Next Article in Journal
Microstructural and Mechanical Characteristics of Alkali-Activated Binders Composed of Milled Fly Ash and Granulated Blast Furnace Slag with µ-Limestone Addition
Previous Article in Journal
New Trends in Separation Techniques of Lithium Isotopes: A Review of Chemical Separation Methods
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Investigation of the Effect of Double-Filler Atoms on the Thermoelectric Properties of Ce-YbCo4Sb12

1
Advanced Materials Convergence R&D Division, Korea Institute of Ceramic Engineering and Technology (KICET), Jinju 52851, Republic of Korea
2
Faculty of Fundamental Sciences, Phenikaa University, Yen Nghia, Ha-Dong District, Hanoi 10000, Vietnam
3
New & Renewable Energy Power Generation Department, Korea South-East Power Co. (KOEN), Yeongcheon 38837, Republic of Korea
4
School of Nano & Materials Science and Engineering, Kyungpook National University, Sangju 37224, Republic of Korea
5
Department of Materials Convergence and System Engineering, School of Materials Science and Engineering, Changwon National University, Changwon 51140, Republic of Korea
6
Device Research Center, Samsung Electronics, Suwon 16678, Republic of Korea
7
Department of Electronic Materials Engineering, Kwangwoon University, Seoul 01897, Republic of Korea
*
Authors to whom correspondence should be addressed.
Materials 2023, 16(10), 3819; https://doi.org/10.3390/ma16103819
Submission received: 1 April 2023 / Revised: 13 May 2023 / Accepted: 15 May 2023 / Published: 18 May 2023
(This article belongs to the Topic Thermoelectric Energy Harvesting)

Abstract

:
Skutterudite compounds have been studied as potential thermoelectric materials due to their high thermoelectric efficiency, which makes them attractive candidates for applications in thermoelectric power generation. In this study, the effects of double-filling on the thermoelectric properties of the CexYb0.2−xCo4Sb12 skutterudite material system were investigated through the process of melt spinning and spark plasma sintering (SPS). By replacing Yb with Ce, the carrier concentration was compensated for by the extra electron from Ce donors, leading to optimized electrical conductivity, Seebeck coefficient, and power factor of the CexYb0.2−xCo4Sb12 system. However, at high temperatures, the power factor showed a downturn due to bipolar conduction in the intrinsic conduction regime. The lattice thermal conductivity of the CexYb0.2−xCo4Sb12 skutterudite system was clearly suppressed in the range between 0.025 and 0.1 for Ce content, due to the introduction of the dual phonon scattering center from Ce and Yb fillers. The highest ZT value of 1.15 at 750 K was achieved for the Ce0.05Yb0.15Co4Sb12 sample. The thermoelectric properties could be further improved by controlling the secondary phase formation of CoSb2 in this double-filled skutterudite system.

1. Introduction

Thermoelectricity is one of the outstanding candidates for waste heat recovery, providing an energy solution for collecting waste heat and converting it to electricity. The performance of thermoelectric materials is evaluated by the figure of merit, ZT, which is denoted as ZT = (S2σT)/k, where S is the Seebeck coefficient, σ is the electrical conductivity, k is the total thermal conductivity composed of lattice and electronic components, and T is the absolute temperature. A high ZT value can be achieved with high electrical conductivity, a large Seebeck coefficient, and low thermal conductivity [1,2,3]. However, each parameter in the formula above has its own correlations with others in terms of carrier concentration [2,4]. For example, increasing the carrier concentration increases the electrical conductivity but decreases the Seebeck coefficient, and increases the electronic thermal conductivity, resulting in an increase in total thermal conductivity [5]. Hence, optimization of the carrier concentration of TE materials is needed to achieve high thermoelectric efficiency.
Each thermoelectric material can be adapted to different operating temperature ranges and requires good properties, such as high thermal stability, high chemical stability, low toxicity of the elements, raw material availability, and, particularly, high performance for practical applications [6]. Among various promising materials for high thermoelectric energy conversion efficiency, skutterudites (SKD) based on Co4Sb12 materials are the best candidates for power generation applications operating in the intermediate temperature range; they are well matched to the above requirements, including good mechanical properties [6,7,8,9]. Intrinsic Co4Sb12-based SKD materials have good electrical properties, such as a high Seebeck coefficient and electrical conductivity, and large carrier mobility [10]. However, the relatively high thermal conductivity of this system results in an overall low thermoelectric performance. Despite this unfavorable feature of thermal conductivity, SKD material systems have an interesting feature; there are two “voids” in their unit cells that can be filled by guest atoms [10]. This affects the charge properties by providing carrier concentration, strongly scattering phonons, and reducing the lattice thermal conductivity [11,12]. The filler atom in the SKD structure not only contributes electrons without changing the band structure [13,14,15], but also significantly reduces the lattice thermal conductivity by loosely binding the atom in the cage, leading to effective phonon scattering [16,17]. A single Yb-filled SKD was found to have an excellent ZT of 1.5 at 850 K, and low lattice thermal conductivity (of less than 1.0 Wm−1K−1) [18]. However, only phonons that have a frequency of resonance close to the rattle frequency will interact and scatter. Therefore, double filling could strongly induce different filler atom vibration frequencies separately, aiming to extend the spectra of resonant phonon scattering [17]. Moreover, Schnelle et al. reported that a lighter atom and smaller ionic radii are more strongly rattled than heavier and large atoms [19]. For enhancing thermoelectric performance, a couple of atoms were inserted into the SKD structure, such as Yb-In (ZT = 0.97 at 750 K, and a lowest lattice of 1.7 Wm−1K−1) [20], Ba-Yb (ZT = 1.36 at 800 K, klat = 0.9 Wm−1K−1) [21], Ca-Yb (ZT = 1.03 at 750 K, klat = 0.7 Wm−1K−1) [16], with promising results. Ce, similarly to Yb, belongs to rare earth metal group and has smaller size of ion, It is lighter than Yb [22], and can be used as a candidate for reducing lattice thermal conductivity through double filling. However, most of these reported experiments use long-lasting heat treatment to synthesize samples. The SKD materials were synthesized using a rapid method that has been published for more than a decade [23,24,25,26,27,28,29], with a shortened processing time of less than 2 days, with a combination of melt spinning and spark plasma sintering [23]. Melt spinning not only saves time, but also refines the nanostructure, which can increase phonon scattering [30,31,32], resulting in a reduction in thermal conductivity and an increase in thermoelectric efficiency. With these practical advantages for synthesizing SKD materials, along with their good thermoelectric properties, their impact on the industry is positive. SKD materials can be used to produce more efficient thermoelectric generators, which can convert waste heat from industrial processes into electricity in the high-temperature range. This can help to reduce the carbon footprint of many industries and make them more sustainable. Besides, replacing Yb with Ce is potentially a commercial option, because Yb is a rare metal (3 ppm), whereas Ce is a relatively abundant element (68 ppm) [33]. Moreover, the price of Ce is lower than that of Yb (Ce: $5000 per ton; Yb: $50,000 per ton) [34]. In this study, the double filling effect was investigated in a series of CexYb0.2−xCo4Sb12 (x = 0, 0.025, 0.05, 0.075, 0.1) samples, with the total filling fraction of Ce-Yb fixed at 0.2, using rapid synthesis of melt spinning followed by spark plasma sintering (SPS).

