Next Article in Journal
Fault Current Characteristics of the DFIG under Asymmetrical Fault Conditions
Next Article in Special Issue
Thermoelectric Properties of Alumina-Doped Bi0.4Sb1.6Te3 Nanocomposites Prepared through Mechanical Alloying and Vacuum Hot Pressing
Previous Article in Journal
Wind Tunnel Studies of a Pedestrian-Level Wind Environment in a Street Canyon between a High-Rise Building with a Podium and Low-Level Attached Houses
Previous Article in Special Issue
Investigation of the Promotion of Wind Power Consumption Using the Thermal-Electric Decoupling Techniques
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhancing Thermoelectric Properties of Si80Ge20 Alloys Utilizing the Decomposition of NaBH4 in the Spark Plasma Sintering Process

1
Department of Physics & Astronomy, Clemson University, Clemson, SC 29634, USA
2
Clemson Nanomaterials Center, and Center for Optical Materials Science and Engineering Technologies, Clemson University, Clemson, SC 29625, USA
3
Department of Materials Science & Engineering, Clemson University, Clemson, SC 29634, USA
*
Authors to whom correspondence should be addressed.
Energies 2015, 8(10), 10958-10970; https://doi.org/10.3390/en81010958
Submission received: 17 July 2015 / Revised: 8 September 2015 / Accepted: 14 September 2015 / Published: 29 September 2015
(This article belongs to the Special Issue Thermoelectric Energy Harvesting)

Abstract

:
The thermoelectric properties of spark plasma sintered, ball-milled, p-type Si80Ge20-(NaBH4)x (x = 0.7,1.7 and 2.7), and Si80Ge20B1.7-y-(NaBH4)y (y = 0.2 and 0.7) samples have been investigated from 30 K to 1100 K. These samples were prepared by spark plasma sintering of an admixture of Si, Ge, B and NaBH4 powders. In particular, the degasing process during the spark plasma sintering process, the combined results of X-ray powder diffraction, Raman spectroscopy, Hall coefficient, electrical resistivity, and Seebeck coefficient measurements indicated that NaBH4 decomposed into Na, B, Na2B29, and H2 during the spark plasma sintering process; Na and B were doped into the SiGe lattice, resulting in favorable changes in the carrier concentration and the power factor. In addition, the ball milling process and the formation of Na2B29 nanoparticles resulted in stronger grain boundary scattering of heat-carrying phonons, leading to a reduced lattice thermal conductivity. As a result, a significant improvement in the figure of merit ZT (60%) was attained in p-type Si80Ge20-(NaBH4)1.7 and Si80Ge20-B1.5(NaBH4)0.7 at 1100 K as compared to the p-type B-doped Si80Ge20 material used in the NASA’s radioactive thermoelectric generators. This single-step “doping-nanostructuring” procedure can possibly be applied to other thermoelectric materials.

