Next Article in Journal
ZnO Nanostructure Templates as a Cost-Efficient Mass-Producible Route for the Development of Cellular Networks
Previous Article in Journal
Recycled PET Nanofibers for Water Filtration Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

High Pressure Synthesis of p-Type CeyFe4−xCoxSb12 Skutterudites

State Key Laboratory of Metastable Materials Science and Technology, Yanshan University, Qinhuangdao 066004, Hebei, China
*
Author to whom correspondence should be addressed.
Materials 2016, 9(4), 257; https://doi.org/10.3390/ma9040257
Submission received: 18 February 2016 / Revised: 17 March 2016 / Accepted: 28 March 2016 / Published: 31 March 2016
(This article belongs to the Section Energy Materials)

Abstract

:
Co-substituted p-Type CeFe4−xCoxSb12 skutterudites were successfully synthesized with a high pressure synthesis method. The structure, composition, and thermoelectric properties were investigated. The obtained CeyFe4−xCoxSb12 samples show the skutterudite structure of I m 3 ¯ symmetry. The hole concentration decreases with elevating Co substitution level, leading to increased Seebeck coefficient and electrical resistivity. Meanwhile, the filling fraction of Ce decreases, which is unfavorable for reducing the lattice thermal conductivity. As a result, the thermoelectric performance of CeyFe4−xCoxSb12 deteriorates with higher Co content. The maximal ZT of 0.91 was achieved at 763 K for the optimal Ce0.92Fe4Sb12 sample.

1. Introduction

Thermoelectric materials have attracted great research interest during the past few decades due to their prospective applications in energy conversion to alleviate the greenhouse effect and environmental pollution [1]. In general, the performance of thermoelectric materials is characterized by the dimensionless figure of merit ZT
Z T = S 2 T / ρ κ
where S, T, ρ, and κ are the Seebeck coefficient, temperature, electrical resistivity, and thermal conductivity, respectively. To make the thermoelectric device competitive with other energy conversion devices, thermoelectric materials with high ZT in the working temperature range must be fabricated, which is a practical challenge because of the entanglements between S, ρ, and κ. Improvement of one parameter usually adversely influences others. To unravel the entanglements, Slack proposed the concept of phonon-glass electron-crystal (PGEC) in 1995 [2]. Skutterudites, a prototype of the PGEC concept, are among the most promising thermoelectric materials working in the intermediate temperature region [3].
Various approaches have been taken to improve the thermoelectric properties of skutterudites, such as void filling [4,5,6,7,8,9,10,11,12] and lattice atom substituting [13,14,15,16,17,18,19]. Compared with n-type CoSb3-based skutterudites for which ZT records are constantly refreshed [4,5,6], the developments in p-type skutterudites relatively lag behind [9,10]. Elemental filled FeSb3-based skutterudites are one of the most important p-type skutterudites. In this class of materials, κ is greatly suppressed due to the rattling filler atoms (with a near unit filling fraction), and a low ρ is ensured by the high hole concentration (>1021 cm−3); meanwhile, a moderate S is maintained due to the large effective mass of hole [20]. Further enhancement of ZT for elemental filled FeSb3-based skutterudites might be possible through enhancing power factor PF and suppressing thermal conductivity.
P F = S 2 / ρ
Substituting framework atoms with electron-richer atoms (such as Co/Ni for Fe) would be a possible method. By such kind of substitutions, the hole concentration can be lowered to benefit enhancement of S. In addition, the impurity scattering of phonons due to mass fluctuation may contribute to further lower κ.
The effects of Co/Ni substitution for Fe on the thermoelectric properties of CeFe4Sb12 were investigated previously [21,22,23,24]. In these studies, the time-consuming solid state reaction method was employed to fabricate the samples, which usually takes several days. Recent works have highlighted pressure as a fundamental thermodynamic variable in synthesizing elemental filled CoSb3 [25,26,27] and Te-doped CoSb3 [28,29,30]. High pressure can lower the reaction temperature and facilitate the synthesis of metastable phase by shifting reaction equilibrium [31]. In this work, Co-substituted CeFe4Sb12 skutterudites were synthesized by using a time-saving high pressure synthesis (HPS) method. We found that the actual filling fraction of Ce decreases slightly with elevating Co substitution level, and a notable decrease in the hole concentration is observed. The hole concentration decrease shows a more pronounced effect on the electrical resistivity, leading to a significantly reduced power factor in the sample with the highest Co level. As a result, the thermoelectric performance deteriorates with higher Co content. The optimal Ce0.92Fe4Sb12 showed a maximal ZT of 0.91 at 763 K.