2. Materials and Methods

A series of double-filled samples, CexYb0.2−xCo4Sb12 (x = 0, 0.025, 0.05, 0.075, 0.1), were prepared using melt spinning followed by a spark plasma sintering method. The raw materials used in this experiment were Ce (Alfa Aesar, ingot, 99.8%), Co (Alfa Aesar, slug, 99.95%), Sb (TTS chemical, slug, 99.9%), and Yb (Alfa Aesar, powder, 99.9%). The chemicals were weighed according to their stoichiometric ratios and placed into carbon-coated quartz tubes, then sealed under a vacuum and subjected to induction melting for 15 min under a power supply of 20 W. The ingots obtained were loaded into cylinder graphite crucibles with an inner diameter of 0.4 mm, for a rapid solidification process (RSP) which was achieved using the melt spinning method under a power supply of 11 kW and a Cu-wheel speed of 2000 rpm (~26.2 m/s) in an argon atmosphere. The ribbons obtained after melt spinning were collected and hand-ground in an argon-filled glovebox to obtain fine powders. The powders were then consolidated using an SPS process at 963 K for 15 min under a pressure of 50 MPa. The relative densities for all samples studied here were measured on the sintered ones, and found to be greater than 95% of the theoretical values.
X-ray diffraction (XRD) data were collected using a Bruker D8 Advance machine (20 kV, 5 mA) with Cu Kα radiation (λ = 1.5418 Å). The microstructure of the sintered samples was analyzed using a scanning electron microscope (SEM, Verios 460 L, FEI, 10 kV for FESEM, 20 kV for EDS). The electrical conductivity and Seebeck coefficient were measured using a commercial ZEM 3 (ULVAC-RIKO) from room temperature to 773 K on rectangular bar-shaped samples with dimensions of approximately 3 mm × 3 mm × 10 mm. The carrier concentration and carrier mobility measurements at room temperature were performed using a Hall measurement system (ResiTest 8400, Tokyo Corporation). The measurements of thermal diffusivity were carried out in a flowing argon atmosphere using laser flash analysis (Laser Flash DLF-1) between 300 K and 773 K, and the thermal conductivity was calculated from the relationship k = αρCp, where α is the thermal diffusivity, ρ is the density, and Cp is the specific heat capacity. The value of 0.23 Jg−1K−1 for Cp is used to calculate the thermal conductivity for all temperatures due to its weak dependence on temperature [18,35].

