Next Article in Journal
Peak Broadening Anisotropy and the Contrast Factor in Metal Alloys
Next Article in Special Issue
The Crystal Orbital Hamilton Population (COHP) Method as a Tool to Visualize and Analyze Chemical Bonding in Intermetallic Compounds
Previous Article in Journal
Lattice Parameters of Optical Damage Resistant In-Doped LiNbO3 Crystals
Previous Article in Special Issue
Lu5Pd4Ge8 and Lu3Pd4Ge4: Two More Germanides among Polar Intermetallics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Optimization of Ca14MgSb11 through Chemical Substitutions on Sb Sites: Optimizing Seebeck Coefficient and Resistivity Simultaneously

Department of Chemistry, University of California, One Shields Avenue, Davis, CA 95616, USA
*
Author to whom correspondence should be addressed.
Crystals 2018, 8(5), 211; https://doi.org/10.3390/cryst8050211
Submission received: 17 April 2018 / Revised: 6 May 2018 / Accepted: 8 May 2018 / Published: 13 May 2018
(This article belongs to the Special Issue Compounds with Polar Metallic Bonding)

Abstract

:
In thermoelectric materials, chemical substitutions are widely used to optimize thermoelectric properties. The Zintl phase compound, Yb14MgSb11, has been demonstrated as a promising thermoelectric material at high temperatures. It is iso-structural with Ca14AlSb11 with space group I41/acd. Its iso-structural analog, Ca14MgSb11, was discovered to be a semiconductor and have vacancies on the Sb(3) sites, although in its nominal composition it can be described as consisting of fourteen Ca2+ cations with one [MgSb4]9− tetrahedron, one Sb37− linear anion and four isolated Sb3− anions (Sb(3) site) in one formula unit. When Sn substitutes Sb in Ca14MgSb11, optimized Seebeck coefficient and resistivity were achieved simultaneously although the Sn amount is small (<2%). This is difficult to achieve in thermoelectric materials as the Seebeck coefficient and resistivity are inversely related with respect to carrier concentration. Thermal conductivity of Ca14MgSb11-xSnx remains almost the same as Ca14MgSb11. The calculated zT value of Ca14MgSb10.80Sn0.20 reaches 0.49 at 1075 K, which is 53% higher than that of Ca14MgSb11 at the same temperature. The band structure of Ca14MgSb7Sn4 is calculated to simulate the effect of Sn substitutions. Compared to the band structure of Ca14MgSb11, the band gap of Ca14MgSb7Sn4 is smaller (0.2 eV) and the Fermi-level shifts into the valence band. The absolute values for density of states (DOS) of Ca14MgSb7Sn4 are smaller near the Fermi-level at the top of valence band and 5p-orbitals of Sn contribute most to the valence bands near the Fermi-level.

Graphical Abstract

1. Introduction

Thermoelectric materials have attracted significant attention as they can improve the efficiency of energy through converting wasted heat into electricity. The efficiency of thermoelectric materials can be evaluated through the figure of merit (zT) by Equation (1).
z T =   α 2 ρ κ T
In the equation, α is the Seebeck coefficient, ρ is electrical resistivity, T is the absolute temperature and κ is thermal conductivity. Defects or tiny amounts of chemical substitutions are important to the optimizations of thermoelectric properties in some typical thermoelectric materials as defects can tune carrier concentrations effectively and adjust the Seebeck coefficient and electrical resistivity. Cationic defects, tuned carrier concentrations and optimized thermoelectric properties were observed in Zintl phase compounds [1,2,3]. Defects can also scatter phonons, make the systems phonon-glass-like and decrease lattice thermal conductivity. The Zn defects in Zn-Sb compounds and Cu defects in Cu-chalcogenides are essential for decreasing lattice thermal conductivity and tuning of carrier concentrations [4,5,6,7,8,9,10]. In some Type-I clathrates, defects have been found in frameworks and the ordering of vacancies are important for the tuning of the Seebeck coefficient, electrical resistivity and thermal conductivity [11,12,13,14,15]. Defects in clathrates can also change the band structure as the bonding and antibonding orbitals of framework elements contribute to the bands near the Fermi-level [11].
Considering the fact that defects are actually replacing atoms with voids, small amounts of chemical substitutions also have similar effects on thermoelectric properties. 1.25% Zn substitutions of IIIA atoms in p-type thermoelectric materials Eu/Sr5In2Sb6 and BaGa2Sb2 add a hole in the structure and are able to significantly change the carrier concentrations, therefore changing electrical resistivity and Seebeck coefficient [16,17,18]. In IV-VI materials, 1–2% of potassium, sodium or thulium doping can improve the thermoelectric properties significantly [19,20,21,22,23]. The benefit of K or Na doping may reside in two aspects. The first one is the change of DOS near the Fermi-level, which can increase the effective mass and lead to an improved Seebeck coefficient. The other benefit is the formation of a nano-composite, which can effectively decrease lattice thermal conductivity. More recently, a GeTe-based material with Pb and Bi2Te3 doping (3 mol %) was discovered to have a zT value of 1.9 at 773K [24]. 1% doping of rare earth elements in TAGS-85 samples increases Seebeck coefficient and therefore leads to larger zT values [25,26]. In half-Heusler alloys, small amounts of substitutions (1%) can also dramatically (~50%) improve thermoelectric properties [27,28,29,30].
The Zintl phase compounds, Yb14MgSb11 and Ca14MgSb11, were reported in 2014 and their thermoelectric properties studied [31]. Both of them have the Ca14AlSb11-type of structure (Figure 1 and Table 1), and Ca14MgSb11 was found to have vacancies (2.6%) on the isolated Sb(3) sites. These sites are coordinated by Ca2+ cations and are shown as the non-bonded Sb atoms in Figure 1. Yb14MgSb11 has a zT of ~1 at 1075 K while Ca14MgSb11 has a semiconductor-like resistivity and a zT of 0.32 at 1075 K. Their iso-structural analog, Yb14MnSb11, has also been discovered as a good thermoelectric material and many studies have been conducted to optimize its thermoelectric properties [2,32]. Samples synthesized by powder metallurgy have slightly larger (~20 Å3 out of ~6000 Å3) unit cell parameters than crystals synthesized by the Sn-flux method based on the refinement of powder X-ray diffraction patterns. This suggests that crystals of Yb14MnSb11 synthesized by the Sn-flux method may have small amounts of Yb or Mn vacancies. The smaller resistivity and lower Seebeck coefficient of Yb14MnSb11 samples synthesized as crystals are also consistent with the above statements. Further studies show that 1% Te or Ge substitutions on Sb site can significantly alter the thermoelectric properties [33,34].
The synthesis and thermoelectric properties of Ca14MgSb11−xSnx are systematically investigated. Sn is used to compensate for Sb vacancies and thereby improve the overall thermoelectric properties of Ca14MgSb11. Sn has one electron less than Sb, which will tune the carrier concentration and resistivity. Calculations of the electronic band structures show that Sb(3) sites contribute most near the Fermi-level and therefore substitution of this site may dramatically change the thermoelectric properties [31].