2. Experimental Methods

Ce (99.9%), Fe (99.5%), Co (99.8%), and Sb (99.999%) powders were mixed according to the atomic ratio of 1:(4−a):a:12 (a = 0, 0.5, and 1, respectively). The mixture was filled into a mold and shaped into a cylinder with a cold press method. The as-prepared cylinder was inserted into an h-BN crucible and loaded into a high pressure apparatus for HPS experiments. The first step of HPS was carried out at 3 GPa and 1100 K for 0.5 h. The intermediate product was ground into powders under argon protection, shaped and subject to the second step of HPS (3 GPa and 873 K for 3 h). The obtained product was ground into powders, washed in acidic solution to remove excess Ce, and sintered into dense pellets by using spark plasma sintering (SPS) at 60 MPa and 853 K for 15 min. Rectangular blocks (2 × 2 × 8 mm3), disks (ø 6 mm × 1 mm), and slices (2.5 × 6 × 0.4 mm3) were cut from the final pellets for the electrical transport, thermal conductivity, and Hall coefficient measurements, respectively.
The chemical composition and crystal structure of the final products were characterized with electron probe microanalysis (EPMA, JEOL JXA-8230, JEOL, Tokyo, Japan) and X-ray diffraction (XRD, Rigaku D/MAX/2500/PC, Rigaku, Tokyo, Japan), respectively. The Seebeck coefficient and electrical resistivity were measured with a ZEM-3 apparatus (Ulvac-Riko, Yokohama, Japan) and the thermal conductivity was measured with a TC-9000H apparatus (Ulvac-Riko, Yokohama, Japan). The uncertainty in S, ρ, and κ measurements is within 5%. The Hall coefficient measurements were performed at room temperature with a Physical Property Measurement System (PPMS, Quantum Design, San Diego, CA, USA). The heat capacity measurements were carried out between 5 and 25 K with the heat capacity option of PPMS.

