and superconductivity in ReGa5. Proc. Natl. Acad. Sci. USA 2015, 112, E7048–E7054. [Google Scholar] [CrossRef] [PubMed]
  15. Dinnebier, R.E.; Billinge, S.J.L. Chapter 1. Principles of Powder Diffraction. In Powder Diffraction; Dinnebier, R.E., Billinge, S.J.L., Eds.; Royal Society of Chemistry: Cambridge, MA, USA, 2008; pp. 1–19. [Google Scholar]
  16. Rodríguez-Carvajal, J. An Introduction to the Program FullProf 2000; Laboratoire Leon Brillouin (CEA-CNRS): Saclay, Paris, France, 2001. [Google Scholar]
  17. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. C Struct. Chem. 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed]
  18. Sheldrick, G.M. SHELXT—Integrated space-group and crystal-structure determination. Acta Crystallogr. A Found. Adv. 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed]
  19. Andersen, O.K.; Jepsen, O. Explicit, First-Principles Tight-Binding Theory. Phys. Rev. Lett. 1984, 53, 2571–2574. [Google Scholar] [CrossRef]
  20. Krier, G.; Jepsen, O.; Burkhardt, A.; Andersen, O. The TB-LMTO-ASA Program; Max-Planck-Institut fur Festkoorperforschung: Stuttgart, Germany, 1995. [Google Scholar]
  21. Deringer, V.L.; Tchougréeff, A.L.; Dronskowski, R. Crystal Orbital Hamilton Population (COHP) Analysis as Projected from Plane-Wave Basis Sets. J. Phys. Chem. A 2011, 115, 5461–5466. [Google Scholar] [CrossRef] [PubMed]
  22. Andersen, O.K. Linear methods in band theory. Phys. Rev. B 1975, 12, 3060–3083. [Google Scholar] [CrossRef] [Green Version]
  23. Lambrecht, W.R.L.; Andersen, O.K. Minimal basis sets in the linear muffin-tin orbital method: Application to the diamond-structure crystals C, Si, and Ge. Phys. Rev. B 1986, 34, 2439–2449. [Google Scholar] [CrossRef] [PubMed]
  24. Grin, Y.; Wedig, U.; Wagner, F.; von Schnering, H.G.; Savin, A. The analysis of “empty space” in the PdGa5 structure. J. Alloys Compd. 1997, 255, 203–208. [Google Scholar] [CrossRef]
  25. Srivichitranond, L.C.; Seibel, E.M.; Xie, W.; Sobczak, Z.; Klimczuk, T.; Cava, R.J. Superconductivity in a new intermetallic structure type based on endohedral Ta@Ir7Ge4 clusters. Phys. Rev. B 2017, 95, 174521. [Google Scholar] [CrossRef]
Figure 1. (a,b) Pictures of the ReGa~5(Sn) single crystals. (c) Powder XRD patterns for ReGa~5(Sn), ReGa~5(Pb) and ReGa~5(Bi).
Figure 1. (a,b) Pictures of the ReGa~5(Sn) single crystals. (c) Powder XRD patterns for ReGa~5(Sn), ReGa~5(Pb) and ReGa~5(Bi).
Crystals 09 00527 g001
Figure 2. Shown in this image is the structure of ReGa~5(Sn) (Re, purple; Ga, blue; Sn, grey) displaying a unit cell of the capped square antiprismatic Re@Ga8 polyhedra projection on the (a) bc plane and (b) ab plane. (c) Crystal structure of ReGa~5(Sn) showing the networks of capped Re2@Ga14 oblong chains. (d) Image comparing the capped Re2@Ga14 oblong chain of ReGa~5(Sn), ReGa~5(Pb) (Pb, purple) and ReGa~5(Bi) (Bi, orange).
Figure 2. Shown in this image is the structure of ReGa~5(Sn) (Re, purple; Ga, blue; Sn, grey) displaying a unit cell of the capped square antiprismatic Re@Ga8 polyhedra projection on the (a) bc plane and (b) ab plane. (c) Crystal structure of ReGa~5(Sn) showing the networks of capped Re2@Ga14 oblong chains. (d) Image comparing the capped Re2@Ga14 oblong chain of ReGa~5(Sn), ReGa~5(Pb) (Pb, purple) and ReGa~5(Bi) (Bi, orange).
Crystals 09 00527 g002
Figure 3. (a,c,e) Heat capacity (Cp(T)/T ~ T2) range 1.85 to 300 K for ReGa~5(Sn), ReGa~5(Pb) and ReGa~5(Bi) to obtain the Sommerfeld parameter (γ) and Debye temperature (θD). (b,d,f) Heat capacity (Cp(T) ~ T) range 1.85 to 300 K for ReGa~5(Sn), ReGa~5(Pb) and ReGa~5(Bi).
Figure 3. (a,c,e) Heat capacity (Cp(T)/T ~ T2) range 1.85 to 300 K for ReGa~5(Sn), ReGa~5(Pb) and ReGa~5(Bi) to obtain the Sommerfeld parameter (γ) and Debye temperature (θD). (b,d,f) Heat capacity (Cp(T) ~ T) range 1.85 to 300 K for ReGa~5(Sn), ReGa~5(Pb) and ReGa~5(Bi).
Crystals 09 00527 g003