3. Results and Discussion

The XRD patterns of a series of CexYb0.2−xCo4Sb12 (x = 0, 0.025, 0.05, 0.075, and 0.1) samples are displayed in Figure 1a. It is shown that the XRD patterns exhibit good crystalline SKD, with diffraction peaks indexed to the SKD structure corresponding to JCPDS #00-065-1791 (body-centered cubic skutterudite phase, space group Im-3). However, small amounts of the secondary phase of CoSb2 are clearly observed around 2θ = 32.2° for every composition, as shown in Figure 1b. The secondary phase CoSb2 is normally formed in skutterudite [23,36,37], especially in melt spinning, because the resulting ribbons contain a mixture of main material and secondary phase [38]. The intensities of the secondary peaks increase with increasing Ce content, and CoSb2 secondary peaks can easily be found in the x = 0.1 sample. Focusing on the main SKD peaks (013) around 2θ = 31°, there is a shift of the peaks toward a higher 2θ angle with increasing Ce content (Figure 1c), which suggests that the lattice constants for CexYb0.2−xCo4Sb12 (x = 0, 0.025, 0.05, 0.075, and 0.1) are reduced. To understand the effect of Ce substitution on the reduction of the lattice parameter, the lattice constant was calculated, and the results are displayed as a function of Ce content in Figure 1d. It was found that the lattice constant was reduced from 9.052 Å for sample x = 0, to 9.041 Å for sample x = 0.10 (Table 1). This reduction can be explained by the Ce ion’s preference for a trivalent state, whereas Yb prefers to be divalent in the skutterudite structure [39,40,41]. Thus, the ionic radius of Ce3+ (103 pm) is smaller than that of Yb2+ (113 pm) [22]. Therefore, by replacing Yb atoms with Ce substitution atoms, the lattice constant could be decreased, and follows Vegard’s law in all samples, as shown in Figure 1d.
The FE-SEM images of the cross-section surface of SPSed bulk samples CexYb0.2−xCo4Sb12 (x = 0, 0.025, 0.05, 0.075, and 0.1) are shown in Figure 2. As shown in Figure 2a–e, the samples with x = 0~0.075 have grain sizes varying between 1 μm and 3 µm. However, in the samples with x = 0.1, as seen in Figure 2e, there are small grains with sizes below 1 μm. During the sintering process under high pressure and temperature, the grains start to grow due to the mobility of atoms at high temperatures, leading to an increase in the size of the grain, which is called grain growth. However, the highest content of Ce (x = 0.1) has a large amount of CoSb2 secondary phase that hinders the grain growth of Co4Sb12 grain. Furthermore, the difference between the samples could also be due to the variation in the microstructure of the ribbons from the RSP. To confirm this, the ribbons of samples Ce = 0.025 and 0.1 were investigated using FE-SEM. In the RSP, the melted composite was pulled out, and contacted the rotating copper wheel to form ribbons. Figure 3a–f show the cross-section of the ribbons from the RSP, and Figure 3a,d show that the thickness of the ribbons is ~12 µm. As shown, the grain size in the surface contacting the copper wheel (Figure 3b,e) is smaller than the one on the opposite surface (Figure 3c,f) (free surface). Based on the detailed contact and free surfaces of each sample, the particle size of the contacted face is around 50 nm, which is much smaller than the size of the free surface at 200 nm. The difference in the size of the particles between the two surfaces can be explained by the different cooling rates. Specifically, on the contacted surfaces, the particles are quickly cooled, and small particles are formed. The free surface has more time to cool, and the particles belonging to these areas are bigger because of the longer growth time. The distribution of each element of sample Ce0.025Yb0.175Co4Sb12 was determined by EDS mapping, as shown in Figure 4. The Sb and Co elements are mainly distributed throughout the surface, not only in this sample, but also in others. The Yb and Ce, which have low amounts of filling, are discretely distributed in all samples. The details for actual compositions and oxygen contents are displayed in Tables S1 and S2.
The carrier concentration and mobility are shown in Figure 5a as a function of Ce content at room temperature. The strong tendencies of the carrier concentration and mobility have been shown in the CexYb0.2−xCo4Sb12 (x = 0, 0.025, 0.05, 0.075, and 0.1) system. It is shown that the carrier concentration increases (from 1.42 × 1020 cm−3 for Ce = 0.000 sample to ~3.41 × 1020 cm−3 for Ce = 0.100 sample) as Ce substitution increases. This suggests that Ce atoms also act as donor dopants. Although Ce and Yb are from the same lanthanide family, they have different valence states of ions in the skutterudite matrix. Thus, Ce fillers will donate more electrons to the matrix than Yb fillers, leading to a gradual increase in carrier concentration. Furthermore, the carrier mobility shows the opposite tendency to the carrier concentration. Adding more Ce content increases the carrier concentration, which gives electrons a greater opportunity to interact with ions. This decreases the mean free path, and thus decreases mobility. The electrical conductivity is the combination of carrier concentration and mobility, as shown in Figure 5b. Overall, the electrical conductivity shows a trend of increasing with increasing Ce fillers, and the electrical conductivity decreases at higher temperatures, indicating the characteristics of a heavily doped semiconductor. The unfilled Ce shows a decrease in its electrical conductivity value, from 2085 S/cm at room temperature to 1185 S/cm at 860 K. These trends are also observed in all other Ce-filled samples.
The temperature-dependent Seebeck coefficient and the power factor for CexYb0.2−xCo4Sb12 (x = 0, 0.025, 0.05, 0.075, and 0.1) are shown in Figure 6. The Seebeck coefficients for all samples in Figure 6a display negative values, indicating the n-type behavior of the semiconductor. The absolute value of the Seebeck coefficient clearly decreases from sample Ce0.025Yb0.175Co4Sb12 (−141.8 µV/K at 300 K) to sample Ce0.1Yb0.1Co4Sb12 (−123.2 µV/K at 300 K) as the carrier concentration increases. The power factor shows a trend of decreasing with more Ce content due to the changing electrical conductivity and Seebeck coefficient, as shown in Figure 6b. The highest power factor for Ce-Yb double filler is 50.6 × 10−4 W/mK2 for sample x = 0.050 at around 715 K. However, at high temperatures above 800 K, the power factor tends to decrease because of the reduction in electrical conductivity. This can be explained by the fact that at high temperatures, the intrinsic carrier is dominantly excited. The high-power factor (around 50 × 10−4 W/mK2) is obtained for samples Ce = 0 and 0.025.
The Seebeck coefficient can also be determined as a function of carrier concentration and effective mass using the Pisarenko relation for degenerate semiconductors, which can be expressed as the equation below. For this analysis, we assume a single parabolic band and an energy-independent carrier scattering approximation for degenerate semiconductors [19,42]:
S = 8 π 2 k B 2 T 3 q h 2 m * π 3 n 2 / 3
where kB is the Boltzmann constant, T is the absolute temperature, h is Planck’s constant, q is the unit charge of the electron, and m* is the effective mass. The effective mass m*/m0 is calculated and listed in Table 1, and Figure 7 shows the relation between the effective mass and carrier concentration at room temperature. In detail, the effective mass has large values that are 1.77 times the free electron mass value m0 for the sample without Ce concentration, and it then increases by 2.77 times, with the index 0.025 of Ce, and reaches 2.95 for the x = 0.100 sample. This proves that the participation of Ce contributes to the large mass effect in the system. The enhancement of effective mass in this system is associated with the valence instability of the 4f electrons in Ce and Yb-based compounds, and the heaviness of charge carriers has been reported in several cases [43,44,45,46].
The temperature-dependent thermal properties of a series of Ce and Yb double-filled samples are displayed in Figure 8a. The total thermal conductivity decreases with an increase in temperature, indicating that phonon–phonon scattering is dominant. In contrast, as the measured temperature increases further, the total thermal conductivity tends to increase due to the contribution of electron–phonon interaction. To explore the effect of Ce-Yb on the lattice’s thermal conductivity, the electronic thermal conductivity was determined using the Wiedemann–Franz relation (ke = LσT), with the Lorenz number, L, calculated using the Seebeck coefficient, following the formula L = 1.5 + Exp[−|S|/116] (Figure S6, Supporting Information) [47]. The lattice thermal conductivity can be determined by klat = ktotalke, and the result is shown in Figure 8b. The participation of Ce leads to a decrease in lattice thermal conductivity. However, the lattice thermal conductivity of sample 0.100 jumps up to a higher value, which is related to the high intensity of the secondary phase in the XRD result. That is, one would expect the reduction of the lattice thermal conductivities in Ce-Yb double-filled samples compared to the Yb single-filled sample. However, as we have clearly seen in the XRD patterns of our samples, the secondary phase of CoSb2, with high lattice thermal conductivity [48], exists in our double-filled samples, resulting in higher values of lattice thermal conductivity compared to the Yb single-filled sample.
The dimensionless figure of merit ZT was calculated for all CexYb0.2−xCo4Sb12 samples, and the results are shown in Figure 9. The lowest ZT was observed for sample x = 0.075, which is related to the lowest power factor and highest total thermal conductivity. The maximum ZT value for the double-filled samples in this study was achieved for sample x = 0.05 at 750 K with a value of 1.15, and this is due to the lowest lattice conductivity and highest power factor. This value is higher than that of double-filled Yb-In (ZT = 0.97) [20] and Yb-Ca (ZT = 1.03) [16] at the same temperature of 750 K, but lower than that of Yb-Ba (ZT = 1.36) [21]. This observation can be explained by the fact that the combination of heavier fillers more effectively scatters lattice phonons compared to the combination of lighter fillers in the SKD materials system, resulting in a greater reduction of klat and a higher ZT value.