3. Results and Discussion

XRD patterns from CeyFe4−xCoxSb12 samples are shown in Figure 1. All these samples are dominated with a skutterudite structure of I m 3 ¯ symmetry. XRD results also reveal trace amounts of FeSb2 and Sb, which are less than 5% as determined by EPMA analyses. A systematical shift of diffraction peaks to higher diffraction angles occurs with increasing Co content, as illustrated in Figure 1b for the magnified (013) peak, indicating that the lattice parameter decreases with elevating Co level. The actual content of Co (x) and filling fraction of Ce (y) in CeyFe4−xCoxSb12 samples determined by EPMA measurements are listed in Table 1. The highest filling fraction of Ce of 0.92 is reached in the sample without Co substitution (Ce0.92Fe4Sb12), which is at the same level as that of 0.91 achieved in an ambient-pressure synthesized sample [11]. Cobalt atom has one more valence electron and smaller atomic size than those of iron atoms. By substituting Co on Fe sites, extra electrons from Co are introduced to the framework meanwhile the lattice parameter is reduced. Both effects contribute adversely to the void-filling process of Ce. Therefore, the higher Co substitution level, and the lower Ce filling fraction [24], and both contribute to a smaller lattice parameter. The actual filling fraction of Ce in our HPS samples decreases slightly with increasing Co substitution level, and is higher than that in the ambient-pressure synthesized sample with the same Co level [24], demonstrating an advantage of high pressure for achieving higher filling fraction.
Room temperature electrical transport properties for different samples are summarized in Table 1. All samples exhibit positive Seebeck coefficient, indicating the majority charge carriers of holes. The hole concentration deduced from the measured Hall coefficient decreases notably with elevating Co substitution level, which is due to Co substitution introducing extra electrons into the system. In addition, the hole concentration in our Ce0.85Fe3.03Co0.97Sb12 sample is 9.5 × 1020 cm−3, less than half of that in the ambient-pressure synthesized Ce0.9Fe3CoSb12 (nominal composition) [21]. The difference can be attributed to the higher filling fraction of Ce in our HPS sample, and accounts for the significant increase in the electrical resistivity of our HPS Ce0.85Fe3.03Co0.97Sb12 sample. The hole concentrations can also be calculated by simple charge counting based on the chemical compositions, which differ from those listed in Table 1. A similar difference was previously reported [11,32], and might be related to the multiple bands near the Fermi level in elemental-filled FeSb3 materials where a simple relationship between carrier concentration and Hall coefficient does not exist. Note that the carrier mobility is continually suppressed with increasing Co level, indicating a decreased hole mean free path caused by impurity scattering.
The temperature dependent thermoelectric properties of CeyFe4−xCoxSb12 samples are shown in Figure 2. The electrical resistivity increases with elevating Co substitution level (Figure 2a), which can be attributed to the decreased hole concentration compensated by the additional electrons from cobalt. The most Co-substituted Ce0.85Fe3.03Co0.97Sb12 exhibits the largest electrical resistivity, which starts to decrease at high temperature due to thermally activated carriers. The other two samples with lower Co substitution level show a heavily doped semiconductor behavior and the resistivities increase with elevating temperature due to stronger scattering. As the hole concentration decreases with higher Co substitution level, an increase in the Seebeck coefficient occurs (Figure 2b). While the most Co-substituted Ce0.85Fe3.03Co0.97Sb12 shows an early onset of intrinsic regime of conduction occurring at ca. 700 K, the Seebeck coefficient for the other two samples increases monotonically with elevating temperature. The tendencies of the Seebeck coefficient with respect to temperature and Co substitution level are consistent with those of the electrical resistivity.
The temperature dependent power factor is presented in Figure 2c. Compared with the increase of S as a function of Co substitution level, the increase of ρ is more significant, leading to decreased peak values of PF with increasing Co substitution level. The maximal PF of 3200 µW m−1 K−2 is achieved for Ce0.92Fe4Sb12, which is comparable to the best values of other single elemental filled FeSb3 skutterudites [11]. We note a remarkably reduced PF in Ce0.85Fe3.03Co0.97Sb12 sample. In this sample, the actual filling fraction of Ce is higher than that in the ambient-pressure synthesized sample with the same Co level [24], leading to a lower hole concentration and a significantly increased electrical resistivity.