2. Experimental Section

Reagents. Elemental Ca pieces (99.5%, Alfa Aesar, Tewksbury, MA, USA), Mg turnings (99.98%, Sigma-Aldrich, St. Louis, MO, USA), Sb (99.999%, Alfa Aesar, Tewksbury, MA, USA) and Sn (99.3%, Alfa Aesar, Tewksbury, MA, USA) were used for the synthesis. Ca was cut into small pieces while Mg and Sn were used as received. All elements were handled using inert atmosphere techniques, including an argon filled glovebox with water levels <0.5 ppm.
Synthesis of Powder. Quantitative yield, high purity samples were synthesized through a powder metallurgy method [2]. Melting Sn and Mg together at 600 °C with the ratio 1:2.2 to produce Mg2Sn as precursor. Samples with the ratio Ca:Mg:Sb:Mg2Sn = 14:1.1 × (1−2x):11−x:x (x is the Sn amount) were loaded into a 50 cm3 tungsten carbide ball mill vial with one large WC ball (diameter = 11 mm) and two small WC balls (diameter = 8 mm). The mixtures of elements were ball milled on a SPEX 8000 M (SPEX SamplePrep, Metuchen, NJ, USA) for one hour and another 30 min after a 30 min break (the 30 min break prevents the reaction mixture from becoming too hot). The fine powder was transferred to a glovebox and then into a niobium tube, which was sealed by arc welding under argon and further jacketed under vacuum in fused silica. Ca14MgSb11−xSnx is annealed at 800 °C for 4 days with a heating rate of 30 °C/h [31].
Powder X-ray Diffraction. Samples were examined using a Bruker zero background holder on a Bruker D8 Advance Diffractometer operated at 40 kV and 40 mA utilizing Cu radiation. radiation is removed by a Ni filter. WinPLOTR (version Jan 2012, part of the FullProf suite of programs, https://www.ill.eu/sites/fullprof/ University of Rennes 1, Rennes, France) software was used for background subtraction and pattern analysis, and EDPCR 2.00 software (part of the FullProf suite of programs, https://www.ill.eu/sites/fullprof/ University of Rennes 1, Rennes, France) was used to perform Le Bail refinement [35].
Consolidation of Powder. The bulk powder samples were consolidated into dense pellets via a Dr. Sinter Lab Jr. SPS-211Lx or SPS-2050 spark plasma sintering (SPS) system (Sumitomo, Tokyo, Japan) in a 12.7 mm high-density graphite die (POCO) under vacuum (<10 Pa). The temperature was increased from room temperature to 1000–1025 K in 15 min, and remained stable for 5–15 min. When the temperature reached the maximum, the force loaded increased from 3 kN to 5–8 kN. The samples were cooled to room temperature afterwards. The geometrical sample density was larger than 95% of the theoretical density.
Thermal Conductivity. Thermal diffusivity (D) measurement was conducted on the pellet obtained from SPS from 300 K to 1075 K on a Netzsch LFA-457 laser flash unit (Netzsch, Burlington, MA, USA). The pellet surfaces were well polished and coated with graphite. The measurement was conducted under dynamic argon atmosphere with a flow rate of 50 mL/min. Thermal conductivity was calculated using the equation κ = D × ρ × Cp. Room-temperature density was measured from a volume method and the high-temperature density was derived using thermal expansion data from a previous paper on Yb14MnSb11 [36]. The Cp was taken from previous papers, which was measured by differential scanning calorimetry (DSC) [2].
Electrical Transport Properties. A Linseis LSR-3 unit (Linseis, Robinsville, NJ, USA) was employed to measure Seebeck coefficient and electrical resistivity via a four-probe method from 325 K to 1075 K under a helium atmosphere on a bar-shaped sample [37]. The sample which had been previously measured on the LFA instrument was cut into a 2 × 2 × 11 mm bar using a Buehler diamond saw and polished before measurement. The probe distance was 8 mm. For convenience and clarity, Seebeck coefficient, electrical resistivity and thermal conductivity were fit to six-order polynomial functions to calculate zT values. Room-temperature Hall coefficient was measured with a Quantum Design physical property measurement system (PPMS) from 7 T to −7 T by 5-point ac technique. Platinum leads were connected to the pressed pellet through silver paste. Carrier concentration was calculated using equation RH = −1/ne using the average of RH from different magnetic fields.
Quantum-chemical calculations. Density functional band structure calculations for Ca14MgSb11 and Ca14MgSb7Sn4 were performed using the linear-muffin tin orbital method (TB-LMTO, Stuttgart, Germany, version 47.1b) within the tight binding approximation [38,39,40,41,42]. The density of states (DOS) and band structures were calculated after convergence of the total energy on a dense k-mesh with 12 × 12 × 12 points, with 163 irreducible k-points. A basis set containing Ca(4s,3d), Mg(3s,3p), and Sb(5s,5p) orbitals was employed for a self-consistent calculation, with Ca(4p), Mg(3d) and Sb(5d,4f) functions being downfolded.