4. Conclusions

A small amount of Ce and Yb was successfully introduced into the SKD structure through the process of melt spinning and SPS. XRD data revealed a gradual reduction in the lattice parameter of SKD with increasing Ce contents. The difference in grain size observed in the SEM images is attributed to the cooling rate between the two faces of the RSP ribbons. By replacing Yb atoms with Ce, the deficiency of the carrier concentration was compensated for by the extra electron from the Ce donors. This led to the optimization of the electrical conductivity, Seebeck coefficient, and power factor of the CexYb0.2−xCo4Sb12 system. At high temperatures, the power factor shows a downturn due to bipolar conduction in the intrinsic conduction regime. Moreover, the lattice thermal conductivity of the CexYb0.2−xCo4Sb12 skutterudite system in the range of 0.025 to 0.1 Ce content was clearly suppressed, as the Ce content increased due to the introduction of the dual phonon scattering center from the Ce and Yb fillers. The highest ZT value for the double-filled skutterudite was achieved with a value of 1.15 at 750 K for the Ce0.05Yb0.15Co4Sb12 sample. However, this value is lower than that of the Yb single-filled SKD, which could be due to the deterioration effect from the secondary phase of CoSb2 in our double-filled SKD system. One would expect an enhancement of thermoelectric performance of the double-filled SKD system, under the combination of the rapid synthesis of melt spinning with the spark plasma sintering process, if the secondary phase formation is well controlled.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ma16103819/s1.