The temperature dependence of thermal conductivity is shown in Figure 2d for CeyFe4−xCoxSb12 samples. The variation tendency of κ with respect to temperature is similar to that of CeyFe4-(x/2)Nix/2Sb12 samples [24]. The total thermal conductivity is a sum of the lattice thermal conductivity (κL) and the carrier contribution κc,
κ = κL + κc
κc can be evaluated with the Wiedemann−Franz law,
κc = LT
where L is the Lorenz number and a value of 2 × 10−8 WΩ K−2 is used here [33]. The temperature dependent κL can then be estimated and is shown as the inset of Figure 2d. The minimum κL of 0.71 Wm−1 K−1 is reached in Ce0.92Fe4Sb12 at 700 K. For all the samples, κL first decreases with increasing temperature, and then increases slightly at high temperature due to the bipolar effect. The onset of bipolar effect shifts to higher temperature for the sample with lower Co substitution level, which can be attributed to the higher hole concentration suppressing the bipolar effect. For CeyFe4−xCoxSb12 samples, it is plausible that both impurity atoms (Co) and rattling atoms (Ce) contribute to phonon scattering and reduce κL. However, κL increases with elevating Co substitution level. Similar results were previously observed in Ni-doped CeFe4Sb12 systems [24]. These results indicate that impurity scattering of phonon due to mass/force constant fluctuation of Fe and Co is less significant for our samples, and the reduction of κL is dominated by the rattling fillers: higher filling fraction leads to lower κL.
One of the effective techniques to confirm such kinds of localized vibrational modes is the low temperature heat capacity (CP) measurement [34,35,36], which was performed on Ce0.92Fe4Sb12 in the temperature range of 5−25 K. The heat capacity as a function of temperature is well described by the sum of an electronic term, γT, and phonon terms, which are the Einstein term for the loosely bonded filler atom and the Debye T3 term for the 16 atoms of the framework. The extra Debye T5 term can be neglected at low temperature [32,37]. As shown in Figure 3, CP is well fitted by
C P = γ T + β T 3 + A x 2 e x / ( e x 1 ) 2 ,
where γ = 0.281 J K−2 mol−1, β = 0.00159 J K−4 mol−1, (corresponding to a Debye temperature of 269 K), A = 22.05 J K−1 mol−1, and x = ΘE/T with an Einstein temperature ΘE = 89.4 K. The fitting quality is significantly good with χ2 = 5.62 × 10−3 and R2 = 0.99996. The Einstein temperature determined in this work is comparable with that (86 K) reported previously [38]. In addition, the filling fraction determined from A/3R (R is the ideal gas constant) is 0.88, which is very close to the value of 0.92 from EPMA measurement and indicates successful filling of Ce atoms into FeSb3 voids.
Figure 4 shows ZT of CeyFe4−xCoxSb12 samples calculated from the measured thermoelectric properties. For the optimal Ce0.92Fe4Sb12 sample, ZT increases with temperature and reaches the highest value of 0.91 at 763 K, which is slightly higher than that achieved in ambient-pressure synthesized sample [11]. It is noted the thermoelectric performance of our HPS CeyFe4−xCoxSb12 samples deteriorates with increasing Co substitution level, which differs from those ambient-pressure synthesized samples. For example, the highest ZT of 0.9 was achieved in Ce0.9Fe3CoSb12 [21] and CeFe3.8Co0.2Sb12 [22]. In our samples, the actual filling fraction of Ce only slightly decreases with increasing Co content, which leads to a much smaller hole concentration and an undesirable increase in ρ compared with those ambient-pressure synthesized samples. Consequentially, an enhancement in thermoelectric performance does not occur through substituting framework atoms with electron-richer atoms alone in our HPS samples.
Nonetheless, HPS methods employed in this study demonstrates an advantage over the solid state reaction method at ambient pressure. HPS significantly accelerates the fabrication process and shortens the total reaction duration to several hours, and allows us to tune the filling fraction in a wider range. In addition, the synthetic pressure and temperature conditions can easily be realized in industrial production [39], another advantage for massive production for thermoelectric applications. Combined with the strategies of multiple filling/substituting and nanostructuring (through, e.g., ball milling and severe plastic deformation) [9,10,40,41], further enhancement of ZT values of p-type skutterudites could be realized with the HPS method.

