Next Article in Journal
Acoustic Performance of Resilient Materials Using Acrylic Polymer Emulsion Resin
Next Article in Special Issue
Effect of Transition Metal Substitution on the Structure and Properties of a Clathrate-Like Compound Eu7Cu44As23
Previous Article in Journal
Solvent-Induced Polymorphism of Iron(II) Spin Crossover Complexes
Previous Article in Special Issue
Synthesis and Structural Characterization of the New Clathrates K8Cd4Ge42, Rb8Cd4Ge42, and Cs8Cd4Ge42
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Binary Alkali-Metal Silicon Clathrates by Spark Plasma Sintering: Preparation and Characterization

1
Max-Planck-Institut für Chemische Physik fester Stoffe, 01187 Dresden, Germany
2
Department of Physics, California Polytechnic State University, San Luis Obispo, CA 93407, USA
3
Department of Physics, University of South Florida, Tampa, FL 33620, USA
*
Authors to whom correspondence should be addressed.
Materials 2016, 9(7), 593; https://doi.org/10.3390/ma9070593
Submission received: 13 June 2016 / Revised: 12 July 2016 / Accepted: 13 July 2016 / Published: 19 July 2016
(This article belongs to the Special Issue Inorganic Clathrate Materials)

Abstract

:
The binary intermetallic clathrates K8-xSi46 (x = 0.4; 1.2), Rb6.2Si46, Rb11.5Si136 and Cs7.8Si136 were prepared from M4Si4 (M = K, Rb, Cs) precursors by spark-plasma route (SPS) and structurally characterized by Rietveld refinement of PXRD data. The clathrate-II phase Rb11.5Si136 was synthesized for the first time. Partial crystallographic site occupancy of the alkali metals, particularly for the smaller Si20 dodecahedra, was found in all compounds. SPS preparation of Na24Si136 with different SPS current polarities and tooling were performed in order to investigate the role of the electric field on clathrate formation. The electrical and thermal transport properties of K7.6Si46 and K6.8Si46 in the temperature range 4–700 K were investigated. Our findings demonstrate that SPS is a novel tool for the synthesis of intermetallic clathrate phases that are not easily accessible by conventional synthesis techniques.

Graphical Abstract

1. Introduction

Binary phases of alkali metals with silicon have been prepared by thermal decomposition of the monosilicides M4Si4 (M = Na, K, Rb, Cs) for over six decades [1,2,3,4]. Structural investigations of the polycrystalline products revealed the existence of clathrate-I M8-xSi46 and clathrate-II M24-xSi136 silicides [5,6]. In both clathrate types, the alkali-metal atoms are enclosed in polyhedral cages formed in the tetrahedrally bonded silicon framework. In the clathrate-I crystal structure, the unit cell contains two Si20 and six Si24 cages, while in the clathrate-II structure there are sixteen Si20 and eight Si28 cages present in the unit cell. In the clathrate phases prepared by thermal decomposition the filling fraction of M in the silicon cages depends explicitly on both the reaction conditions and the size of the M atom relative to the size of the cage [6] (Figure 1).
Alternatively, redox techniques have been developed to obtain metastable clathrate phases [7,8,9,10] including Ge and Si allotropes with empty cages [11,12]. Defect-free single crystals of alkali metal silicon clathrates large enough for structure investigations have only been obtained in exceptional cases [4]. For example, crystals of K8Si46 were grown on α-Si exposed to potassium vapor [13], while crystals of ternary clathrate-II phases Rb8Na16Si136 and Cs8Na16Si136 form in closed ampoules directly from the melt [14,15]. Single-crystals suitable for measurement of intrinsic transport properties were first prepared by the electrochemically-driven transformation of Na4Si4 to Na24Si136 during the spark plasma treatment (SPS) [16,17]. Another approach—the kinetically controlled thermal decomposition (KCTD) method—can also provide relatively large single crystals [18,19]. Recently, a thermal decomposition at lower temperatures was employed for the preparation of Na24-x(SiyGe1-y)136 and proved to be useful for the synthesis of silicon clathrates [20]. Extensions of the SPS and KCTD approaches have also recently reported, including use of multi-phase precursors or simultaneous ion-exchange and electrochemical reactions to prepare ternary or quaternary clathrates [21,22,23].
The unique advantage of SPS for the synthesis of clathrates by producing scalable, compact bulk materials also continues to be of great interest [4]. In this work we discuss both the preparation and consolidation of alkali-metal silicon clathrates by SPS. We report on the preparation and crystal structure of different clathrate phases (Table 1) as well as SPS densification of K7.6Si46 and K6.8Si46 powders by SPS that allow for the investigation of their transport properties over a large temperature range. The formation and evaporation the free alkali metals during the reaction allow us to work only with a small amount (less than 200 mg) of precursors and, in the results, after processing, we obtained 10–20 mg of the materials.