Author Contributions

Conceptualization, J.Y.C. and W.H.S.; methodology, N.V.B., N.V.D.; validation, N.L., M.K. and S.H.R.; investigation, M.L., D.S. and W.H.N.; data curation, N.V.B., N.V.D. and J.Y.C.; writing—original draft preparation, N.V.B., J.W.R. and S.L.; writing—review and editing, J.Y.C., W.H.S., S.Y.K. and S.-M.K.; visualization, S.Y.K. and S.-M.K.; funding acquisition, W.H.N., J.W.R. and W.H.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Korea Institute of Ceramic Engineering and Technology KPP22002, the National Research Foundation of Korea (NRF-2021R1A5A8033165), and the Technology Innovation Program RS-2022-00144027 funded by the Ministry of Trade, Industry & Energy (MOTIE, Korea). The present research has been also conducted by the Research Grant of Kwangwoon University in 2022.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available upon request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Riffat, S.B.; Ma, X. Thermoelectrics: A review of present and potential applications. Appl. Ther. Eng. 2003, 23, 913–935. [Google Scholar] [CrossRef]
  2. Tritt, T.M.; Subramanian, M.A. Thermoelectric materials, phenomena, and applications: A bird’s eye view. MRS Bull. 2006, 31, 188–198. [Google Scholar] [CrossRef]
  3. Snyder, G.J.; Toberer, E.S. Complex thermoelectric materials. In Materials for Sustainable Energy: A Collection of Peer-Reviewed Research and Review Articles from Nature Publishing Group; World Scientific: Singapore, 2011; pp. 101–110. [Google Scholar]
  4. Sales, B.; Mandrus, D.; Williams, R.K. Filled skutterudite antimonides: A new class of thermoelectric materials. Science 1996, 272, 1325–1328. [Google Scholar] [CrossRef] [PubMed]
  5. Dughaish, Z. Lead telluride as a thermoelectric material for thermoelectric power generation. Phys. B Condens. Matter 2002, 322, 205–223. [Google Scholar] [CrossRef]
  6. Rull-Bravo, M.; Moure, A.; Fernandez, J.; Martín-González, M. Skutterudites as thermoelectric materials: Revisited. RSC Adv. 2015, 5, 41653–41667. [Google Scholar] [CrossRef]
  7. 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]
  8. Dahal, T.; Kim, H.S.; Gahlawat, S.; Dahal, K.; Jie, Q.; Liu, W.; Lan, Y.; White, K.; Ren, Z. Transport and mechanical properties of the double-filled p-type skutterudites La0.68Ce0.22Fe4−xCoxSb12. Acta Mater. 2016, 117, 13–22. [Google Scholar] [CrossRef]
  9. Music, D.; Geyer, R.W.; Keuter, P. Thermomechanical response of thermoelectrics. Appl. Phys. Lett. 2016, 109, 223903. [Google Scholar] [CrossRef]
  10. Nolas, G.; Slack, G.; Morelli, D.; Tritt, T.; Ehrlich, A. The effect of rare-earth filling on the lattice thermal conductivity of skutterudites. J. Appl. Phys. 1996, 79, 4002–4008. [Google Scholar] [CrossRef]
  11. Keppens, V.; Mandrus, D.; Sales, B.C.; Chakoumakos, B.; Dai, P.; Coldea, R.; Maple, M.; Gajewski, D.; Freeman, E.; Bennington, S. Localized vibrational modes in metallic solids. Nature 1998, 395, 876–878. [Google Scholar] [CrossRef]
  12. Hermann, R.P.; Grandjean, F.; Long, G.J. Einstein oscillators that impede thermal transport. Am. J. Phys. 2005, 73, 110–118. [Google Scholar] [CrossRef]
  13. Rowe, D.M. CRC Handbook of Thermoelectrics; CRC Press: Boca Raton, FL, USA, 2018. [Google Scholar]
  14. Sales, B.; Mandrus, D.; Chakoumakos, B.; Keppens, V.; Thompson, J. Filled skutterudite antimonides: Electron crystals and phonon glasses. Phys. Rev. B 1997, 56, 15081. [Google Scholar] [CrossRef]
  15. Tang, Y.; Gibbs, Z.M.; Agapito, L.A.; Li, G.; Kim, H.-S.; Nardelli, M.B.; Curtarolo, S.; Snyder, G.J. Convergence of multi-valley bands as the electronic origin of high thermoelectric performance in CoSb3 skutterudites. Nat. Mater. 2015, 14, 1223–1228. [Google Scholar] [CrossRef] [PubMed]
  16. Salvador, J.; Yang, J.; Wang, H.; Shi, X. Double-filled skutterudites of the type YbxCayCo4Sb12: Synthesis and properties. J. Appl. Phys. 2010, 107, 043705. [Google Scholar] [CrossRef]
  17. Yang, J.; Zhang, W.; Bai, S.; Mei, Z.; Chen, L. Dual-frequency resonant phonon scattering in BaxRyCo4Sb12 (R = La, Ce, and Sr). Appl. Phys. Lett. 2007, 90, 192111. [Google Scholar] [CrossRef]