4. Conclusions

p-type CeyFe4−xCoxSb12 skutterudites were successfully synthesized by using an efficient HPS method within several hours. The highest filling fraction of Ce (y = 0.92) was reached in the Co-free sample. With increasing Co substitution level, both hole concentration and Ce filling fraction decrease. The enhancement in S due to a smaller hole concentration is offset by the increase in ρ. Moreover, the suppression of κL seems dominated by the rattling filler atoms, and lower filling fraction of Ce with higher Co content is unfavorable to reduce κL. The maximal ZT of 0.91 was achieved in the optimal Ce0.92Fe4Sb12 sample, and the thermoelectric performance of CeyFe4−xCoxSb12 deteriorates with higher Co content.

Acknowledgments

This work was supported by the National Science Foundation of China (51172196, 51201149, 51332005, 51421091, and 51525205), and the Natural Science Foundation for Distinguished Young Scholars of Hebei Province of China (E2014203150).

Author Contributions

Bo Xu conceived the project; Yadi Liu, Long Zhang, Yongjun Tian, and Bo Xu designed the experiments; Yadi Liu, Xiaohui Li, Qian Zhang, and Dongli Yu performed the HPS experiments, Yadi Liu, Xiaohui Li, Qian Zhang, and Long Zhang performed the measurements; Yadi Liu, Long Zhang, Dongli Yu, Yongjun Tian, and Bo Xu analyzed the data; Yadi Liu, Long Zhang, and Bo Xu co-wrote the paper; All authors discussed the results and commented on the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bell, L.E. Cooling, heating, generating power, and recovering waste heat with thermoelectric systems. Science 2008, 321, 1457–1461. [Google Scholar] [CrossRef] [PubMed]
  2. Slack, G.A. New Materials and Performance Limits for Thermoeletric Cooling. In CRC Handbook of Thermoelectrics; Rowe, D.M., Ed.; CRC Press: Boca Raton, FL, USA, 1995; pp. 407–440. [Google Scholar]
  3. Rull-Bravo, M.; Moure, A.; Fernandez, J.F.; Martin-Gonzalez, M. Skutterudites as thermoelectric materials: Revisited. RSC Adv. 2015, 5, 41653–41667. [Google Scholar] [CrossRef]
  4. Pei, Y.Z.; Yang, J.; Chen, L.D.; Zhang, W.; Salvador, J.R.; Yang, J. Improving thermoelectric performance of caged compounds through light-element filling. Appl. Phys. Lett. 2009, 95, 042101. [Google Scholar] [CrossRef]
  5. Shi, X.; Yang, J.; Salvador, J.R.; Chi, M.; Cho, J.Y.; Wang, H.; Bai, S.; Yang, J.; Zhang, W.; Chen, L. Multiple-Filled Skutterudites: High Thermoelectric Figure of Merit through Separately Optimizing Electrical and Thermal Transports. J. Am. Chem. Soc. 2011, 133, 7837–7846. [Google Scholar] [CrossRef] [PubMed]
  6. Rogl, G.; Grytsiv, A.; Rogl, P.; Peranio, N.; Bauer, E.; Zehetbauer, M.; Eibl, O. n-Type skutterudites (R, Ba, Yb)yCo4Sb12 (R = Sr, La, Mm, DD, SrMm, SrDD) approaching ZT ≈ 2.0. Acta Mater. 2014, 63, 30–43. [Google Scholar] [CrossRef]
  7. Liu, R.; Yang, J.; Chen, X.; Shi, X.; Chen, L.; Uher, C. p-Type skutterudites RxMyFe3CoSb12 (R, M = Ba, Ce, Nd, and Yb): Effectiveness of double-filling for the lattice thermal conductivity reduction. Intermetallics 2011, 19, 1747–1751. [Google Scholar] [CrossRef]
  8. Yang, J.; Meisner, G.P.; Rawn, C.J.; Wang, H.; Chakoumakos, B.C.; Martin, J.; Nolas, G.S.; Pedersen, B.L.; Stalick, J.K. Low temperature transport and structural properties of misch-metal-filled skutterudites. J. Appl. Phys. 2007, 102, 083702. [Google Scholar] [CrossRef]