2. Results and Discussion

As first described in reference [16], clathrates are formed during SPS redox-treatment by oxidation of Si44− at the anode while the alkali metal is reduced at the cathode (Figure 2a,b). Clathrate-II formation, according to
34 M4Si4M24Si136 + 112 M
can therefore be described by the half-reactions
34 Si4 4− − 112 e → Si136 24− (anode)
112 M+ + 112 e → 112 M0 (cathode)
For clathrate-I formation according to
23 M4Si4 → 2 M8Si46 + 76 M
should be considered by means of the half-reactions
23 Si4 4− – 76 e → 2 Si46 8− (anode)
76 M+ + 76 e → 76 M0 (cathode)
Reactions (1) and (2) describe the basics of the clathrate formation by SPS, whereby the process is driven by the DC field and the evaporation of the alkali metal at the cathode and its absorption by the graphite die (Figure 2b). In order to qualitatively investigate the influence of electric current on clathrate formation we performed three control experiments with Na4Si4 as the precursor. First, the polarity of the DC current was reversed, upon which it was found that the clathrate phase, Na24Si136 in this case, consistently forms at the anode (Figure 2c). In a second experiment, equal volumes of Al2O3 powder were placed above and below the precursor to create an electrically insulating barrier. The electrical current then presumably flowed through the low resistance graphite die such that the current passing through the Na4Si4 precursor is minimized. In this case no reaction was observed under the identical conditions used to form the clathrate Na24Si136 (Figure 2d). In the third experiment, the use of stainless steel punches combined with a die manufactured from boron nitride, a high electrical-resistivity material, presumably forced most of the current to pass through the precursor (as opposed to the surrounding die) (Figure 2e). A more rapid reaction than in the case of using the graphite die was observed. Optical photographs of cross-section of SPS-processed pellets (Figure 2f,h) clearly show that formation of the clathrate phases occurred at the anode during SPS processing. This was also the case for all clathrate compositions prepared in this study.
The crystal structures of the SPS-prepared clathrate phases were investigated by the Rietveld method using powder X-ray diffraction (PXRD) data. Experimental, calculated, and difference diffraction patterns for the clathrate-I and II samples are shown on Figure 3 and Figure 4, respectively, and the resulting crystallographic data are presented in Table 2 and Table 3, respectively.
The lattice parameters for K7.6Si46, Rb6.2Si46, and Cs7.8Si136 show good agreement with those synthesized by other preparation methods [7,16,18,19,24,25,26]. The lattice parameter for Rb11.5Si136, as expected, has an intermediate value between that reported for K17.2Si136 [19] and Cs8Si136 [6]. In all cases the silicon framework sites were found to be fully occupied (no vacancies), in agreement with prior results for binary silicon clathrates [7,16,18,19,26,27,28]. The stable covalent Si–Si bonds in silicon clathrates allow for a large excess of electrons that fill antibonding states. In a rigid-band model, a high concentration of conduction electrons are provided by charge transfer from the alkali-metal guests, thus inducing metallic behavior, as confirmed experimentally [15,29]. In contrast to SPS-prepared Na8Si46 and Na24Si136, which show full occupancy of the polyhedral cages by sodium [19], the SPS-prepared clathrates of potassium, rubidium and caesium exhibit partial occupancy of the alkali-atom positions (the specific refined compositions are given in Table 2 and Table 3).
The compositions of clathrates-II prepared in this study together with previously reported in the literature [16,19,29,30] allow for a comparison of the lattice parameter versus guest-atom content in clathrate-II phases. As shown in Figure 5, in all cases the lattice parameter increases with those increasing M content in the Si20 (crystallographic site 16c) cage. In addition, the lattice parameter also increases with increasing alkali-ion radius [31] for a specific M concentration. Preferential occupancy of the larger Si28 (crystallographic site 8b) cages was observed for the case of Rb11.5Si136, as well as for for NaxSi136 [30,32].
In order to further characterize the products formed by SPS we measured the electrical and thermal transport properties of K7.6Si46 and K6.8Si46. The density of the pellets (only K8-xSi46) was approximately 90% of the calculated density based on PXRD data. Temperature-dependent electrical resistivity ρ (Figure 6, top), Seebeck coefficient S (Figure 6, middle), and thermal conductivity κ (Figure 6, bottom) in the temperature range from 10 to 700 K are shown in Figure 6. The agreement between the high- and low-temperature data, measured with two different setups on two different specimens, is an indication of the homogeneity of the materials prepared by SPS. Both materials K7.6Si46 and K6.8Si46 show metallic temperature dependence of ρ and S, in agreement with the seminal work by Cros et al. [6]. The fact that these materials show metallic conduction follows from the simple assertion that each K atom donates one electron to the silicon framework resulting in an overall high electron concentration. Measured ρ of cold-pressed K7.6Si46, obtained from the thermal decomposition of K4Si4, was reported to be much higher: ρ ~ 140 mΩ·cm at room temperature with weak temperature dependence [28]. The large discrepancy between those results and the data shown in Figure 6 may be attributed to the microstructure of the specimens and the typically poor inter-granular contact that is achieved using cold pressing [28]. Nevertheless, the ρ values shown in Figure 6 are significantly larger than expected for the presumed high electron concentration. In addition, the residual resistance ratio, ρ300K/ρ10K = 3.6, is also relatively low for a good metal and significantly smaller than the values obtained for single crystals of Na8Si46 (ρ300K/ρ10K = 36) [33] and Na24Si136 (ρ300K/ρ10K = 14) [17]. These results reiterate the need for high-density polycrystalline specimens in order to minimize the effects of grain boundary scattering on the transport properties of intermetallic clathrates [34].