  18. Wang, S.; Salvador, J.R.; Yang, J.; Wei, P.; Duan, B.; Yang, J. High-performance n-type YbxCo4Sb12: From partially filled skutterudites towards composite thermoelectrics. NPG Asia Mater. 2016, 8, e285. [Google Scholar] [CrossRef]
  19. Hu, P.; Wei, T.-R.; Qiu, P.; Cao, Y.; Yang, J.; Shi, X.; Chen, L. Largely enhanced Seebeck coefficient and thermoelectric performance by the distortion of electronic density of states in Ge2Sb2Te5. ACS Appl. Mater. Interfaces 2019, 11, 34046–34052. [Google Scholar] [CrossRef]
  20. Peng, J.; Alboni, P.; He, J.; Zhang, B.; Su, Z.; Holgate, T.; Gothard, N.; Tritt, T. Thermoelectric properties of (In, Yb) double-filled CoSb3 skutterudite. J. Appl. Phys. 2008, 104, 053710. [Google Scholar] [CrossRef]
  21. Shi, X.; Kong, H.; Li, C.-P.; Uher, C.; Yang, J.; Salvador, J.; Wang, H.; Chen, L.; Zhang, W. Low thermal conductivity and high thermoelectric figure of merit in n-type BaxYbyCo4Sb12 double-filled skutterudites. Appl. Phys. Lett. 2008, 92, 182101. [Google Scholar] [CrossRef]
  22. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta crystallogr. Sect. A Cryst. Phys. Diffr. Theor. Gen. Crystallogr. 1976, 32, 751–767. [Google Scholar] [CrossRef]
  23. Li, H.; Tang, X.; Zhang, Q.; Uher, C. Rapid preparation method of bulk nanostructured Yb0.3Co4Sb12+y compounds and their improved thermoelectric performance. Appl. Phys. Lett. 2008, 93, 252109. [Google Scholar] [CrossRef]
  24. Jie, Q.; Wang, H.; Liu, W.; Wang, H.; Chen, G.; Ren, Z. Fast phase formation of double-filled p-type skutterudites by ball-milling and hot-pressing. Phys. Chem. Chem. Phys. 2013, 15, 6809–6816. [Google Scholar] [CrossRef] [PubMed]
  25. Li, H.; Tang, X.; Su, X.; Zhang, Q. Preparation and thermoelectric properties of high-performance Sb additional Yb0.2Co4Sb12+y bulk materials with nanostructure. Appl. Phys. Lett. 2008, 92, 202114. [Google Scholar] [CrossRef]
  26. Yu, J.; Zhao, W.; Zhou, H.; Wei, P.; Zhang, Q. Rapid preparation and thermoelectric properties of Ba and In double-filled p-type skutterudite bulk materials. Scr. Mater. 2013, 68, 643–646. [Google Scholar] [CrossRef]
  27. Tan, G.; Liu, W.; Wang, S.; Yan, Y.; Li, H.; Tang, X.; Uher, C. Rapid preparation of CeFe4Sb12 skutterudite by melt spinning: Rich nanostructures and high thermoelectric performance. J. Mater. Chem. A 2013, 1, 12657–12668. [Google Scholar] [CrossRef]
  28. Salvador, J.R.; Waldo, R.A.; Wong, C.A.; Tessema, M.; Brown, D.N.; Miller, D.J.; Wang, H.; Wereszczak, A.A.; Cai, W. Thermoelectric and mechanical properties of melt spun and spark plasma sintered n-type Yb-and Ba-filled skutterudites. Mater. Sci. Eng. B 2013, 178, 1087–1096. [Google Scholar] [CrossRef]
  29. Thompson, D.R.; Liu, C.; Yang, J.; Salvador, J.R.; Haddad, D.B.; Ellison, N.D.; Waldo, R.A.; Yang, J. Rare-earth free p-type filled skutterudites: Mechanisms for low thermal conductivity and effects of Fe/Co ratio on the band structure and charge transport. Acta Mater. 2015, 92, 152–162. [Google Scholar] [CrossRef]
  30. He, J.; Sootsman, J.R.; Girard, S.N.; Zheng, J.-C.; Wen, J.; Zhu, Y.; Kanatzidis, M.G.; Dravid, V.P. On the origin of increased phonon scattering in nanostructured PbTe based thermoelectric materials. J. Am. Chem. Soc. 2010, 132, 8669–8675. [Google Scholar] [CrossRef]
  31. Rogl, G.; Grytsiv, A.; Rogl, P.; Bauer, E.; Hochenhofer, M.; Anbalagan, R.; Mallik, R.; Schafler, E. Nanostructuring of p- and n-type skutterudites reaching figures of merit of approximately 1.3 and 1.6, respectively. Acta Mater. 2014, 76, 434–448. [Google Scholar] [CrossRef]
  32. Nandihalli, N.; Gregory, D.H.; Mori, T. Energy-Saving Pathways for Thermoelectric Nanomaterial Synthesis: Hydrothermal/Solvothermal, Microwave-Assisted, Solution-Based, and Powder Processing. Adv. Sci. 2022, 9, 2106052. [Google Scholar] [CrossRef]
  33. Emsley, J. Nature’s Building Blocks: An AZ Guide to the Elements; Oxford University Press: Oxford, UK, 2011. [Google Scholar]
  34. Tang, Y.; Hanus, R.; Chen, S.-W.; Snyder, G.J. Solubility design leading to high figure of merit in low-cost Ce-CoSb3 skutterudites. Nat. Commun. 2015, 6, 7584. [Google Scholar] [CrossRef] [PubMed]