  9. Rogl, G.; Grytsiv, A.; Rogl, P.; Bauer, E.; Zehetbauer, M. A new generation of p-type didymium skutterudites with high ZT. Intermetallics 2011, 19, 546–555. [Google Scholar] [CrossRef]
  10. Rogl, G.; Grytsiv, A.; Heinrich, P.; Bauer, E.; Kumar, P.; Peranio, N.; Eibl, O.; Horky, J.; Zehetbauer, M.; Rogl, P. New bulk p-type skutterudites DD0.7Fe2.7Co1.3Sb12−xXx (X = Ge, Sn) reaching ZT > 1.3. Acta Mater. 2015, 91, 227–238. [Google Scholar] [CrossRef]
  11. Qiu, P.F.; Yang, J.; Liu, R.H.; Shi, X.; Huang, X.Y.; Snyder, G.J.; Zhang, W.; Chen, L.D. High-temperature electrical and thermal transport properties of fully filled skutterudites RFe4Sb12 (R = Ca, Sr, Ba, La, Ce, Pr, Nd, Eu, and Yb). J. Appl. Phys. 2011, 109, 063713. [Google Scholar] [CrossRef]
  12. Dong, Y.; Puneet, P.; Tritt, T.M.; Nolas, G.S. High-temperature thermoelectric properties of p-type skutterudites Ba0.15YbxCo3FeSb12 and YbyCo3FeSb9As3. J. Mater. Sci. 2015, 50, 34–39. [Google Scholar] [CrossRef]
  13. Caillat, T.; Kulleck, J.; Borshchevsky, A.; Fleurial, J.P. Preparation and thermoelectric properties of the skutterudite-related phase Ru0.5Pd0.5Sb3. J. Appl. Phys. 1996, 79, 8419–8426. [Google Scholar] [CrossRef]
  14. Anno, H.; Matsubara, K.; Notohara, Y.; Sakakibara, T.; Tashiro, H. Effects of doping on the transport properties of CoSb3. J. Appl. Phys. 1999, 86, 3780–3786. [Google Scholar] [CrossRef]
  15. Liu, W.S.; Zhang, B.P.; Li, J.F.; Zhang, H.L.; Zhao, L.D. Enhanced thermoelectric properties in CoSb3-xTex alloys prepared by mechanical alloying and spark plasma sintering. J. Appl. Phys. 2007, 102, 103717. [Google Scholar] [CrossRef]
  16. Su, X.L.; Li, H.; Wang, G.Y.; Chi, H.; Zhou, X.Y.; Tang, X.F.; Zhang, Q.J.; Uher, C. Structure and transport properties of double-doped CoSb2.75Ge0.25–xTex (x = 0.125–0.20) with in situ nanostructure. Chem. Mater. 2011, 23, 2948–2955. [Google Scholar] [CrossRef]
  17. Kim, I.H.; Ur, S.C. Electronic transport properties of Fe-doped CoSb3 prepared by encapsulated induction melting. Mater. Lett. 2007, 61, 2446–2450. [Google Scholar] [CrossRef]
  18. Katsuyama, S.; Shichijo, Y.; Ito, M.; Majima, K.; Nagai, H. Thermoelectric properties of the skutterudite Co1−xFexSb3 system. J. Appl. Phys. 1998, 84, 6708–6712. [Google Scholar] [CrossRef]
  19. Duan, B.; Zhai, P.; Xu, C.; Ding, S.; Li, P.; Zhang, Q. Thermoelectric performance of tellurium and sulfur double-substituted skutterudite materials. J. Mater. Sci. 2014, 49, 4445–4452. [Google Scholar] [CrossRef]
  20. Sales, B.C.; Mandrus, D.; Chakoumakos, B.C.; Keppens, V.; Thompson, J.R. Filled skutterudite antimonides: Electron crystals and phonon glasses. Phys. Rev. B 1997, 56, 15081–15089. [Google Scholar] [CrossRef]
  21. Liu, R.; Qiu, P.; Chen, X.; Huang, X.; Chen, L. Composition optimization of p-type skutterudites CeyFexCo4−xSb12 and YbyFexCo4−xSb12. J. Mater. Res. 2011, 26, 1813–1819. [Google Scholar] [CrossRef]
  22. Tan, G.J.; Wang, S.Y.; Yan, Y.G.; Li, H.; Tang, X.F. Effects of Cobalt Substitution for Fe on the Thermoelectric Properties of p-Type CeFe4−xCoxSb12 Skutterudites. J. Electron. Mater. 2012, 41, 1147–1152. [Google Scholar] [CrossRef]
  23. Tan, G.J.; Wang, S.Y.; Tang, X.F. Thermoelectric Performance Optimization in p-Type CeyFe3CoSb12 Skutterudites. J. Electron. Mater. 2014, 43, 1712–1717. [Google Scholar] [CrossRef]
  24. Qiu, P.F.; Liu, R.H.; Yang, J.; Shi, X.; Huang, X.Y.; Zhang, W.; Chen, L.D.; Yang, J.; Singh, D.J. Thermoelectric properties of Ni-doped CeFe4Sb12 skutterudites. J. Appl. Phys. 2012, 111, 023705. [Google Scholar] [CrossRef]