3. Materials and Methods

3.1. Synthesis of Precursors

Binary precursors M4Si4 (M = Na, K, Rb, or Cs) were prepared from alkali metal pieces (ChemPur, Karlsruhe, Germany 99.9 wt %) and coarsely ground silicon (ChemPur, 99.999 wt %), by sealing of the precursor mixture under argon in small, welded tantalum ampoules that were enclosed under vacuum inside silica tubes and heated to 800 °C at an average rate of 3 °C/min. The ampoules were held at 800 °C for 4 h before being slowly cooled to room temperature (furnace cooled). The reaction products were determined to be single phase M4Si4 by powder X-ray diffraction (PXRD). All sample manipulations were performed in an argon-filled glove box (H2O < 1 ppm; O2 < 1 ppm). CAUTION: M4Si4 compounds are sensitive to air and moisture (pyrophoric) and must be kept under inert atmosphere at all times.

3.2. Spark Plasma Synthesis

SPS experiments were conducted under vacuum (background pressure <10 Pa) using a dedicated SPS setup in a protective argon atmosphere (Figure 2a; SPS: 515 ET, Syntex-Fuji, Japan; glove box: MBraun, Garching, Germany) [24]. Table 1 contains the preparation conditions for the clathrates synthesized within this study. The precursors were ground to fine powder and loaded into graphite or boron nitride (h-BN) dies with an inner diameter of 8 mm. Graphite or stainless steel punches were used to apply the uniaxial pressure. Tantalum foil was employed to surround the powder specimen on all sides, isolating the specimen from the die and punches during SPS processing. Graphite dies worked well for Na, however they were observed to react with alkali-metal vapor for precursors of K, Rb, and Cs and often cracked the graphite die during SPS processing. For these alkali metals h-BN was found to be most suitable. After SPS the products were isolated from the residual precursor by washing first with ethanol and then distilled water under flowing argon.

3.3. Sample Characterization

Powder X-ray diffraction (PXRD) data were collected using the Guinier technique (Huber G670 camera, Rimsting, Germany, Cu-Kα1 radiation, λ = 1.540598 Å, graphite monochromator, 5° ≤ 2θ ≤ 100°, ∆2θ = 0.005°). Crystal structure refinements were performed using the software WinCSD [25]. For the transport measurements, the K7.6Si46 and K6.8Si46 powders were first washed then consolidated by SPS using tungsten carbide tool (8 mm inner diameter) at 300 °C and a uniaxial pressure of 600 MPa under vacuum (background pressure of approximately 10 Pa) for 10 min. The density of the pellets was approximately 90% of the calculated density based on PXRD data. The specimens were cut with a diamond wire saw into parallelepiped pieces 1.5 mm × 1.5 mm × 8.0 mm for electrical transport measurements and low-temperature thermal conductivity measurements, and 1.5 mm thick and 8 mm in diameter disks for thermal diffusivity measurements. Electrical resistivity (ρ) and Seebeck coefficient (S) measurements from 22 to 450 °C on SPS densified K7.6Si46 and K6.8Si46 specimens were performed with a commercial ZEM-3 setup (ULVAC-RIKO, Munich, Germany). Specific heat (Cp) measurements were performed on a Netzsch Pegasus DSC (Selb, Germany) while heating to 450 °C at a heating rate of 10 °C/min in an Ar atmosphere. Thermal diffusivity, D, was measured from 22 °C to 450 °C using the laser-flash technique (LFA 457 Micro Flash, Netzsch, Selb, Germany). The temperature dependent thermal conductivity, κ, was calculated using the relation κ = Cp × D × d where d is the density in g/cm3. Low temperature ρ, S, and κ data were collected using the TTO-option (Thermal Transport Option) of a PPMS (Physical Property Measurement System) (Quantum Design, San Diego, CA, USA) using a four-point method.