3. Results and Discussion

Synthesis and Structure. Mg2Sn used in the reaction as a precursor was verified by powder X-ray diffraction. There may be very small peaks indicating the existence of unreacted Mg since extra Mg is used in the precursor synthesis. As mentioned in a previous paper, Ca14MgSb11 synthesized by powder metallurgy may contain a minor amount of impurity Ca11Sb10 and its existence has a limited effect on the thermoelectric properties [31,43]. Ca14MgSb11−xSnx (x = 0.05 and 0.10) also contained minor amounts of impurities Ca11Sb10 while the Le Bail refinement of the powder X-ray diffraction pattern for Ca14MgSb10.80Sn0.20 indicates a phase pure sample (Figure 2). These samples are air-sensitive and oxidized rapidly upon exposure, leading to poor quality of powder X-ray diffraction patterns, not suitable for Rietveld refinement. The refined unit cell parameters are listed in Table 2. Generally, the unit cell parameters show a slight increase with the increasing Sn amounts.
Thermoelectric Properties.Figure 3 shows the results of thermoelectric properties measurement of Ca14MgSb10.95Snx (x = 0.05, 0.1 and 0.2). The Seebeck coefficient and resistivity of these samples change dramatically with the small changes of Sn compositions. The Seebeck coefficient of Ca14MgSb10.95Sn0.05 has a similar trend to that of Ca14MgSb11, but the values are much lower. Ca14MgSb10.90Sn0.10 and Ca14MgSb10.80Sn0.20 have almost the same Seebeck coefficient values, which are slightly higher than the Seebeck coefficient of Ca14MgSb11. The linear increase of Seebeck coefficients of Ca14MgSb10.90Sn0.10 and Ca14MgSb10.80Sn0.20 within the measured temperature region is the most significant change caused by Sn substitutions. The Seebeck coefficient of Ca14MgSb11, which decreases at low temperature and increases at high temperature, is attributed to a combination of increasing temperature and change of carrier concentration based on Equation (2) [19,31].
α = 8 π 2 k B 2 3 e h 2 m * T   ( π 3 n ) 2 3
In the equation, kB is the Boltzmann constant, h is the Planck constant, m* is the effective mass of carriers and n is carrier concentration. In Ca14MgSb10.90Sn0.10 and Ca14MgSb10.80Sn0.20, no effect of carrier concentration change is observed in the Seebeck coefficient, which indicates that the substitution of Sn decreases the band gap.
The resistivity of these samples are semiconductor-like, but the details are different. Ca14MgSb10.95Sn0.05 has a larger resistivity than that of Ca14MgSb11 at room temperature and a 33% lower resistivity than Ca14MgSb11 at high temperatures. The resistivity of Ca14MgSb10.90Sn0.10 is higher than the other two samples, especially in the high temperature region. The resistivity of Ca14MgSb10.80Sn0.20 follows the expectation, which is lower than that of Ca14MgSb11 for the entire temperature region. The activation energy (Ea) can be calculated based on Equation (3) and are listed in Table 3 [31].
ln ρ = ln ρ0 + Ea/2kBT
It can be seen that the activation energy drops from x = 0.05 to x = 0.10. The carrier concentration of Ca14MgSb10.80Sn0.20 is measured to be 3.5 × 1019/cm3 at room temperature, which is slightly higher than that of Ca14MgSb11. Therefore, the carrier concentration has a small change within the three compositions. The Sn substitution of Sb is expected to make Ca14MgSb11 more conductive as Sn has one electron less than Sb and this increases the carrier concentration of p-type semiconductor Ca14MgSb11. However, the carrier concentration change is complex due to the combinations of defects, hole additions caused by Sn doping and the band structure change. The addition of Sn changes the activation energy, which corresponds to the band gap, but this cannot change the intrinsic semiconductor properties of Ca14MgSb11.
The thermal conductivities of Ca14MgSb11−xSnx (x = 0.05, 0.10 and 0.20) are listed in Figure 2c. The sample of x = 0.05 has the largest thermal conductivity and sample of x = 0.20 has almost the same thermal conductivity as Ca14MgSb11. Lorenz numbers of these samples at different temperatures are calculated using Equations (4) and (5) [4,44].
L =   ( k B e ) 2 3 F 0 ( η ) F 2 ( η ) 4 F 1 ( η ) 2 F 0 ( η ) 2
α =   k B e [ 2 F 1 ( η ) F 0 ( η ) η ]
κ l a t =   κ   κ e =   κ L σ T  
In the equations, L is the Lorenz number, η is the reduced Fermi-level, Fn(η) is the Fermi-Dirac integral, κlat is the lattice thermal conductivity and κe is electronic thermal conductivity from electrical conductivity. The lattice thermal conductivity can be calculated by subtracting the electronic thermal conductivity from the total thermal conductivity. The calculated lattice thermal conductivity shows that the lattice term contributes the most to the total thermal conductivity and the electronic term has larger contributions at high temperature regions due to the decrease of resistivity. The calculated zT values show that Ca14MgSb10.80Sn0.20 has the largest maximum zT value of 0.49 at 1075 K, which is 53% higher than that of Ca14MgSb11 due to optimized resistivity. Ca14MgSb10.95Sn0.05 has a maximum zT value of 0.04 at 1075 K due to a lower Seebeck coefficient and Ca14MgSb10.90Sn0.10 has a maximum zT value of 0.12 due to a much larger resistivity.