Figure 4. Electronic structure calculations with a black solid line indicating the Fermi level for theoretical ReGa4.5 and a dashed blue, pink and green line indicating the Fermi energy level for ReGa4.96(Sn), ReGa5.08(Pb) and ReGa5.13(Bi), respectively, for the (a) density of states (DOS) curves, obtained from Local-Density Approximation (LDA), (b) band structure and (c) Orbital Hamiltonian Population Curves (COHP), for Re-Ga interactions.
Figure 4. Electronic structure calculations with a black solid line indicating the Fermi level for theoretical ReGa4.5 and a dashed blue, pink and green line indicating the Fermi energy level for ReGa4.96(Sn), ReGa5.08(Pb) and ReGa5.13(Bi), respectively, for the (a) density of states (DOS) curves, obtained from Local-Density Approximation (LDA), (b) band structure and (c) Orbital Hamiltonian Population Curves (COHP), for Re-Ga interactions.
Crystals 09 00527 g004
Table 1. Single crystal refinement data for ReGa4.96(Sn), ReGa5.08(Pb) and ReGa5.13(Bi).
Table 1. Single crystal refinement data for ReGa4.96(Sn), ReGa5.08(Pb) and ReGa5.13(Bi).
Refined Formula ReGa4.96(9)(Sn) ReGa5.08(5)(Pb)ReGa5.13(5)(Bi)
FW (g/mol) 532.19 540.55543.86
Space group; Z P 4/m n c; 4 P 4/m n c; 4 P 4/m n c; 4
a (Å) 6.4680(9) 6.4830(13) 6.474(2)
c (Å) 10.166(2) 10.229(2) 10.213(4)
V (Å3) 425.29(15) 429.9(2) 428.0(3)
Extinction Coefficient 0.0017(4) 0.0037(3) 0.0044(3)
θ range (°) 3.734–33.135 3.721–33.166 3.726–33.240
No. reflections; Rint 5007; 0.0646 5731; 0.0700 5798; 0.1039
No. independent reflections 429 436 438
No. parameters 30 29 30
R1; ωR2 (I > 2δ(I)) 0.0397; 0.0932 0.0284; 0.0552 0.0263; 0.0464
Goodness of fit 1.311 1.197 0.999
Diffraction peak and hole (e3) 2.713; −2.981 1.572; −2.3591.822; −1.863
Table 2. Atomic coordinates and equivalent isotropic displacement parameters for ReGa4.96(Sn), ReGa5.08(Pb) and ReGa5.13(Bi) (Ueq is defined as one-third of the trace of the orthogonalized Uij tensor (Å2)).
Table 2. Atomic coordinates and equivalent isotropic displacement parameters for ReGa4.96(Sn), ReGa5.08(Pb) and ReGa5.13(Bi) (Ueq is defined as one-third of the trace of the orthogonalized Uij tensor (Å2)).
AtomWyck.xyzOcc.Ueq
ReGa4.96(9)(Sn)
Re4e000.3451(1)10.008(1)
Ga28h0.4535(3)0.1808(3)010.016(1)
Ga38g0.7431(3)0.2431(3)¼10.045(1)
Ga44e000.051(3)0.18(3)0.021(8)
Ga516i0.0426(18)0.096(2)0.0947(8)0.20(1)0.028(4)
ReGa5.08(5)(Pb)
Re4e000.3453(1)10.007(1)
Ga28h0.4622(2)0.1816(2)010.016(1)
Ga38g0.7462(2)0.2462(2)¼10.040(1)
Ga42a0000.75(1)0.022(1)
Ga516i0.0389(11)0.0883(19)0.0970(5)0.18(1)0.024(2)
ReGa5.13(5)(Bi)
Re4e000.3451(1)10.006(1)
Ga28h0.4692(1)0.1821(1)010.018(1)
Ga38g0.7467(1)0.2467(1)¼10.033(1)
Ga44e000.0318(3)0.62(1)0.012(1)
Ga516i0.039(2)0.079(5)0.1001(6)0.13(1)0.027(3)
Table 3. List of atomic distances for ReGa4.96(Sn), ReGa5.08(Pb) and ReGa5.13(Bi).
Table 3. List of atomic distances for ReGa4.96(Sn), ReGa5.08(Pb) and ReGa5.13(Bi).
ReGa4.96(Sn)
Atom1Atom2Distances (Å)
ReGa22.614(2)
ReGa32.484(2)
ReGa42.99(4)
ReGa52.635(9)
Ga5Ga51.36(3) (diagonally along the b axis)
ReGa5.08(Pb)
Atom1Atom2Distances (Å)
ReGa22.612(1)
ReGa32.491(1)
ReGa43.5322(6)
ReGa5 2.616(6)
Ga5Ga51.25(1) (diagonally along the b axis)
ReGa5.13(Bi)
Atom1Atom2Distances (Å)
ReGa22.604(1)
ReGa32.487(1)
ReGa43.200(4)
ReGa52.566(8)
Ga5Ga51.14(4) (diagonally along the b axis)

Share and Cite

MDPI and ACS Style

Marshall, M.; Górnicka, K.; Mudiyanselage, R.S.D.; Klimczuk, T.; Xie, W. New Tetragonal ReGa5(M) (M = Sn, Pb, Bi) Single Crystals Grown from Delicate Electrons Changing. Crystals 2019, 9, 527. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst9100527

AMA Style

Marshall M, Górnicka K, Mudiyanselage RSD, Klimczuk T, Xie W. New Tetragonal ReGa5(M) (M = Sn, Pb, Bi) Single Crystals Grown from Delicate Electrons Changing. Crystals. 2019; 9(10):527. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst9100527

Chicago/Turabian Style

Marshall, Madalynn, Karolina Górnicka, Ranuri S. Dissanayaka Mudiyanselage, Tomasz Klimczuk, and Weiwei Xie. 2019. "New Tetragonal ReGa5(M) (M = Sn, Pb, Bi) Single Crystals Grown from Delicate Electrons Changing" Crystals 9, no. 10: 527. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst9100527

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