4. Conclusions

Binary intermetallic clathrates K8-xSi46 (x = 0.4; 1.2), Rb6.2Si46, Rb11.5Si46, and Cs7.8Si136 were synthesized by spark plasma synthesis from the respective M4Si4 precursors. The formation of polycrystalline products at the anode of the SPS setup was observed for all compounds. The preparation of the new clathrate-II phase Rb11.5Si136 is further evidence that SPS is useful for the inorganic synthesis. The role the electric field is in driving the chemical conversion of the precursor into the clathrate phase at the anode during SPS processing is emphasized. Conventional thermal decomposition (CTC) [6,28,35], kinetically controlled thermal decomposition (KCTD) [18,19], chemical oxidation (CO) [7,27], and SPS yield different intermetallic clathrate phases from the same M4Si4 precursor (Table 4). While the influence of temperature clearly plays a role in clathrate formation, a definitive understanding of the local formation mechanisms in each of these synthetic approaches [36] is of great interest and is an ongoing task.

Acknowledgments

M.Be. (while at USF) and G.S.N. acknowledge support from the US Department of Energy, Basic Energy Sciences, Division of Materials Science and Engineering, under Award No. DE-FG02-04ER46/45 for SPS synthesis. M.Be. also acknowledges support from Deutscher Akademischer Austausch Dienst. M.Ba. and Y.G. acknowledge support from DFG priority program SPP1415.