4. DOS Calculation

The DOS diagram for Ca14MgSb7Sn4 is shown in Figure 4. This composition is assumed for calculating the DOS of Sn-substituted Ca14MgSb11, as the Sb(3) site is assumed to be occupied by Sn. Several reasons support this hypothesis as the Sb(3) site is known to show vacancies and Sb(3) is the isolated Sb3− site with no direct covalent bonding to other atoms, which is the easiest to be substituted. Because of the large unit cell of Ca14MgSb11, Sb(3) is assumed to be fully replaced by Sn although actually at most only 5% of Sb(3) is occupied by Sn. Therefore, the composition used in the calculations is Ca14MgSb7Sn4.
The band structure of Ca14MgSb7Sn4 (Figure 4) has similar features to that of Ca14MgSb11 [31]. The Fermi-level falls into the valence band, to which p-orbitals of Sb and Sn have dominating contributions. Sn makes the largest contribution at the top of the valence band near the Fermi-level. The conduction band is mainly from orbitals of Ca, especially when energy is >1.2 eV, and there is a sharp peak at the bottom of the conduction band, which originates from the anti-bonding orbitals of the linear Sb37− units and orbitals of Ca. The band gap between the top of the valence band (+0.58 eV) and the bottom of the conduction band (+0.78 eV) is 0.20 eV. This value is smaller than the band gap of Ca14MgSb11 (0.6 eV) but is similar to the activation energy calculated from resistivity. It is also noticeable that there is an energy gap of 0.15 eV between the sharp peak and the states at higher energy in the conduction band. Compared to Ca14MgSb11, Ca14MgSb7Sn4 has smaller absolute values for the density of states at the top of valence band near the Fermi-level. Therefore, when Sb is replaced by Sn, two major changes happen to the valence band. The first one is the shift of Fermi-level into the valence band and the other one is the decrease of absolute values of DOS near Fermi-level.
In the experimental samples Ca14MgSb11−xSnx (x = 0.05, 0.10 and 0.20), the Sn amount is much lower than that used in the calculations for the DOS. Comparing the electronic calculations for Ca14MgSb11 and Ca14MgSb7Sn4, a small decrease of band gap is expected. This is supported by the experimental results. The Seebeck coefficient strongly depends on the density of states at the Fermi-level [19,45,46].
S =   π 2 3 k B q k B T [ 1 n d n ( E ) d E + 1 μ d μ ( E ) d E ] E = E F
S =   π 2 3 k B q k B T { 1 n d [ g ( E ) f ( E ) ] d E + 1 μ d μ ( E ) d E } E = E F
In Equations (7) and (8), μ is the mobility, n is the carrier density, g(E) is density of states and f(E) is Fermi function [19]. In Ca14MgSb11−xSnx (x = 0.05, 0.10 and 0.20), as mentioned above, the Fermi level shifts into valence band and the change of DOS at the Fermi level is difficult to determine. Based on the measured Seebeck coefficient, it can be concluded that Fermi-level of Ca14MgSb10.95Sn0.05 falls into a valley of DOS and leads to a small Seebeck coefficient, while Ca14MgSb10.9Sn0.1 and Ca14MgSb10.8Sn0.2 have Fermi-levels with a large DOS and a large Seebeck coefficient. Mobility is inversely related to resistivity by the equation, 1/ρ = σ = neμ (n is carrier density, e is the charge carrier, u is mobility). Mobility depends on the band structure and affects resistivity significantly as the carrier densities for these samples are approximately the same.