1. Introduction

Thermoelectric materials are of technological interest owing to their ability of direct heat-to-electricity energy conversion. Currently, Silicon-Germanium (SiGe) alloys are the only thermoelectric materials that have found applications in power generation in the temperature range of 900 K < T < 1300 K [1,2]. Many efforts have been exerted to enhance the dimensionless thermoelectric figure of merit Z T = ( α 2 σ / κ ) T of SiGe compounds [3,4], where α is the Seebeck coefficient, σ the electrical conductivity, κ the thermal conductivity, T the absolute temperature, and P F = α 2 σ T the power factor. The p-type (Boron, B-doped) and n-type (Phosphorus, P-doped) SiGe material used in NASA’s radioisotope thermoelectric generators (RTGs) possess a ZT of 0.5 and 0.9, respectively [2]. In the past few decades, there have been many theoretical [5,6,7] and experimental efforts [8,9,10,11] toward further enhancing the ZT of SiGe via an enhancement of the PF and/or a reduction in κ . Some results are noteworthy. For example, particle size distribution was found to be crucial to achieve dense homogeneous samples [12]; polycrystalline SiGe possesses a better thermoelectric performance than single crystalline SiGe [13]; mechanical ball milling is effective in producing single phased SiGe powder [14,15]. A significant improvement in ZT with peak values about 1.3 and 0.95 at 1200 K in n-type and p-type respectively were achieved via reducing the lattice thermal conductivity ( κ L ) using a high energy ball milling method to produce nanostructured SiGe [11,16]. Recently we have employed the spark plasma sintering (SPS) technique to synthesize high relative density SiGe directly from single elemental (SE) Si and Ge powders, which possess thermoelectric performance comparable to those used in NASA’s RTGs [17].
Doping plays a central role in optimizing the thermoelectric performance of SiGe as dopants optimize the carrier concentration while simultaneously acting as point defects which can strongly scatter the heat-carrying phonons at elevated temperatures. Previously the effects of different dopants, including P, Ga, B, GaP, In, Sb, and InSb, on the thermoelectric properties of Si80Ge20 alloys have been investigated [18,19]. Doping typically governs the PF, which is at the core and often the bottle-neck of enhancing ZT in many thermoelectric materials. It is generally accepted that a strongly energy-dependent differential electrical conductivity leads to enhancement of the PF [20]. Such strong energy dependence arises from the density of states and/or from the relaxation time of charge carriers [21,22]. In addition to doping, the “compositing or nanostructuring” process is another way to enhance the PF. Bergman and Fel showed that the PF could be enhanced by making a two-phase composite in a parallel slab micromorphology or a core-shell micromorphology [23]. Doping and nanostructuring combined in this modulation-doping approach to enhance the PF. For example, the PF of p-type Si80Ge20B1.5 was improved by a factor of 40% by embedding 30 vol. % B-doped SiGe nanoparticles in the intrinsic SiGe host matrix to form (Si80Ge20)0.7(Si100B5)0.3 composites [24]. Furthermore, grain boundary engineering is another effective approach to enhancing the PF in several cases. For example, the alkali metal salt NaBH4 processing appears to lead to thermoelectrically-favorable grain boundaries in p-type Bi2Te3 [25] and Pbo.75Sn0.25Te [26].
In this work, we have simultaneously achieved Na-doping, B-doping, formation of Na2B29 nanoparticles, and sample densification via the decomposition of NaBH4 in a single-step synergistic “doping-nanostructuring-sintering” process. As a result, the PF is enhanced while the total thermal conductivity, κ , is reduced, thereby leading to a significantly enhanced ZT.