  25. Yang, J.Q.; Zhang, L.; Liu, Y.D.; Chen, C.; Li, J.H.; Yu, D.L.; He, J.L.; Liu, Z.Y.; Tian, Y.J.; Xu, B. Investigation of skutterudite MgyCo4Sb12: High pressure synthesis and thermoelectric properties. J. Appl. Phys. 2013, 113, 113703. [Google Scholar] [CrossRef]
  26. Zhang, J.J.; Xu, B.; Wang, L.M.; Yu, D.L.; Liu, Z.Y.; He, J.L.; Tian, Y.J. Great thermoelectric power factor enhancement of CoSb3 through the lightest metal element filling. Appl. Phys. Lett. 2011, 98, 072109. [Google Scholar] [CrossRef]
  27. Zhang, J.J.; Xu, B.; Wang, L.M.; Yu, D.L.; Yang, J.Q.; Yu, F.R.; Liu, Z.Y.; He, J.L.; Wen, B.; Tian, Y.J. High-pressure synthesis of phonon-glass electron-crystal featured thermoelectric LixCo4Sb12. Acta Mater. 2012, 60, 1246–1251. [Google Scholar] [CrossRef]
  28. Deng, L.; Wang, L.B.; Jia, X.P.; Ma, H.A.; Qin, J.M.; Wan, Y.C. Improvement of thermoelectric performance for Te-doped CoSb3 by higher synthesis pressure. J. Alloys Compd. 2014, 602, 117–121. [Google Scholar] [CrossRef]
  29. Deng, L.; Jia, X.P.; Su, T.C.; Jiang, Y.P.; Zheng, S.Z.; Guo, X.; Ma, H.A. The thermoelectric properties of Co4Sb12-xTex synthesized at different pressure. Mater. Lett. 2011, 65, 1057–1059. [Google Scholar] [CrossRef]
  30. Zhang, Q.; Li, X.H.; Kang, Y.L.; Zhang, L.; Yu, D.L.; He, J.L.; Liu, Z.Y.; Tian, Y.J.; Xu, B. High pressure synthesis of Te-doped CoSb3 with enhanced thermoelectric performance. J. Mater. Sci. Mater. Electron. 2015, 26, 385–391. [Google Scholar] [CrossRef]
  31. Badding, J.V. High-pressure synthesis, characterization, and tuning of solid state materials. Annu Rev. Mater. Sci. 1998, 28, 631–658. [Google Scholar] [CrossRef]
  32. Schnelle, W.; Leithe-Jasper, A.; Rosner, H.; Cardoso-Gil, R.; Gumeniuk, R.; Trots, D.; Mydosh, J.; Grin, Y. Magnetic, thermal, and electronic properties of iron-antimony filled skutterudites MFe4Sb12 (M = Na, K, Ca, Sr, Ba, La, Yb). Phys. Rev. B 2008, 77, 094421. [Google Scholar] [CrossRef]
  33. Dyck, J.S.; Chen, W.; Uher, C.; Chen, L.; Tang, X.; Hirai, T. Thermoelectric properties of the n-type filled skutterudite Ba0.3Co4Sb12 doped with Ni. J. Appl. Phys. 2002, 91, 3698–3705. [Google Scholar] [CrossRef]
  34. Dimitrov, I.K.; Manley, M.E.; Shapiro, S.M.; Yang, J.; Zhang, W.; Chen, L.D.; Jie, Q.; Ehlers, G.; Podlesnyak, A.; Camacho, J.; Li, Q. Einstein modes in the phonon density of states of the single-filled skutterudite Yb0.2Co4Sb12. Phys. Rev. B 2010, 82, 174301. [Google Scholar] [CrossRef]
  35. Hermann, R.; Jin, R.; Schweika, W.; Grandjean, F.; Mandrus, D.; Sales, B.; Long, G. Einstein oscillators in thallium filled antimony skutterudites. Phys. Rev. Lett. 2003, 90, 135505. [Google Scholar] [CrossRef] [PubMed]
  36. Keppens, V.; Mandrus, D.; Sales, B.C.; Chakoumakos, B.C.; Dai, P.; Coldea, R.; Maple, M.B.; Gajewski, D.A.; Freeman, E.J.; Bennington, S. Localized vibrational modes in metallic solids. Nature 1998, 395, 876–878. [Google Scholar]
  37. Schnelle, W.; Leithe-Jasper, A.; Schmidt, M.; Rosner, H.; Borrmann, H.; Burkhardt, U.; Mydosh, J.; Grin, Y. Itinerant iron magnetism in filled skutterudites CaFe4Sb12 and YbFe4Sb12: Stable divalent state of ytterbium. Phys. Rev. B 2005, 72, 020402. [Google Scholar] [CrossRef]
  38. Cao, D.; Bridges, F.; Chesler, P.; Bushart, S.; Bauer, E.; Maple, M. Evidence for rattling behavior of the filler atom (L) in the filled skutterudites LT4X12 (L = Ce, Eu, Yb; T = Fe, Ru; X = P, Sb) from EXAFS studies. Phys. Rev. B 2004, 70, 094109. [Google Scholar] [CrossRef]