Author Contributions

Y.G. and G.S.N. developed the concept of this study; I.V. and M.Be. conceived and designed the SPS experiments; I.A. and W.S. performed the physical measurements experiments; I.V., M.Be., and M.Ba. refined the crystal structures from PXRD experiments; all authors contributed to the text of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hohmann, E. Silicide und Germanide der Alkalimetalle. Z. Anorg. Chem. 1948, 257, 113–126. [Google Scholar] [CrossRef]
  2. Cros, C.; Pouchard, M.; Hagenmuller, P. Two new phases of the silicon-sodium system. C. R. Hebd. Seances Acad. Sci. 1965, 260, 4764–4767. [Google Scholar]
  3. Cros, S.; Pouchard, M. Sur les phases de type clathrate du silicium et des éléments apparentés (C, Ge, Sn): Une approche historique. C. R. Chim. 2009, 12, 1014–1056. [Google Scholar] [CrossRef]
  4. Nolas, G.S. (Ed.) The Physics and Chemistry of Inorganic Clathrates; Springer: Berlin, Germany, 2014.
  5. Kasper, J.S.; Hagenmuller, P.; Pouchard, M.; Cros, C. Clathrate Structure of Silicon Na8Si46 and NaxSi136 (x < 11). Science 1965, 150, 1713–1714. [Google Scholar] [PubMed]
  6. Cros, C.; Pouchard, M.; Hagenmuller, P. Sur une nouvelle famille de clathrates minéraux isotypes des hydrates de gaz et de liquides. Interprétation des résultats obtenus. J. Solid State Chem. 1970, 2, 570–581. [Google Scholar] [CrossRef]
  7. Böhme, B.; Guloy, A.; Tang, Z.; Schnelle, W.; Burkhardt, U.; Baitinger, M.; Grin, Y. Oxidation of M4Si4 (M = Na, K) to Clathrates by HCl or H2O. J. Am. Chem. Soc. 2007, 129, 5348–5349. [Google Scholar] [CrossRef] [PubMed]
  8. Böhme, B.; Hoffmann, S.; Baitinger, M.; Grin, Y. Application of n-Dodecyltrimethylammonium Chloride for the Oxidation of Intermetallic Phases. Z. Naturforsch. 2011, 66, 230–238. [Google Scholar] [CrossRef]
  9. Liang, Y.; Böhme, B.; Reibold, M.; Schnelle, W.; Schwarz, U.; Baitinger, M.; Lichte, H.; Grin, Y. Synthesis of the Clathrate-I Phase Ba8-xSi46 via Redox Reactions. Inorg. Chem. 2011, 50, 4523–4528. [Google Scholar] [CrossRef] [PubMed]
  10. Krishna, L.; Chai, P.; Koh, C.A.; Toberer, E.S.; Nolas, G.S. Synthesis and structural properties of type I potassium SiGe alloy clathrates. Mater. Lett. 2015, 149, 123–126. [Google Scholar] [CrossRef]
  11. Ammar, A.; Cros, C.; Pouchard, M.; Jaussaud, N.; Bassat, J.-M.; Villeneuve, G.; Duttine, M.; Ménétrier, M.; Reny, E. On the clathrate form of elemental silicon, Si136: Preparation and characterisation of NaxSi136 (x → 0). Solid State Sci. 2004, 6, 393–400. [Google Scholar] [CrossRef]
  12. Guloy, A.M.; Ramlau, R.; Tang, Z.; Schnelle, W.; Baitinger, M.; Grin, Y. A guest-free germanium clathrate. Nature 2006, 443, 320–323. [Google Scholar] [CrossRef] [PubMed]
  13. Gallmeier, J.; Schäfer, H.; Weiss, A. Cage structure as a common building principle of compounds K8E46 (E = silicon, germanium, tin). Z. Naturforsch. B 1969, 24, 665–667. [Google Scholar]
  14. Bobev, S.; Sevov, S.C. Synthesis and Characterization of Stable Stoichiometric Clathrates of Silicon and Germanium: Cs8Na16Si136 and Cs8Na16Ge136. J. Am. Chem. Soc. 1999, 121, 3795–3796. [Google Scholar] [CrossRef]
  15. Nolas, G.S.; Vanderveer, D.G.; Wilkinson, A.P.; Cohn, J.L. Temperature dependent structural and transport properties of the type II clathrates A8Na16E136 (A = Cs or Rb and E = Ge or Si). J. Appl. Phys. 2002, 91, 8970–8973. [Google Scholar] [CrossRef]
  16. Beekman, M.; Baitinger, M.; Borrmann, H.; Schnelle, W.; Meier, K.; Nolas, G.S.; Grin, Y. Preparation and Crystal Growth of Na24Si136. J. Am. Chem. Soc. 2009, 131, 9642–9643. [Google Scholar] [CrossRef] [PubMed]
  17. Beekman, M.; Schnelle, W.; Borrmann, H.; Baitinger, M.; Grin, Y.; Nolas, G.S. Intrinsic Electrical and Thermal Properties from Single Crystals of Na24Si136. Phys. Rev. Lett. 2010, 104. [Google Scholar] [CrossRef] [PubMed]