  35. Khovaylo, V.; Korolkov, T.; Voronin, A.; Gorshenkov, M.; Burkov, A. Rapid preparation of InxCo4Sb12 with a record-breaking ZT = 1.5: The role of the In overfilling fraction limit and Sb overstoichiometry. J. Mater. Chem. A 2017, 5, 3541–3546. [Google Scholar] [CrossRef]
  36. Li, H.; Tang, X.; Su, X.; Zhang, Q.; Uher, C. Nanostructured bulk YbxCo4Sb12 with high thermoelectric performance prepared by the rapid solidification method. J. Phys. D Appl. Phys. 2009, 42, 145409. [Google Scholar] [CrossRef]
  37. Aversano, F.; Branz, S.; Bassani, E.; Fanciulli, C.; Ferrario, A.; Boldrini, S.; Baricco, M.; Castellero, A. Effect of rapid solidification on the synthesis and thermoelectric properties of Yb-filled Co4Sb12 skutterudite. J. Alloys Compd. 2019, 796, 33–41. [Google Scholar] [CrossRef]
  38. Kruszewski, M.; Zybała, R. Review of rapid fabrication methods of skutterudite materials. Mater. Today Proc. 2021, 44, 3475–3482. [Google Scholar] [CrossRef]
  39. Yang, K.; Cheng, H.; Hng, H.; Ma, J.; Mi, J.; Zhao, X.; Zhu, T.; Zhang, Y. Synthesis and thermoelectric properties of double-filled skutterudites CeyYb0.5−yFe1.5Co2.5Sb12. J. Alloys Compd. 2009, 467, 528–532. [Google Scholar] [CrossRef]
  40. Bérardan, D.; Godart, C.; Alleno, E.; Bauer, E. Chemical properties and thermopower of the new series of skutterudite Ce1−pYbpFe4Sb12. J. Alloys Compd. 2003, 351, 18–23. [Google Scholar] [CrossRef]
  41. Berardan, D.; Alleno, E.; Godart, C.; Puyet, M.; Lenoir, B.; Lackner, R.; Bauer, E.; Girard, L.; Ravot, D. Improved thermoelectric properties in double-filled Cey/2Yby/2Fe4−x(Co/Ni)xSb12 skutterudites. J. Appl. Phys. 2005, 98, 033710. [Google Scholar] [CrossRef]
  42. 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, 554–557. [Google Scholar] [CrossRef]
  43. Fisk, Z.; Sarrao, J.; Smith, J.; Thompson, J. The physics and chemistry of heavy fermions. Proc. Natl. Acad. Sci. USA 1995, 92, 6663–6667. [Google Scholar] [CrossRef]
  44. Cho, J.Y.; Thomas, E.L.; Nambu, Y.; Capan, C.; Karki, A.B.; Young, D.P.; Kuga, K.; Nakatsuji, S.; Chan, J.Y. Crystal Growth, Structure, and Physical Properties of Ln(Cu,Ga)13−x(Ln = La−Nd, Eu; x ≈ 0.2). Chem. Mater. 2009, 21, 3072–3078. [Google Scholar] [CrossRef]
  45. Steglich, F. Superconductivity and magnetism in heavy-fermion compounds. J. Phys. Soc. Jpn. 2005, 74, 167–177. [Google Scholar] [CrossRef]
  46. Miyake, K.; Narikiyo, O.; Onishi, Y. Superconductivity of Ce-based heavy fermions under pressure: Valence fluctuation mediated pairing associated with valence instability of Ce. Phys. B Condens. Matter 1999, 259, 676–677. [Google Scholar] [CrossRef]
  47. Kim, H.-S.; Gibbs, Z.M.; Tang, Y.; Wang, H.; Snyder, G.J. Characterization of Lorenz number with Seebeck coefficient measurement. APL Mater. 2015, 3, 041506. [Google Scholar] [CrossRef]
  48. Goto, Y.; Miyao, S.; Kamihara, Y.; Matoba, M. Electrical/thermal transport and electronic structure of the binary cobalt pnictides CoPn2 (Pn = As and Sb). AIP Adv. 2015, 5, 067147. [Google Scholar] [CrossRef]
Figure 1. (a) Powder XRD of CexYb0.2−xCo4Sb12 after SPS, (b) the secondary phase peak of CoSb2, (c) zoomed in XRD pattern between 31° and 31.8°, and (d) calculated lattice parameter of CexYb0.2−xCo4Sb12 samples as a function of Ce content.
Figure 1. (a) Powder XRD of CexYb0.2−xCo4Sb12 after SPS, (b) the secondary phase peak of CoSb2, (c) zoomed in XRD pattern between 31° and 31.8°, and (d) calculated lattice parameter of CexYb0.2−xCo4Sb12 samples as a function of Ce content.
Materials 16 03819 g001
Figure 2. FE-SEM images of CexYb0.2−xCo4Sb12 samples after SPS for (a) x = 0.000, (b) x = 0.025, (c) x = 0.050, (d) x = 0.075, and (e) x = 0.100.
Figure 2. FE-SEM images of CexYb0.2−xCo4Sb12 samples after SPS for (a) x = 0.000, (b) x = 0.025, (c) x = 0.050, (d) x = 0.075, and (e) x = 0.100.
Materials 16 03819 g002
Figure 3. FE-SEM images of ribbons from the RSP process. (a) The cross section, (b) contact surface, and (c) free surface of sample Ce0.025Yb0.175Co4Sb12; (d) the cross section, (e) contact, and (f) free surface of sample Ce0.1Yb0.1Co4Sb12.