2. Experimental Results and Data Analyses

In this work, doping is achieved via decomposition of NaBH4 in the SPS process. Three samples were treated with sodium boron hydride (NaBH4)x (x = 0.7, 1.0, and 2.7), while two samples were treated with single element boron (B) in addition to sodium boron hydride B1.7-y(NaBH4)y (y = 0.2 and 0.7). The purpose of preparing the Si80Ge20B1,5(NaBH4)0.2 and Si80Ge20B1(NaBH4)0.7 sample with increasing percentage of NaBH4 and decreasing B, is to investigate the extent of the NaBH4 decomposition with the same total amount of B.
In general, the thermal decomposition of NaBH4 follows:
N a B H 4 i s o l i d i ( N a , B , H ) + j g a s j ( B , H )
where the products can be single elemental Na, B, H2, or binary phases, e.g., Na–B such as Na2B29, Na2B30, Na3B20 or Na–H, or ternary phases, Na–B–H [27].
The thermal decomposition of NaBH4 occurs between 600 °C and 700 °C [27], significantly below our SPS temperature of 1020 °C. Indeed, we observed the presence of B-rich Na–B compounds namely Na2B29 by SEM, EDX, and XRD measurements. Importantly, degasing was observed in spark plasma sintering NaBH4-added SiGe samples, but not in NaBH4-free samples, confirming the formation of H2. An immediate question arises as to where exactly the Na resides after this decomposition. Figure 1a shows the XRD pattern of NaBH4 –added Si80Ge20 alloys after the SPS process. The lattice constant was determined using the (111) and (220) peaks by Bragg’s law. The peaks from Na2B29 were observed and marked with asterisk in Figure 1a. The ICSD PDF file number of Na2B29 is 01-071-2824, and the space group is I 1 m 1 [27]. No other secondary phases were observed.
Interestingly, the XRD peaks of the sample treated with B and NaBH4 {Si80Ge20B1.5(NaBH4)0.2} shifted to higher angle, contrary to those samples treated with only NaBH4 {Si80Ge20(NaBH4)1.7}, indicating that the dominant dopant is B. The peak of the sample treated with only NaBH4 {Si80Ge20(NaBH4)1.7}, on the other hand, is shifted to lower angle, consistent with a dominant Na doping (Figure 1b). This opposite shifting is attributed to the ionic radii of the Na (99 pm) and B (35 pm) comparing to Si or Ge (39 pm) [28]. Shifts of Raman peaks track with those of XRD peaks, confirming the doping by Na and B in the SiGe host matrix (Figure 1c). The decomposition of NaBH4 in the SPS process opens a new way to dope Na into the SiGe host matrix, which is actually a (Si80Ge20) composition.
The TEM image (Figure 2a) shows that the ball milled admixture of Si and Ge powders has grain size from 0.3 to 1.0 μm, compared to 1–20 μm and 44 μm for pristine Si and Ge powders before the ball milling process. Despite the grain coarsening during the SPS process in each sample, we adopted the same ball milling and SPS conditions for all samples; it is, thus, plausible to assume the grain size of SiGe coarse grains should be nearly the same for all samples after SPS process, which is a fixed parameter in our study of the doping effects. Figure 2b shows the Na2B29 nanoparticles on the grain boundaries of SiGe host matrix grains.
Figure 1. (a) XRD patterns of the sample dominantly doped with B, the sample doped dominantly with Na, and pristine SiGe; (b) magnification of the (111) peak shift upon the Na and B doping, note that the peaks shift to the opposite direction; and (c) Raman peak shifting tracks with XRD peak shifting.
Figure 1. (a) XRD patterns of the sample dominantly doped with B, the sample doped dominantly with Na, and pristine SiGe; (b) magnification of the (111) peak shift upon the Na and B doping, note that the peaks shift to the opposite direction; and (c) Raman peak shifting tracks with XRD peak shifting.
Energies 08 10958 g001
Figure 2. (a) TEM image shows the grain size of ball milled SiGe powder before SPS; and (b) SEM image of a fracture surface shows the agglomerate of Na2B29 nanoparticles at the grain boundary.
Figure 2. (a) TEM image shows the grain size of ball milled SiGe powder before SPS; and (b) SEM image of a fracture surface shows the agglomerate of Na2B29 nanoparticles at the grain boundary.
Energies 08 10958 g002
Figure 3a–c present the temperature dependence of the electrical resistivity Seebeck coefficient, and power factor α 2 / ρ , respectively, of the dense bulk NaBH4-treated p-type Si80Ge20 in comparison to the p-leg material used in NASA’s RTGs [2]. All of the samples exhibit similar trends in the temperature dependence of their physical properties. The sign of the Seebeck coefficient and that of the Hall coefficient confirm a p-type conduction. As shown in Figure 3b, the Seebeck coefficient, of most of the as-prepared samples, rivals that of the reference (i.e., the p-leg material used in NASA RTGs), except the 2.7% NaBH4 sample that shows a lower α. Since the Seebeck coefficient is positively correlated to the effective mass m*, and inversely proportional to the carrier concentration n, one possible explanation for the decreased α in the sample with 2.7% NaBH4 is the increasing of n. A significant enhancement in the PF, compared to the reference, is attained for all the samples (see Figure 3c).
Figure 3. Temperature dependence of (a) the electrical conductivity; (b) Seebeck coefficient; and (c) power factor α 2 T / ρ , of five SPS prepared nanostructured dense bulk Si80Ge20 alloy samples doped with (NaBH4)x (x = 0.7,1.7 and 2.7) and B1.7-y(NaBH4)y (y = 0.2 and 0.7) in comparison to the p-leg material used in NASA’s RTGs.