  18. Stefanoski, S.; Beekman, M.; Wong-Ng, W.; Zavalij, P.; Nolas, G.S. Simple Approach for Selective Crystal Growth of Intermetallic Clathrates. Chem. Mater. 2011, 23, 1491–1495. [Google Scholar] [CrossRef]
  19. Stefanoski, S.; Nolas, G.S. Synthesis and Structural Characterization of Single-Crystal K7.5Si46 and K17.8Si136 Clathrates. Cryst. Growth Des. 2011, 11, 4533–4537. [Google Scholar] [CrossRef]
  20. Baranowski, L.L.; Krishna, L.; Martinez, A.D.; Raharjo, T.; Stevanovic, V.; Tamboli, A.C.; Toberer, E.S. Synthesis and optical band gaps of alloyed Si–Ge type II clathrates. J. Mater. Chem. B 2014, 2, 3231–3237. [Google Scholar] [CrossRef]
  21. Dong, Y.; Nolas, G.S. Rapid crystal growth of type-II clathrates A8Na16Si136 (A = K, Rb, Cs) by spark plasma sintering. Cryst. Eng. Commun. 2015, 17, 2242–2244. [Google Scholar] [CrossRef]
  22. Dong, Y.; Chai, P.; Beekman, M.; Zeng, X.; Tritt, T.M.; Nolas, G.S. Precursor Routes to Complex Ternary Intermetallics: Single-Crystal and Microcrystalline Preparation of Clathrate-I Na8Al8Si38 from NaSi + NaAlSi. Inorg. Chem. 2015, 54, 5316–5321. [Google Scholar] [CrossRef] [PubMed]
  23. Wei, K.; Dong, Y.; Nolas, G.S. Precursor routes to quaternary intermetallics: Synthesis, crystal structure, and physical properties of clathrate-II Cs8Na16Al24Si112. J. Solid State Chem. 2016, 237, 81–85. [Google Scholar] [CrossRef]
  24. Reinfried, N.; Höhn, P.; Grin, Y. Spark-Plasma Synthesis an Inert Gas Atmosphere; Max-Planck-Institute for Chemical Physics of Solids: Dresden, Germany, 2006; p. 29. [Google Scholar]
  25. Akselrud, L.; Grin, Y. WinCSD: Software package for crystallographic calculations (Version 4). J. Appl. Crystallogr. 2014, 47, 803–805. [Google Scholar] [CrossRef]
  26. Stefanoski, S.; Blosser, M.C.; Nolas, G.S. Pressure Effects on the Size of Type-I and Type-II Si-Clathrates Synthesized by Spark Plasma Sintering. Cryst. Growth Des. 2013, 13, 195–197. [Google Scholar] [CrossRef]
  27. Blosser, M.; Nolas, G.S. Synthesis of Na8Si46 and Na24Si136 by oxidation of Na4Si4 from ionic liquid decomposition. Mater. Lett. 2013, 99, 161–163. [Google Scholar] [CrossRef]
  28. Ramachandran, G.K.; McMillan, P.F.; Dong, J.; Sankey, O.F. K7.62(1)Si46 and Rb6.15(2)Si46: Two Structure I Clathrates with Fully Occupied Framework Sites. J. Solid State Chem. 2000, 154, 626–634. [Google Scholar] [CrossRef]
  29. Bobev, S.; Sevov, S.C. Clathrates of Group 14 with Alkali Metals: An Exploration. J. Solid State Chem. 2000, 153, 92–105. [Google Scholar] [CrossRef]
  30. Beekman, M.; Nenghabi, E.N.; Biswas, K.; Myles, C.W.; Baitinger, M.; Grin, Y.; Nolas, G.S. Framework Contraction in Na-Stuffed Si(cF136). Inorg. Chem. 2010, 49, 5338–5340. [Google Scholar] [CrossRef] [PubMed]
  31. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr. A 1976, 32, 751–767. [Google Scholar] [CrossRef]
  32. Reny, E.; Gravereau, P.; Cros, C.; Pouchard, M. Structural characterisations of the NaxSi136 and Na8Si46 silicon clathrates using the Rietveld method. J. Mater. Chem. 1998, 8, 2839–2844. [Google Scholar] [CrossRef]
  33. Stefanoski, S.; Martin, J.; Nolas, G.S. Low temperature transport properties and heat capacity of single-crystal Na8Si46. J. Phys. Condens. Matter 2010, 22. [Google Scholar] [CrossRef] [PubMed]
  34. Beekman, M.; Nolas, G.S. Transport Properties of the Binary Type I Clathrate K8Ge442. Int. J. Appl. Ceram. Technol. 2007, 4, 332–338. [Google Scholar] [CrossRef]
  35. Cros, C.; Benejat, J.-C. Preparation and properties of a clathrate having an extensive domain of existence. Sodium silicide NaxSi136 compounds. Bull. Soc. Chim. Fr. 1972, 5, 1739–1743. [Google Scholar]
  36. Hutchins, P.T.; Leynaud, O.; O’Dell, L.A.; Smith, M.E.; Barnes, P.; McMillan, P.F. Time-Resolved in Situ Synchrotron X-ray Diffraction Studies of Type 1 Silicon Clathrate Formation. Chem. Mater. 2011, 23, 5160–5167. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Crystal structures of (a) Na4Si4; (b) M4Si4 (M = K, Rb, or Cs); (c) M8-xSi46; and (d) M24-xSi136.