5. Summary

Ca14MgSb11−xSnx (x = 0.05, 0.1 and 0.2) solid solutions were synthesized by powder metallurgy with Mg2Sn as precursor. Although the Sn amount in the samples is small (<2%), optimized Seebeck coefficient and resistivity were achieved simultaneously with similar thermal conductivity. This is very rare, as the Seebeck coefficient and resistivity are inversely related with respected to carrier concentration. The Seebeck coefficient of Ca14MgSb10.95Sn0.05 has a similar temperature dependence compared with Ca14MgSb11, while the Seebeck coefficients of Ca14MgSb10.90Sn0.10 and Ca14MgSb10.80Sn0.20 linearly increase from room temperature to high temperature, different from the trend of Ca14MgSb11. Their resistivity shows semiconductor behavior and the activation energy decreases with Sn amount. The zT value of Ca14MgSb10.80Sn0.20 reaches 0.49 at 1075 K, which is 53% higher than that of Ca14MgSb11 at the same temperature. The band structure of Ca14MgSb7Sn4 is calculated to better understand the effect of Sn substitution. The band gap of Ca14MgSb7Sn4 is 0.2 eV and the Fermi-level shifts into the valence band. 5p-orbitals of Sn contribute most to the valence bands near Fermi-level at the top of valence band and the overall DOS of Ca14MgSb7Sn4 are smaller in the valence band compared to Ca14MgSb11. The substitution of Sn increases the carrier concentration and decreases both the Seebeck coefficient and resistivity as expected.

Author Contributions

Y.H. and S.M.K conceived and designed the experiments. Y.H. performed the experiments and analyzed the data. K.A.L. contributed to the DOS calculations. Y.H. wrote the first draft of the manuscript; all co-authors contributed to the final version of the manuscript.