Figure 3. Temperature dependence of (a) the electrical conductivity; (b) Seebeck coefficient; and (c) power factor α 2 T / ρ , of five SPS prepared nanostructured dense bulk Si80Ge20 alloy samples doped with (NaBH4)x (x = 0.7,1.7 and 2.7) and B1.7-y(NaBH4)y (y = 0.2 and 0.7) in comparison to the p-leg material used in NASA’s RTGs.
Energies 08 10958 g003
The carrier concentration of p-type Si80Ge20 is about 1.67 × 1020 cm−3 [11,13]. The decrease of the electrical resistivity ρ for the samples can be explained in terms of the increasing of both the carrier concentration n and the Hall mobility (n ~ 2.5 × 1020 cm−3 and μ ~ 34 cm2·V−1·s−1), especially for the sample treated with 2.7% NaBH4 that has higher n and μ (n ~ 4×1020 cm−3 and μ ~ 40 cm2·V−1·s−1 at room temperature), as shown in Figure 4a,b. The carrier concentration exhibits an essentially temperature-independent behavior, while the mobility shows a weak, but well-discerned, negative temperature coefficient. Among the conventional scattering mechanisms, such as the charge carrier-phonon scattering, charge carrier-charge neutral point defect scattering, charge carrier-charged point defect scattering, and charge carrier-grain boundary, only the charge carrier-phonon scattering mechanism gives rises to a negative temperature coefficient [29]. Hence, we conclude that the charge carrier-phonon scattering is the dominant carrier scattering mechanism, though the charged and charge neutral point defects may coexist. Note that the magnitude change of the carrier concentration agrees with a hole-doping scenario, as expected for the Na-doping and B-doping.
Figure 4. (a) Temperature dependent of carrier concentration, n; and (b) Hall mobility, μ.
Figure 4. (a) Temperature dependent of carrier concentration, n; and (b) Hall mobility, μ.
Energies 08 10958 g004
Figure 5a shows the temperature dependent total thermal conductivity of all the samples. With the exception of the Si80Ge20(NaBH4)2.7 sample the composited samples have lower total thermal conductivity than the reference. The sample treated with 2.7% NaBH4 shows a higher thermal conductivity than that of the reference, which is attributed to the large electronic thermal conductivity κ e (Figure 5b). The electronic thermal conductivity is estimated by the Wiedemann-Franz relation and subtracted from the total thermal conductivity to derive the lattice thermal conductivity. The lattice thermal conductivity κ L as a function of temperature is plotted in Figure 5c. Once again, the reference has the highest κ L , while the Si80Ge20(NaBH4)1.7 sample shows the lowest lattice thermal conductivity. The lattice thermal conductivity is reduced due to an increase in the number of grain boundaries [30] by ball milling and the formation of nanoparticles, and the extra point defects introduced by the Na and B dopants. The best result was attained by (NaBH4)1.7 treatment whose thermal conductivity is ~20%–25% lower than the reference over a wide temperature range between 300 K and 1100 K.
Figure 5. Temperature dependence of (a) the total thermal conductivity; (b) the electronic thermal conductivity; and (c) the lattice thermal conductivity.
Figure 5. Temperature dependence of (a) the total thermal conductivity; (b) the electronic thermal conductivity; and (c) the lattice thermal conductivity.
Energies 08 10958 g005
As mentioned earlier, the purpose of treating two samples with B and NaBH4 {Si80Ge20B1,5(NaBH4)0.2 and Si80Ge20B1(NaBH4)0.7} with increased percentage of NaBH4 and decreased B, is to investigate whether NaBH4 completely decomposes to only B, Na, and H2 or also to boron-rich sodium phase. We found that increasing the NaBH4 percentage leads to a larger reduction in the total and lattice thermal conductivities without a degradation of the electrical properties, which can be correlated with the increasing amount of Na2B29 nanoparticles at the grain boundaries. For example, the lowest thermal conductivity of Si80Ge20(NaBH4)1.7 sample that has most Na2B29 has the lowest lattice thermal conductivity. Figure 2b shows the Na2B29 nanoparticles imbedded in the fracture surface of the sample after SPS, where numerous nanoparticles with a size ~20 nm have emerged at the grain boundary. As the percentage of NaBH4 is increased (consequently increased Na2B29) the peak of the lattice thermal conductivity decreased especially for the Si80Ge20(NaBH4)1.7 sample that has a higher concentration of boron-rich sodium nanoparticles.
Dimensionless figure of merit ZT values higher than those of NASA’s RTG materials were attained (Figure 6) for all of the samples studied. This enhancement can be attributed to the enhancement in the PF and also the reduction of thermal conductivity, especially the lattice component. The ZT value of Si80Ge20-(NaBH4)1.7 and Si80Ge20-B1.5(NaBH4)0.2 shows a maximum of about 0.8 at 1100 K, which is about 45% higher than that of the p-leg material of NASA’s RTG (ZT ~ 0.5).
Figure 6. Dimensionless figure of merit ZT of all the samples compared to the p-leg material of NASA’s RTG.
Figure 6. Dimensionless figure of merit ZT of all the samples compared to the p-leg material of NASA’s RTG.
Energies 08 10958 g006