Figure 1. Crystal structures of (a) Na4Si4; (b) M4Si4 (M = K, Rb, or Cs); (c) M8-xSi46; and (d) M24-xSi136.
Materials 09 00593 g001
Figure 2. (a) SPS setup for the preparation of clathrates; (be) Location and relative amounts of the reaction products for different polarities and different die and punch materials; (fh) Optical micrographs of the cross-sections from fractured post-reaction pellets illustrating the directional growth of clathrates at the anode.
Figure 2. (a) SPS setup for the preparation of clathrates; (be) Location and relative amounts of the reaction products for different polarities and different die and punch materials; (fh) Optical micrographs of the cross-sections from fractured post-reaction pellets illustrating the directional growth of clathrates at the anode.
Materials 09 00593 g002
Figure 3. Powder X-ray diffraction (PXRD) patterns of the SPS-prepared clathrates-I K7.6Si46 (550 °C) and K6.8Si46 (600 °C). Red asterisks mark the reflections of α-Si (<0.2 mass % in K7.6Si46 and 1.5 mass % in K6.8Si46). The experimental intensities are shown as a solid line. The difference between the experimental and calculated intensities is shown below the experimental data; the tick marks represent the positions of the diffraction reflections of the clathrate-I phase.
Figure 3. Powder X-ray diffraction (PXRD) patterns of the SPS-prepared clathrates-I K7.6Si46 (550 °C) and K6.8Si46 (600 °C). Red asterisks mark the reflections of α-Si (<0.2 mass % in K7.6Si46 and 1.5 mass % in K6.8Si46). The experimental intensities are shown as a solid line. The difference between the experimental and calculated intensities is shown below the experimental data; the tick marks represent the positions of the diffraction reflections of the clathrate-I phase.
Materials 09 00593 g003
Figure 4. Powder X-ray diffraction (PXRD) patterns of the Rb6.2Si46 and Rb11.1Si136 mixed product and clathrate-II Cs7.8Si136. The experimental intensities are shown as a solid line. The differences between the experimental and calculated intensities are shown below the experimental data. The tick marks represent the positions of the diffraction reflections. Red asterisks mark the reflections from α-Si (<0.2 mass % in Rb6.2Si46 and Rb11.1Si136 and 2.5 mass % in Cs7.8Si136).
Figure 4. Powder X-ray diffraction (PXRD) patterns of the Rb6.2Si46 and Rb11.1Si136 mixed product and clathrate-II Cs7.8Si136. The experimental intensities are shown as a solid line. The differences between the experimental and calculated intensities are shown below the experimental data. The tick marks represent the positions of the diffraction reflections. Red asterisks mark the reflections from α-Si (<0.2 mass % in Rb6.2Si46 and Rb11.1Si136 and 2.5 mass % in Cs7.8Si136).
Materials 09 00593 g004
Figure 5. Lattice parameters versus M content, x for clathrate-I NaxSi136 (0 < x < 24) [28], Na24Si136 [16], K17.2Si136 [19], Rb11.5Si136 (this work), and Cs7.8Si136 (this work), as well as ternary Rb8Na16Si136 and Cs8Na16Si136 [27].
Figure 5. Lattice parameters versus M content, x for clathrate-I NaxSi136 (0 < x < 24) [28], Na24Si136 [16], K17.2Si136 [19], Rb11.5Si136 (this work), and Cs7.8Si136 (this work), as well as ternary Rb8Na16Si136 and Cs8Na16Si136 [27].
Materials 09 00593 g005
Figure 6. Transport properties of polycrystalline K7.6Si46 and K6.8Si46. The low temperature, 2 to 300 K, and high temperature, 300 to 700 K, are described in the text. (Top) electrical resistivity ρ; (Middle) Seebeck coefficient S; (Bottom) thermal conductivity κ.
Figure 6. Transport properties of polycrystalline K7.6Si46 and K6.8Si46. The low temperature, 2 to 300 K, and high temperature, 300 to 700 K, are described in the text. (Top) electrical resistivity ρ; (Middle) Seebeck coefficient S; (Bottom) thermal conductivity κ.
Materials 09 00593 g006aMaterials 09 00593 g006b
Table 1. Spark-plasma synthesis (SPS) conditions for the silicon clathrates.