Acknowledgments

The authors thank Julia Zaikina for her help in TB-LMTO calculations. This work was supported by the NASA Science Mission Directorate’s Radioisotope Power Systems. Financial support from NEUP is gratefully acknowledged.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Pomrehn, G.S.; Zevalkink, A.; Zeier, W.G.; van de Walle, A.; Snyder, G.J. Defect-Controlled Electronic Properties in AZn2Sb2 Zintl Phases. Angew. Chem. In. Ed. 2014, 53, 3422–3426. [Google Scholar] [CrossRef] [PubMed]
  2. Grebenkemper, J.H.; Hu, Y.; Barrett, D.; Gogna, P.; Huang, C.-K.; Bux, S.K.; Kauzlarich, S.M. High Temperature Thermoelectric Properties of Yb14MnSb11 Prepared from Reaction of MnSb with the Elements. Chem. Mater. 2015, 27, 5791. [Google Scholar] [CrossRef]
  3. Zevalkink, A.; Zeier, W.G.; Cheng, E.; Snyder, J.; Fleurial, J.-P.; Bux, S. Nonstoichiometry in the Zintl Phase Yb1−δZn2Sb2 as a Route to Thermoelectric Optimization. Chem. Mater. 2014, 26, 5710. [Google Scholar] [CrossRef]
  4. Toberer, E.S.; Cox, C.A.; Brown, S.R.; Ikeda, T.; May, A.F.; Kauzlarich, S.M.; Snyder, G.J. Traversing the Metal-Insulator Transition in a Zintl Phase: Rational Enhancement of Thermoelectric Efficiency in Yb14Mn1−xAlxSb11. Adv. Funct. Mater. 2008, 18, 2795. [Google Scholar] [CrossRef]
  5. Toberer, E.S.; Rauwel, P.; Gariel, S.; Taftø, J.; Jeffrey Snyder, G. Composition and the Thermoelectric Performance of β-Zn4Sb3. J. Mater. Chem. 2010, 20, 9877. [Google Scholar] [CrossRef]
  6. Snyder, G.J.; Christensen, M.; Nishibori, E.; Caillat, T.; Iversen, B.B. Disordered Zinc in Zn4Sb3 with Phonon-glass and Electron-crystal Thermoelectric Properties. Nat. Mater. 2004, 3, 458. [Google Scholar] [CrossRef] [PubMed]
  7. He, Y.; Zhang, T.; Shi, X.; Wei, S.H.; Chen, L. High Thermoelectric Performance in Copper Telluride. NPG Asia Mater. 2015, 7, e210. [Google Scholar] [CrossRef]
  8. He, Y.; Day, T.; Zhang, T.; Liu, H.; Shi, X.; Chen, L.; Snyder, G.J. High Thermoelectric Performance in Non-Toxic Earth-abundant Copper Sulfide. Adv. Mater. 2014, 26, 3974. [Google Scholar] [CrossRef] [PubMed]
  9. Ge, Z.H.; Zhang, B.P.; Chen, Y.X.; Yu, Z.X.; Liu, Y.; Li, J.F. Synthesis and Transport Property of Cu1.8S as a Promising Thermoelectric Compound. Chem. Commun. 2011, 47, 12697. [Google Scholar] [CrossRef] [PubMed]
  10. Liu, H.; Shi, X.; Xu, F.; Zhang, L.; Zhang, W.; Chen, L.; Li, Q.; Uher, C.; Day, T.; Snyder, G.J. Copper Ion Liquid-like Thermoelectrics. Nat. Mater. 2012, 11, 422. [Google Scholar] [CrossRef] [PubMed]
  11. Kaltzoglou, A.; Fässler, T.; Christensen, M.; Johnsen, S.; Iversen, B.; Presniakov, I.; Sobolev, A.; Shevelkov, A. Effects of the Order–disorder Phase Transition on the Physical Properties of A8Sn442 (A = Rb, Cs). J. Mater. Chem. 2008, 18, 5630. [Google Scholar] [CrossRef]
  12. Kaltzoglou, A.; Fässler, T.F.; Gold, C.; Scheidt, E.W.; Scherer, W.; Kume, T.; Shimizu, H. Investigation of Substitution Effects and the Phase Transition in Type-I clathrates RbxCs8–xSn442 (1.3 ≤ x ≤ 2.1) Using Single-crystal X-ray Diffraction, Raman Spectroscopy, Heat Capacity and Electrical Resistivity Measurements. J. Solid State Chem. 2009, 182, 2924. [Google Scholar] [CrossRef]
  13. Chung, D.; Hogan, T.; Brazis, P.; Roccilane, M.; Kannewurf, C.; Bastea, M.; Uher, C.; Kanatzidis, M.G. CsBi4Te6: A High-Performance Thermoelectric Material for Low-Temperature Applications. Science 2000, 287, 1024. [Google Scholar] [CrossRef] [PubMed]
  14. Dolyniuk, J.A.; Owens-Baird, B.; Wang, J.; Zaikina, J.V.; Kovnir, K. Clathrate Thermoelectrics. Mater. Sci. Eng. R 2016, 108, 1. [Google Scholar] [CrossRef]
  15. Zaikina, J.V.; Kovnir, K.A.; Sobolev, A.V.; Presniakov, I.A.; Prots, Y.; Baitinger, M.; Schnelle, W.; Olenev, A.V.; Lebedev, O.I.; Van Tendeloo, G.; et al. Sn20.53.5As22I8: A Largely Disordered Cationic Clathrate with a New Type of Superstructure and Abnormally Low Thermal Conductivity. Chem. Eur. J. 2007, 13, 5090. [Google Scholar] [CrossRef] [PubMed]
  16. Aydemir, U.; Zevalkink, A.; Ormeci, A.; Gibbs, Z.M.; Bux, S.; Snyder, G.J. Thermoelectric Enhancement in BaGa2Sb2 by Zn Doping. Chem. Mater. 2015, 27, 1622. [Google Scholar] [CrossRef]
  17. Chanakian, S.; Aydemir, U.; Zevalkink, A.; Gibbs, Z.M.; Fleurial, J.-P.; Bux, S.; Snyder, G.J. High Temperature Thermoelectric Properties of Zn-doped Eu5In2Sb6. J. Mater. Chem. C 2015, 3, 10518. [Google Scholar] [CrossRef]
  18. Chanakian, S.; Zevalkink, A.; Aydemir, U.; Gibbs, Z.M.; Pomrehn, G.; Fleurial, J.-P.; Bux, S.; Snyder, G.J. Enhanced Thermoelectric Properties of Sr5In2Sb6 via Zn-doping. J. Mater. Chem. A 2015, 3, 10289. [Google Scholar] [CrossRef]
  19. Heremans, J.P.; Jovovic, V.; Toberer, E.S.; Saramat, A.; Kurosaki, K.; Charoenphakdee, A.; Yamanaka, S.; Snyder, G.J. Enhancement of Thermoelectric Efficiency in PbTe by Distortion of the Electronic Density of States. Science 2008, 321, 554. [Google Scholar] [CrossRef] [PubMed]