3. Experimental Details

S i 80 G e 20 alloys were fabricated by ball milling (BM) process followed by spark plasma sintering (SPS) procedure, which is both time- and cost-efficient, and easy to scale up. Small grain size powders were selected: 1–20 μm silicon powder (99.9985% Alfa Aesar®, 26 Parkridge Rd, Ward Hill, MA 01835, USA) and germanium powder −100 mesh (Alfa Aesar® 99.999%) and used in this process. To avoid oxidation, the powders were loaded into a milling container inside a glove box. Then the powders were ball milled for 12 h to further refine the grains and thoroughly mix the powders. The ball milled powders were then divided into five batches, mixed with appropriate amount of NaBH4 and B according to the nominal formulas Si80Ge20(NaBH4)x and Si80Ge20 B1.7-y(NaBH4)y, where x = 0.7, 1.7, 2.7, y = 0.2 and 0.7 in a 3-D mixer. These samples were sintered using a Dr. Sinter SPS-515S (Fuji Electronic Industrial Co.®) (Tokyo, Japan) and characterized regarding their thermoelectric properties [17]. The Archimedes method measurements showed that the densities of the as-pressed samples were least 98% of the theoretical density (2.99 ± 0.1 g/cc).
The phase purity and micromorphology of these samples before and after SPS were then characterized by X-ray diffraction (XRD) using a Rigaku® Miniflex, The Woodlands, TX, USA) and Hitachi® S-3400N (Hitachi America, Troy, NY USA) equipped with an Oxford X-act® energy dispersive X-ray spectroscopy (EDX). Oxford Instruments, Raleigh, NC, USA) In order to investigate the effect of the dopants, Raman spectra were acquired with Dilor® XY triple grating (Dilor®, Lille, France) and Renishaw® InVia Raman microscopes (Renishaw Inc.®, Hoffman Estates, IL, USA) with Elaser = 2.33 eV. The incident laser beam was focused using a 50× objective, and the laser power on the samples was kept to a minimum to avoid heating. The thermal diffusivity measurements were performed on the densified pellet with 12.7 mm diameter and 2 mm thickness before cutting to approximately 10 × 2 × 2 mm3 bars for other transport measurements. High-temperature thermal diffusivity measurements were made on a Netzsch® LFA 457 (Burlington, MA, USA) laser flash apparatus using the transient method from 300 K to 1100 K. The high-temperature thermal conductivity of the samples was calculated using κ = d D C V , where d is the thermal diffusivity, D the density and C V the specific heat at constant volume. The high-temperature specific heat at constant pressure, C P , was measured on a Netzsch® (Burlington, MA USA) differential scanning calorimeter 404 “Pegasus”. The specific heat at constant pressure,   C P , and the specific heat at constant volume, C V , were considered to be essentially the same for the calculation of the total thermal conductivity. High-temperature resistivity ( ρ = 1 / σ ) and Seebeck coefficient were measured using the commercially-available Ulvac® ZEM-3 (Boston, MA, USA) from 300 K to 1100 K. The low temperature thermal conductivity was measured, on the same samples, from ~ 10 K to room temperature using a custom-designed steady-state technique [31]. The lattice thermal conductivity ( κ L ) was obtained by applying the Wiedemann–Franz relationship, κ e   =   L 0 σ   T , with the Lorenz number for a degenerate semiconductor,   L 0 = 2 × 10 8 W Ω/ K 2 , σ the electrical conductivity, and the formula κ L   =   κ t o t a l     κ e . The low temperature Seebeck coefficient and electrical resistivity were measured on a custom-designed system from 30 K to 300 K [32]. Hall-effect measurements were performed on a Quantum Design® (San Diego, CA, USA) physical properties measurement system (PPMS) using a five-probe configuration by sweeping the magnetic field between ± 1 Tesla, and the carrier concentration was calculated from the Hall data R H = 1 / n e , where R H is the Hall coefficient, e the electron charge, and n the carrier concentration. The electron mobility was then calculated from μ = RH/ρ. Additionally a thermal stability test was carried out by annealing the samples at 1100 K for 48 h in evacuated tubes before reinvestigating their thermoelectric properties. Importantly, no significant degradation of the physical properties was found.