Figure 3. FE-SEM images of ribbons from the RSP process. (a) The cross section, (b) contact surface, and (c) free surface of sample Ce0.025Yb0.175Co4Sb12; (d) the cross section, (e) contact, and (f) free surface of sample Ce0.1Yb0.1Co4Sb12.
Materials 16 03819 g003
Figure 4. EDS images of sample Ce0.025Yb0.175Co4Sb12. (a) The SEM image, the distribution of (b) Co, (c) Sb, (d) Ce, (e) Yb and (f) O (EDS images of other compositions are displayed in Figures S1–S5).
Figure 4. EDS images of sample Ce0.025Yb0.175Co4Sb12. (a) The SEM image, the distribution of (b) Co, (c) Sb, (d) Ce, (e) Yb and (f) O (EDS images of other compositions are displayed in Figures S1–S5).
Materials 16 03819 g004
Figure 5. (a) The carrier concentration and mobility of CexYb0.2−xCo4Sb12 (x = 0, 0.025, 0.05, 0.075, and 0.1) at room temperature, and (b) the temperature dependence of electrical conductivity.
Figure 5. (a) The carrier concentration and mobility of CexYb0.2−xCo4Sb12 (x = 0, 0.025, 0.05, 0.075, and 0.1) at room temperature, and (b) the temperature dependence of electrical conductivity.
Materials 16 03819 g005
Figure 6. The temperature dependence of (a) the Seebeck coefficient and (b) the power factor of CexYb0.2−xCo4Sb12 (x = 0, 0.025, 0.05, 0.075, and 0.1).
Figure 6. The temperature dependence of (a) the Seebeck coefficient and (b) the power factor of CexYb0.2−xCo4Sb12 (x = 0, 0.025, 0.05, 0.075, and 0.1).
Materials 16 03819 g006
Figure 7. The effective mass m*/m0 at room temperature as a function of carrier concentration for CexYb0.2−xCo4Sb12.
Figure 7. The effective mass m*/m0 at room temperature as a function of carrier concentration for CexYb0.2−xCo4Sb12.
Materials 16 03819 g007
Figure 8. The temperature dependence of (a) the total thermal conductivity and (b) the lattice thermal conductivity of sample CexYb0.2−xCo4Sb12 (x = 0, 0.025, 0.05, 0.075, and 0.1).
Figure 8. The temperature dependence of (a) the total thermal conductivity and (b) the lattice thermal conductivity of sample CexYb0.2−xCo4Sb12 (x = 0, 0.025, 0.05, 0.075, and 0.1).
Materials 16 03819 g008
Figure 9. The temperature dependence of the ZT of sample CexYb0.2−xCo4Sb12 (x = 0, 0.025, 0.05, 0.075, and 0.1).
Figure 9. The temperature dependence of the ZT of sample CexYb0.2−xCo4Sb12 (x = 0, 0.025, 0.05, 0.075, and 0.1).
Materials 16 03819 g009
Table 1. Nominal composition and room temperature values of the lattice parameter (a), carrier concentration (n), electrical conductivity (σ), Seebeck coefficient (S), thermal conductivity (k) and effective mass (m*/m0) for CexYb0.2−xCo4Sb12.
Table 1. Nominal composition and room temperature values of the lattice parameter (a), carrier concentration (n), electrical conductivity (σ), Seebeck coefficient (S), thermal conductivity (k) and effective mass (m*/m0) for CexYb0.2−xCo4Sb12.
Compositiona (Å)n (×1020 cm−3)σ (Sm−1)S (µVK−1)k (Wm−1K−1)m*/m0
Yb0.2Co4Sb129.0521.42085.5−133.12.41.77
Ce0.025Yb0.175Co4Sb129.0472.51756.1−143.23.02.77
Ce0.05Yb0.15Co4Sb129.0472.71783.9−141.82.52.92
Ce0.075Yb0.125Co4Sb129.0443.22016.9−128.92.62.94
Ce0.1Yb0.1Co4Sb129.0413.41932.2−123.24.32.95
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Binh, N.V.; Van Du, N.; Lee, N.; Kang, M.; Ryu, S.H.; Lee, M.; Seo, D.; Nam, W.H.; Roh, J.W.; Lee, S.; et al. Investigation of the Effect of Double-Filler Atoms on the Thermoelectric Properties of Ce-YbCo4Sb12. Materials 2023, 16, 3819. https://doi.org/10.3390/ma16103819

AMA Style

Binh NV, Van Du N, Lee N, Kang M, Ryu SH, Lee M, Seo D, Nam WH, Roh JW, Lee S, et al. Investigation of the Effect of Double-Filler Atoms on the Thermoelectric Properties of Ce-YbCo4Sb12. Materials. 2023; 16(10):3819. https://doi.org/10.3390/ma16103819

Chicago/Turabian Style

Binh, Nguyen Vu, Nguyen Van Du, Nayoung Lee, Minji Kang, So Hyeon Ryu, Munhwi Lee, Deokcheol Seo, Woo Hyun Nam, Jong Wook Roh, Soonil Lee, and et al. 2023. "Investigation of the Effect of Double-Filler Atoms on the Thermoelectric Properties of Ce-YbCo4Sb12" Materials 16, no. 10: 3819. https://doi.org/10.3390/ma16103819

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