  20. Zhang, Q.; Cao, F.; Liu, W.; Lukas, K.; Yu, B.; Chen, S.; Opeil, C.; Broido, D.; Chen, G.; Ren, Z. Heavy Doping and Band Engineering by Potassium to Improve the Thermoelectric Figure of Merit in p-Type PbTe, PbSe, and PbTe1-ySey. J. Am. Chem. Soc. 2012, 134, 10031. [Google Scholar] [CrossRef] [PubMed]
  21. Pei, Y.; LaLonde, A.; Iwanaga, S.; Snyder, G.J. High Thermoelectric Figure of Merit in Heavy Hole Dominated PbTe. Energy Environ. Sci. 2011, 4, 2085. [Google Scholar] [CrossRef]
  22. Girard, S.N.; He, J.; Zhou, X.; Shoemaker, D.; Jaworski, C.M.; Uher, C.; Dravid, V.P.; Heremans, J.P.; Kanatzidis, M.G. High Performance Na-doped PbTe-PbS Thermoelectric Materials: Electronic Density of States Modification and Shape-controlled Nanostructures. J. Am. Chem. Soc. 2011, 133, 16588. [Google Scholar] [CrossRef] [PubMed]
  23. Zhao, L.D.; Tan, G.; Hao, S.; He, J.; Pei, Y.; Chi, H.; Wang, H.; Gong, S.; Xu, H.; Dravid, V.P.; et al. Ultrahigh Power Factor and Thermoelectric Performance in Hole-doped Single-Crystal SnSe. Science 2016, 351, 141. [Google Scholar] [CrossRef] [PubMed]
  24. Wu, D.; Zhao, L.D.; Hao, S.; Jiang, Q.; Zheng, F.; Doak, J.W.; Wu, H.; Chi, H.; Gelbstein, Y.; Uher, C.; et al. Origin of the High Performance in GeTe-Based Thermoelectric Materials upon Bi2Te3 Doping. J. Ame. Chem. Soc. 2014, 136, 11412. [Google Scholar] [CrossRef] [PubMed]
  25. Levin, E.M.; Cook, B.A.; Harringa, J.L.; Bud’ko, S.L.; Venkatasubramanian, R.; Schmidt-Rohr, K. Analysis of Ce- and Yb-Doped TAGS-85 Materials with Enhanced Thermoelectric Figure of Merit. Adv. Funct. Mater. 2011, 21, 441. [Google Scholar] [CrossRef]
  26. Levin, E.M.; Bud’ko, S.L.; Schmidt-Rohr, K. Enhancement of Thermopower of TAGS-85 High-Performance Thermoelectric Material by Doping with the Rare Earth Dy. Adv. Funct. Mater. 2012, 22, 2766. [Google Scholar] [CrossRef]
  27. Bhattacharya, S.; Pope, A.L.; Littleton, R.T.; Tritt, T.M.; Ponnambalam, V.; Xia, Y.; Poon, S.J. Effect of Sb Doping on the Thermoelectric Properties of Ti-based half-Heusler Compounds, TiNiSn1−xSbx. Appl. Phys. Lett. 2000, 77, 2476. [Google Scholar] [CrossRef]
  28. Shen, Q.; Chen, L.; Goto, T.; Hirai, T.; Yang, J.; Meisner, G.P.; Uher, C. Effects of Partial Substitution of Ni by Pd on the Thermoelectric Properties of ZrNiSn-based half-Heusler Compounds. Appl. Phys. Lett. 2001, 79, 4165. [Google Scholar] [CrossRef]
  29. Appel, O.; Zilber, T.; Kalabukhov, S.; Beeri, O.; Gelbstein, Y. Morphological Effects on the Thermoelectric Properties of Ti0.3Zr0.35Hf0.35Ni1+δSn Alloys Following Phase Separation. J. Mater. Chem. C 2015, 3, 11653. [Google Scholar] [CrossRef]
  30. Casper, F.; Graf, T.; Chadov, S.; Balke, B.; Felser, C. Half-Heusler Compounds: Novel Materials for Energy and Spintronic Applications. Semicond. Sci. Technol. 2012, 27, 063001. [Google Scholar] [CrossRef]
  31. Hu, Y.; Wang, J.; Kawamura, A.; Kovnir, K.; Kauzlarich, S.M. Yb14MgSb11 and Ca14MgSb11—New Mg-Containing Zintl Compounds and Their Structures, Bonding, and Thermoelectric Properties. Chem. Mater. 2015, 27, 343. [Google Scholar] [CrossRef]
  32. Toberer, E.S.; Brown, S.R.; Ikeda, T.; Kauzlarich, S.M.; Jeffrey Snyder, G. High Thermoelectric Efficiency in Lanthanum Doped Yb14MnSb11. Appl. Phys. Lett. 2008, 93, 062110. [Google Scholar] [CrossRef]
  33. Yi, T.; Abdusalyamova, M.N.; Makhmudov, F.; Kauzlarich, S.M. Magnetic and Transport Properties of Te Doped Yb14MnSb11. J. Mater. Chem. 2012, 22, 14378. [Google Scholar] [CrossRef]
  34. Rauscher, J.F.; Cox, C.A.; Yi, T.; Beavers, C.M.; Klavins, P.; Toberer, E.S.; Snyder, G.J.; Kauzlarich, S.M. Synthesis, Structure, Magnetism, and High Temperature Thermoelectric Properties of Ge Doped Yb14MnSb11. Dalton Trans. 2010, 39, 1055. [Google Scholar] [CrossRef] [PubMed]
  35. Rodríquez-Carvajal, J. Recent Advances in Magnetic Structure Determination by Neutron Powder Diffraction. Physica B 1993, 192, 55. [Google Scholar] [CrossRef]
  36. Ravi, V.; Firdosy, S.; Caillat, T.; Brandon, E.; Van Der Walde, K.; Maricic, L.; Sayir, A. Thermal Expansion Studies of Selected High-Temperature Thermoelectric Materials. J. Electron. Mater. 2009, 38, 1433. [Google Scholar] [CrossRef]
  37. Mackey, J.; Dynys, F.; Sehirlioglu, A. Uncertainty Analysis for Common Seebeck and Electrical Resistivity Measurement Systems. Rev. Sci. Instrum. 2014, 85, 085119. [Google Scholar] [CrossRef] [PubMed]
  38. Andersen, O.K. Linear Methods in Band Theory. Phys. Rev. B 1975, 12, 3060. [Google Scholar] [CrossRef] [Green Version]
  39. Andersen, O.K.; Jepsen, O. Explicit, First-Principles Tight-Binding Theory. Phys. Rev. Lett. 1984, 53, 2571. [Google Scholar] [CrossRef]