4. Conclusions

We successfully reduced the lattice thermal conductivity and enhanced the power factor of p-type SiGe alloys via decomposing NaBH4 and densifying samples in a single-step spark plasma sintering process. The decomposition of NaBH4 led to Na and B doping into the SiGe lattice and the formation of Na2B29 nanoparticles on the grain boundary of SiGe coarse grain. The reduction of the electrical resistivity is attributed to the increased carrier concentration. The reduction in the lattice thermal conductivity is mainly due to enhanced phonon scattering due to the increased point defect scattering, the increased number of grain boundaries, as well as the scattering from the boron-rich sodium nanoparticles. The present work presents a new method for doping Na into SiGe, which can, in principle, be extended to other alkali metal salts and in other existing thermoelectric materials.

Acknowledgments

Ali Lahwal acknowledges a special program of Ministry of Higher Education funded by the Libyan Government. We also acknowledge the early contributions of Daniel Thompson while at Clemson University and currently at General Motors R&D. Menghan Zhou, Dale Hitchcock, and Jian He would like to acknowledge the support of NSF DMR 1307740. Terry M. Tritt would like to acknowledge a support under a DoE subcontract with Nanosonics Inc. (Blacksburg, VA).

Author Contributions

Ali Lahwal measured the thermoelectric properties, analyzed the data of the thermoelectric properties and Hall measurements and drafted the manuscript.
Xiaoyu Zeng, and Menghan Zhou contributed in measuring the low and high temperature measurements of the thermal conductivity, Seebeck coefficient and electrical resistivity.
Sriparna Bhattacharya contributed in XRD and DES analysis.
Dale Hitchcock and Jian He contributed in Hall coefficient measurements and reviewing the results and the manuscript.
Mehmet Karakaya and Apparao M. Rao conducted the Raman measurements and data analysis.
Terry M. Tritt designed the instruments and reviewed the results and the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wood, C. Materials for Thermoelectric Energy Conversion. Rep. Prog. Phys. 1988, 51, 459–539. [Google Scholar] [CrossRef]
  2. Vining, C.B. Silicon Germanium. In CRC Handbook of Thermoelectrics; Rowe, D.M., Ed.; CRC Press: Boca Raton, FL, USA, 1995; pp. 329–338. [Google Scholar]
  3. Bhandari, C.M.; Rowe, D.M. Silicon–Germanium Alloys as High-Temperature Thermoelectric Materials. Contemp. Phys. 1980, 21, 219–242. [Google Scholar] [CrossRef]
  4. Vining, C.B. A Model for the High-Temperature Transport Properties of Heavily Doped n-Type Silicon-Germanium Alloys. J. Appl. Phys. 1991, 69, 331–341. [Google Scholar] [CrossRef]
  5. Hicks, L.D.; Dresselhaus, M.S. Thermoelectric Figure of Merit of a One-Dimensional Conductor. Phys. Rev. B 1993, 47, 16631–16634. [Google Scholar] [CrossRef]
  6. Chen, G. Phonon Transport in Low Dimensional Structures. In Recent Trends in Thermoelectric Materials Research III; Semiconductors and Semimetals; Tritt, T.M., Ed.; Academic: New York, NY, USA, 2001; Volume 71, pp. 203–259. [Google Scholar]
  7. Venkatasubramanian, R. Phonon Blocking Electron Transmitting Superlattice Structures as Thin Film Thermoelectric Materials. In Recent Trends in Thermoelectric Materials Research III; Semiconductors and Semimetals; Tritt, T.M., Ed.; Academic: New York, NY, USA, 2001; Volume 71, pp. 175–201. [Google Scholar]
  8. Venkatasubramanian, R.; Siivola, E.; Colpitts, T.; O’Quinn, B. Thin-Film Thermoelectric Devices with High Room-Temperature Figures of Merit. Nature 2001, 413, 597–602. [Google Scholar] [CrossRef] [PubMed]
  9. Poudeu, P.F.P.; D’Angelo, J.; Downey, A.D.; Short, J.L.; Hogan, T.P.; Kanatzidis, M.G. Nanostructures versus Solid Solutions: Low Lattice Thermal Conductivity and Enhanced Thermoelectric Figure of Merit in Pb9.6Sb0.2Te10–xSex Bulk Materials. J. Am. Chem. Soc. 2006, 128, 14347–14355. [Google Scholar] [CrossRef] [PubMed]
  10. Hochbaum, A.I.; Chen, R.; Delgado, R.D.; Liang, W.; Garnett, E.C.; Najarian, M.; Majumdar, A.; Yang, P. Enhanced thermoelectric performance of rough silicon nanowires. Nature 2008, 451, 163–167. [Google Scholar] [CrossRef] [PubMed]
  11. Wang, X.W.; Lee, H.; Lan, Y.C.; Zhu, G.H.; Joshi, G.; Wang, D.Z.; Yang, J.; Muto, A.J.; Tang, M.Y.; Klatsky, J.; et al. Enhanced thermoelectric figure of merit in nanostructured n-type silicon germanium bulk alloy. Appl. Phys. Lett. 2008, 93, 193121. [Google Scholar] [CrossRef]