Table 1. Spark-plasma synthesis (SPS) conditions for the silicon clathrates.
PrecursorTR (°C) atR (min) bDie/Punches cProducts d
Na4Si4550180C/CNa24Si136 (M) + Na8Si46 (m)
600180C/CNa24Si136 (M)
700180C/CNa24Si136 (M) + Si (m)
K4Si450060BN/SSK8-xSi46 (tr)
55060BN/SSK8-xSi46 (M)
60060BN/SSK8-xSi46 (M) + Si (m)
65060BN/SSSi (M) + K8-xSi46 (m)
Rb4Si445060BN/SSRb8-xSi46 (M) + Rb24-xSi136 (m) + Si (tr)
50060BN/SSSi (M) + Rb8-xSi46 (m) + Rb24-xSi136 (m)
Cs4Si435060BN/SSCs8-xSi136 (M) + Si (tr)
a TR is the reaction temperature; b tR is the treatment time; c Materials for the die and punches: C = graphite, BN = hp-boron nitride, SS = stainless steel; d Excluding unreacted precursor. M = majority phase, m = minority phase, tr = trace.
Table 2. Crystallographic information on the clathrate-I phases prepared by SPS (PXRD data).
Table 2. Crystallographic information on the clathrate-I phases prepared by SPS (PXRD data).
PhaseK7.6(1)Si46K6.8(1)Si46Rb6.2(1)Si46
Lattice parameter a, Å10.2776 (1)10.2747 (1)10.2848 (1)
Radiation, wavelength λ, ÅCu Kα1, 1.54056
Maximal diffraction angle 2θ, °100.30
Residuals RI/RP0.04/0.100.04/0.090.02/0.10
M1 2a (0 0 0)Occ * = 0.783 (1)Occ = 0.695 (4)Occ = 0.244 (3)
M2 6d (¼ ½ 0)Occ = 1.013 (3) **Occ = 0.918 (3)Occ = 0.961 (2)
Si1 6c (¼ 0 ½)
Si2 16i (x x x)x = 0.1845 (1)x = 0.1846 (1)x = 0.1840 (1)
Si3 24k (0 y z)y = 0.3068 (1)y = 0.3060 (1)y = 0.3043 (1)
z = 0.1183 (1)z = 0.1185 (1)z = 0.1190 (1)
*: Occ—occupancy factor; **: The value 1.013 (3) for the K2 occupancy in K7.6Si46 reflects the real error of the refinement (≌4 e.s.d.) and helps to understand the reliability of the occupancy values for other clathrates.
Table 3. Crystallographic information on the clathrate-II phases prepared by SPS (PXRD data).
Table 3. Crystallographic information on the clathrate-II phases prepared by SPS (PXRD data).
PhaseRb11.1(1)Si136Cs7.8(1)Si136
Lattice parameter a, Å14.7142 (9)14.6733 (3)
Radiation, wavelength λ, ÅCu Kα1, 1.54056
Maximal diffraction angle 2θ, °100.30
Residuals RI/RP0.10/0.100.05/0.13
M1 8b (⅜ ⅜ ⅜)Occ * = 0.85 (2)Occ = 0.968 (2)
M2 16c (0 0 0)Occ = 0.272 (1)Occ = 0.0
Si1 8a (⅛ ⅛ ⅛)
Si2 32e (x x x)x = 0.2198 (6)x = 0.2168 (2)
Si3 96g (x x z)x = 0.1839 (2)x = 0.1822 (1)
z = 0.3698 (6)z = 0.3699 (2)
*: Occ—occupancy factor.
Table 4. Clathrate phases obtained by different synthesis techniques from M4Si4 precursors.
Table 4. Clathrate phases obtained by different synthesis techniques from M4Si4 precursors.
PrecursorPreparation Technique aClathrate Phases ObtainedReferences
Na4Si4CTCNa8Si46 and Na24-xSi136[6,33]
KCTDNa8Si46 and Na24-xSi136[18]
CONa8Si46 and Na24-xSi136[7,25]
SPSNa8Si46 and Na24-xSi136[16,24], this work
K4Si4CTCK8-xSi46[6,26]
KCTDK8-xSi46 and K24-xSi136[19]
COK8-xSi46[7]
SPSK8-xSi46this work
Rb4Si4CTCRb8-xSi46[6,26]
SPSRb8-xSi46 and Rb24-xSi136this work
Cs4Si4CTCCs24-xSi136[6]
SPSCs24-xSi136this work
a CO = chemical oxidation, KCTD = kinetically controlled thermal decomposition, CTC = conventional thermal decomposition, SPS = spark plasma synthesis.

Share and Cite

MDPI and ACS Style

Veremchuk, I.; Beekman, M.; Antonyshyn, I.; Schnelle, W.; Baitinger, M.; Nolas, G.S.; Grin, Y. Binary Alkali-Metal Silicon Clathrates by Spark Plasma Sintering: Preparation and Characterization. Materials 2016, 9, 593. https://doi.org/10.3390/ma9070593

AMA Style

Veremchuk I, Beekman M, Antonyshyn I, Schnelle W, Baitinger M, Nolas GS, Grin Y. Binary Alkali-Metal Silicon Clathrates by Spark Plasma Sintering: Preparation and Characterization. Materials. 2016; 9(7):593. https://doi.org/10.3390/ma9070593

Chicago/Turabian Style

Veremchuk, Igor, Matt Beekman, Iryna Antonyshyn, Walter Schnelle, Michael Baitinger, George S. Nolas, and Yuri Grin. 2016. "Binary Alkali-Metal Silicon Clathrates by Spark Plasma Sintering: Preparation and Characterization" Materials 9, no. 7: 593. https://doi.org/10.3390/ma9070593

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