  40. Andersen, O.K.; Pawlowska, Z.; Jepsen, O. Illustration of the Linear-muffin-tin-orbital Tight-binding Representation: Compact Orbitals and Charge Density in Si. Phys. Rev. B 1986, 34, 5253. [Google Scholar] [CrossRef]
  41. Nowak, H.J.; Andersen, O.K.; Fujiwara, T.; Jepsen, O.; Vargas, P. Electronic-structure Calculations for Amorphous Solids Using the Recursion Method and Linear Muffin-tin Orbitals: Application to Fe80B20. Phys. Rev. B 1991, 44, 3577. [Google Scholar] [CrossRef]
  42. 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. [Google Scholar] [CrossRef]
  43. Brown, S.R.; Kauzlarich, S.M.; Gascoin, F.; Jeffrey Snyder, G. High-temperature Thermoelectric Studies of A11Sb10 (A=Yb, Ca). J. Solid State Chem. 2007, 180, 1414. [Google Scholar] [CrossRef]
  44. Fistul, V.I. Heavily Doped Semiconductor; Plenum Press: New York, NY, USA, 1969. [Google Scholar]
  45. Cutler, M.; Mott, N.F. Observation of Anderson Localization in an Electron Gas. Phys. Rev. 1969, 181, 1336. [Google Scholar] [CrossRef]
  46. Mahan, G.D.; Sofo, J.O. The Best Thermoelectric. Proc. Natl. Acad. Sci. USA 1996, 93, 7436. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Unit cell of Ca14MgSb11 projected along the c-axis. Ca and Sb atoms are represented by blue and red spheres, Sb37 ions are shown with yellow bonds and the green tetrahedra are [MgSb4]9 clusters.
Figure 1. Unit cell of Ca14MgSb11 projected along the c-axis. Ca and Sb atoms are represented by blue and red spheres, Sb37 ions are shown with yellow bonds and the green tetrahedra are [MgSb4]9 clusters.
Crystals 08 00211 g001
Figure 2. Refinement of the powder X-ray diffraction patterns of (a) Ca14MgSb10.95Sn0.05, (b) Ca14MgSb10.90Sn0.10 and (c) Ca14MgSb10.80Sn0.20.
Figure 2. Refinement of the powder X-ray diffraction patterns of (a) Ca14MgSb10.95Sn0.05, (b) Ca14MgSb10.90Sn0.10 and (c) Ca14MgSb10.80Sn0.20.
Crystals 08 00211 g002
Figure 3. (a) Seebeck coefficient, (b) resistivity (insert: ln ρ vs. 1/T plots), (c) thermal conductivity with lattice thermal conductivities shown in dashed lines and (d) zT values of Ca14MgSb10.95Sn0.05, Ca14MgSb10.90Sn0.10 and Ca14MgSb10.80Sn0.20. In all plots, data for Ca14MgSb11 are shown in black curves as references.
Figure 3. (a) Seebeck coefficient, (b) resistivity (insert: ln ρ vs. 1/T plots), (c) thermal conductivity with lattice thermal conductivities shown in dashed lines and (d) zT values of Ca14MgSb10.95Sn0.05, Ca14MgSb10.90Sn0.10 and Ca14MgSb10.80Sn0.20. In all plots, data for Ca14MgSb11 are shown in black curves as references.
Crystals 08 00211 g003
Figure 4. Density of states (DOS) diagram for Ca14MgSb7Sn4. (a) Total DOS together with partial contributions from Ca and 5p-orbitals of Sn and Sb (b) Partial DOS showing the contributions of only 5p-orbitals of Sn and Sb.
Figure 4. Density of states (DOS) diagram for Ca14MgSb7Sn4. (a) Total DOS together with partial contributions from Ca and 5p-orbitals of Sn and Sb (b) Partial DOS showing the contributions of only 5p-orbitals of Sn and Sb.
Crystals 08 00211 g004
Table 1. Wyckoff positions and atomic coordinates [×104] of atoms for Ca14MgSb11.
Table 1. Wyckoff positions and atomic coordinates [×104] of atoms for Ca14MgSb11.
AtomWyckoff Positionsxyz
Ca132g9580(2)9274(2)8281(1)
Ca232g9771(2)1264(2)79(1)
Ca316e3553(2)02500
Ca432g1781(2)4030(2)8439(1)
Mg18a025008750
Sb116f1364(1)3864(1)1250
Sb232g37(1)1100(1)8059(1)
Sb3 *32g8705(1)9751(1)9516(1)
Sb48b025001250
* Sb3 site has 2.6% vacancy.
Table 2. Unit cell parameters from refinement of powder X-ray diffraction patterns.
Table 2. Unit cell parameters from refinement of powder X-ray diffraction patterns.
Sn Amount Used in Synthesisa (Å)c (Å)V3)
0.0016.73(1)22.54(1)6309(1)
0.0516.72(1)22.60(1)6318(3)
0.1016.73(1)22.59(1)6323(3)
0.2016.73(1)22.62(1)6331(3)
Table 3. Calculated activation energy of Ca14MgSb11−xSnx.
Table 3. Calculated activation energy of Ca14MgSb11−xSnx.
CompositionActivation Energy (eV)
x = 0.000.15
x = 0.050.17
x = 0.100.08
x = 0.200.10

Share and Cite

MDPI and ACS Style

Hu, Y.; Lee, K.; Kauzlarich, S.M. Optimization of Ca14MgSb11 through Chemical Substitutions on Sb Sites: Optimizing Seebeck Coefficient and Resistivity Simultaneously. Crystals 2018, 8, 211. https://doi.org/10.3390/cryst8050211

AMA Style

Hu Y, Lee K, Kauzlarich SM. Optimization of Ca14MgSb11 through Chemical Substitutions on Sb Sites: Optimizing Seebeck Coefficient and Resistivity Simultaneously. Crystals. 2018; 8(5):211. https://doi.org/10.3390/cryst8050211

Chicago/Turabian Style

Hu, Yufei, Kathleen Lee, and Susan M. Kauzlarich. 2018. "Optimization of Ca14MgSb11 through Chemical Substitutions on Sb Sites: Optimizing Seebeck Coefficient and Resistivity Simultaneously" Crystals 8, no. 5: 211. https://doi.org/10.3390/cryst8050211

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