  12. Baughman, R.J.; McVay, G.L.; Lefever, R.A. Preparation of hot-pressed silicon germanium ingots: Part I–III. Mater. Res. Bull. 1974, 9, 685–692, 735–744, 863–872. [Google Scholar] [CrossRef]
  13. Vining, C.B.; Laskow, W.; Hanson, J.; van der Beck, R.R.; Gorsuch, P. Thermoelectric properties of pressure-sintered Si0.8Ge0.2 thermoelectric alloys. J. Appl. Phys. 1991, 69, 4333–4340. [Google Scholar] [CrossRef]
  14. Cook, B.A.; Harringa, J.L.; Han, S.H.; Beaudry, B.J. Parasitic, effects of oxygen on the thermoelectric properties of Si80Ge20 doped with GaP and P. J. Appl. Phys. 1992, 72, 1423–1428. [Google Scholar] [CrossRef]
  15. Cook, B.A.; Harringa, J.L.; Han, S.H.; Vining, C.B. Si80Ge20 thermoelectric alloys prepared with GaP additions. J. Appl. Phys. 1995, 78, 5474–5480. [Google Scholar] [CrossRef]
  16. Joshi, G.; Lee, H.; Lan, Y.C.; Wang, X.W.; Zhu, G.H.; Wang, D.Z.; Gould, R.W.; Cuff, D.C.; Tang, M.Y.; Dresselhaus, M.S.; et al. Enhanced thermoelectric figure-of-merit in nanostructured p-type silicon germanium bulk alloys. Nano Lett. 2008, 8, 4670–4674. [Google Scholar] [CrossRef] [PubMed]
  17. Thompson, D.; Hitchcock, D.; Lahwal, A.; Tritt, T.M. Single-element spark plasma sintering of silicon germanium. Emerg. Mater. Res. 2012, 1, 299–305. [Google Scholar] [CrossRef]
  18. Ying, G.; Jiang, H.; Zhang, C.; Wu, X.; Niu, S. Thermoelectric properties on n-type Si80Ge20 with different dopants. In Proceedings of the 25th International conference of Thermoelectrics, Vienna, Austria, 6–10 August 2006; pp. 272–275.
  19. Pisharody, P.K.; Garvey, L.P. Modified silicon-germanium alloys with improved performance. In Proceedings of the 7th Intersociety Energy Conversion Engineering Conference, San Diego, CA, USA, 20–25 August 1978; IEEE: New York, NY, USA, 1978; pp. 1963–1968. [Google Scholar]
  20. Ravich, Y.I. Selective Carrier Scattering in Thermoelectric Materials. In CRC Handbook of Thermoelectrics; Rowe, D.M., Ed.; CRC Press: Boca Raton, FL, USA, 1995; pp. 67–81. [Google Scholar]
  21. Heremans, J.P.; Jovovic, V.; Toberer, E.S.; Saramat, A.; Kurosaki, K.; Chroenphakdee, A.; Yamanaka, S.; Snyder, G.F. Enhancement of thermoelectric performance in PbTe by distortion of the electronic density of states. Science 2008, 321, 554–557. [Google Scholar] [CrossRef] [PubMed]
  22. Jeong, C.; Lundstorm, M. On electronic structure engineering and thermoelectric performance. J. Electron. Mater. 2011, 40, 738–743. [Google Scholar] [CrossRef]
  23. Bergman, D.J.; Fel, L.G. Enhancement of thermoelectric power factor in composite thermoelectrics. J. Appl. Phys. 1999, 85, 8205–8216. [Google Scholar] [CrossRef]
  24. Zehbarjadi, M.; Joshi, G.; Zhu, G.; Yu, B.; Minnich, A.; Lan, Y.C.; Wang, X.; Dresselhaus, M.S.; Ren, Z.F.; Chen, G. Power Factor Enhancement by Modulation Doping in Bulk Nanocomposites. Nano Lett. 2011, 11, 2225–2230. [Google Scholar] [CrossRef] [PubMed]
  25. Ji, X.; He, J.; Su, Z.; Gothard, N.; Tritt, T.M. Improved thermoelectric performance in polycrystalline p-type Bi2Te3 via an alkali metal salt hydrothermal nanocoating treatment approach. J. Appl. Phys. 2008, 104, 0349071–0349076. [Google Scholar] [CrossRef]
  26. Ji, X.; Zhang, B.; Su, Z.; Holgate, T.; He, J.; Tritt, T.M. Nanoscale granular boundaries in polycrystalline Pb0.75Sn0.25Te: An innovative approach enhance the thermoelectric figure of merit. Phys. Status Solidi A 2009, 206, 221–228. [Google Scholar] [CrossRef]
  27. Martelli, P.; Caputo, R.; Remhof, A.; Mauron, P.; Borgschulte, A.; Zuttel, A. Stability and Decomposition of NaBH4. J. Phys. Chem. C 2010, 114, 7173–7177. [Google Scholar] [CrossRef]
  28. Dean, J.A. Lange’s Handbook of Chemistry; McGraw-Hill Inc.: New York, NY, USA, 1999; pp. 313–323. [Google Scholar]
  29. Nag, B.R. Electron Transport in Compound Semiconductors; Springer: Berlin/Heidelberg, Germany, 1980; p. 30. [Google Scholar]
  30. Klemens, P.G. Phonon scattering and thermal resistance due to grain boundaries. Int. J. Thermophysics 1994, 15, 1345–1351. [Google Scholar] [CrossRef]
  31. Pope, A.L.; Zawilski, B.M.; Tritt, T.M. Description of removable sample mount apparatus for rapid thermal conductivity measurements. Cryogenics 2001, 14, 725–731. [Google Scholar] [CrossRef]
  32. Pope, A.L.; Littleton, R.T., IV; Tritt, T.M. Apparatus for the rapid measurement of electrical transport properties for both “needle-like” and bulk materials. Rev. Sci. Instrum. 2001, 72, 3129–3131. [Google Scholar] [CrossRef]