  39. Brazhkin, V.V. High-pressure synthesized materials: Treasures and hints. High Press. Res. 2007, 27, 333–351. [Google Scholar] [CrossRef]
  40. Rogl, G.; Grytsiv, A.; Rogl, P.; Bauer, E.; Kerber, M.B.; Zehetbauer, M.; Puchegger, S. Multifilled nanocrystalline p-type didymium-Skutterudites with ZT > 1.2. Intermetallics 2010, 18, 2435–2444. [Google Scholar] [CrossRef]
  41. Rogl, G.; Grytsiv, A.; Rogl, P.; Royanian, E.; Bauer, E.; Horky, J.; Setman, D.; Schafler, E.; Zehetbauer, M. Dependence of thermoelectric behaviour on severe plastic deformation parameters: A case study on p-type skutterudite DD0.60Fe3CoSb12. Acta Mater. 2013, 61, 6778–6789. [Google Scholar] [CrossRef]
Figure 1. (a) XRD patterns (Cu Kα) for CeyFe4−xCoxSb12 samples; (b) the magnified (013) peaks emphasizing the continuous lattice parameter contraction with elevating Co substitution level.
Figure 1. (a) XRD patterns (Cu Kα) for CeyFe4−xCoxSb12 samples; (b) the magnified (013) peaks emphasizing the continuous lattice parameter contraction with elevating Co substitution level.
Materials 09 00257 g001
Figure 2. Temperature dependent thermoelectric properties of CeyFe4−xCoxSb12 samples. (a) electrical resistivity ρ; (b) the Seebeck coefficient S; (c) power factor PF; and (d) thermal conductivity κ. The inset to panel d shows the lattice thermal conductivity κL.
Figure 2. Temperature dependent thermoelectric properties of CeyFe4−xCoxSb12 samples. (a) electrical resistivity ρ; (b) the Seebeck coefficient S; (c) power factor PF; and (d) thermal conductivity κ. The inset to panel d shows the lattice thermal conductivity κL.
Materials 09 00257 g002
Figure 3. Low-temperature molar heat capacity of Ce0.92Fe4Sb12 (see main text for the fitting details).
Figure 3. Low-temperature molar heat capacity of Ce0.92Fe4Sb12 (see main text for the fitting details).
Materials 09 00257 g003
Figure 4. ZT as a function of temperature for CeyFe4−xCoxSb12 samples.
Figure 4. ZT as a function of temperature for CeyFe4−xCoxSb12 samples.
Materials 09 00257 g004
Table 1. Co content (x), Ce filling fraction (y), and room temperature lattice constant (a), Seebeck coefficient (S), electrical resistivity (ρ), carrier concentration (p), and carrier mobility (μ) for HPS CeyFe4−xCoxSb12 samples.
Table 1. Co content (x), Ce filling fraction (y), and room temperature lattice constant (a), Seebeck coefficient (S), electrical resistivity (ρ), carrier concentration (p), and carrier mobility (μ) for HPS CeyFe4−xCoxSb12 samples.
xya (nm)S (μV K−1)ρ (μΩ m)p (1020 cm−3)μ (cm2 V−1 s−1)
00.92 (2)0.9167 (3)76.25.021.55.8
0.49 (1)0.88 (2)0.9153 (3)84.36.917.15.3
0.97 (2)0.85 (2)0.9140 (3)99.414.39.54.6

Share and Cite

MDPI and ACS Style

Liu, Y.; Li, X.; Zhang, Q.; Zhang, L.; Yu, D.; Xu, B.; Tian, Y. High Pressure Synthesis of p-Type CeyFe4−xCoxSb12 Skutterudites. Materials 2016, 9, 257. https://doi.org/10.3390/ma9040257

AMA Style

Liu Y, Li X, Zhang Q, Zhang L, Yu D, Xu B, Tian Y. High Pressure Synthesis of p-Type CeyFe4−xCoxSb12 Skutterudites. Materials. 2016; 9(4):257. https://doi.org/10.3390/ma9040257

Chicago/Turabian Style

Liu, Yadi, Xiaohui Li, Qian Zhang, Long Zhang, Dongli Yu, Bo Xu, and Yongjun Tian. 2016. "High Pressure Synthesis of p-Type CeyFe4−xCoxSb12 Skutterudites" Materials 9, no. 4: 257. https://doi.org/10.3390/ma9040257

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