Share and Cite

MDPI and ACS Style

Lahwal, A.; Zeng, X.; Bhattacharya, S.; Zhou, M.; Hitchcock, D.; Karakaya, M.; He, J.; Rao, A.M.; Tritt, T.M. Enhancing Thermoelectric Properties of Si80Ge20 Alloys Utilizing the Decomposition of NaBH4 in the Spark Plasma Sintering Process. Energies 2015, 8, 10958-10970. https://doi.org/10.3390/en81010958

AMA Style

Lahwal A, Zeng X, Bhattacharya S, Zhou M, Hitchcock D, Karakaya M, He J, Rao AM, Tritt TM. Enhancing Thermoelectric Properties of Si80Ge20 Alloys Utilizing the Decomposition of NaBH4 in the Spark Plasma Sintering Process. Energies. 2015; 8(10):10958-10970. https://doi.org/10.3390/en81010958

Chicago/Turabian Style

Lahwal, Ali, Xiaoyu Zeng, Sriparna Bhattacharya, Menghan Zhou, Dale Hitchcock, Mehmet Karakaya, Jian He, Apparao M. Rao, and Terry M. Tritt. 2015. "Enhancing Thermoelectric Properties of Si80Ge20 Alloys Utilizing the Decomposition of NaBH4 in the Spark Plasma Sintering Process" Energies 8, no. 10: 10958-10970. https://doi.org/10.3390/en81010958

Article Metrics

Back to TopTop