Increased Phonon Scattering by Nanograins and Point Defects in Nanostructured Silicon with a Low Concentration of Germanium

Similar documents
Thermionic power generation at high temperatures using SiGe/ Si superlattices

Charge Transport and Thermoelectric Properties of P-type Bi 2-x Sb x Te 3 Prepared by Mechanical Alloying and Hot Pressing

Thermal conductivity of GaAs/Ge nanostructures

Reduced Lattice Thermal Conductivity in Bi-doped Mg 2 Si 0.4 Sn 0.6

New Directions for Low-Dimensional Thermoelectric Materials**

Nanostructured Thermoelectric Materials: From Superlattices to Nanocomposites

Enhancement of thermoelectric figure-of-merit by resonant states of aluminium doping in lead selenide

Solar Thermoelectric Energy Conversion

University of California Postprints

Model of transport properties of thermoelectric nanocomposite materials

ENERGY NANOTECHNOLOGY --- A Few Examples

Thermoelectric materials. Presentation in MENA5010 by Simen Nut Hansen Eliassen

Thermal conductivity of symmetrically strained Si/Ge superlattices

Supplementary Information for On-chip cooling by superlattice based thin-film thermoelectrics

Supporting Information

Thermal Conductivity of Core-Shell Nanocomposites for Enhancing Thermoelectric Performance. S. J. Poon, A. S. Petersen, and Di Wu

Enhancing thermoelectric performance in hierarchically structured BiCuSeO by. increasing bond covalency and weakening carrier-phonon coupling

Thermoelectric power generators are clean, noise free

Half Heusler Alloys for Efficient Thermoelectric Power Conversion

Thermoelectric and Transport Properties of In-filled and Ni-doped CoSb 3 Skutterudites

Seebeck Enhancement Through Miniband Conduction in III V Semiconductor Superlattices at Low Temperatures

Thermal conductivity of bulk nanostructured lead telluride. in the view of phonon gas kinetics

Thermoelectric and mechanical properties of Ag_2Te and Sb_<1.6>Bi_<0.4>Te_3

Reduction of Thermal Conductivity by Nanoscale 3D Phononic Crystal

Control of Nano Structure by Multi Films for Nano-structured Thermoelectric Materials

Anisotropic thermoelectric properties of layered compound In 2 Te 5 single crystal

Nanowires. Department of Physics, Case Western Reserve University, Cleveland, Ohio 44106, United

Nanoporous Si as an Efficient Thermoelectric Material

Thermal transport in SiGe superlattice thin films and nanowires: Effects of specimen and periodic lengths

Nanoscale Heat Transfer and Information Technology

Supporting Information

Doping optimization for the power factor of bipolar thermoelectric materials. Abstract

Influence of Dimensionality on Thermoelectric Device Performance

efficiency can be to Carnot primarily through the thermoelectric figure of merit, z, defined by

Supporting Information

High-temperature thermoelectric behavior of lead telluride

A Facile Synthetic Approach for Copper Iron Sulfide. Nanocrystals with Enhanced Thermoelectric Performance

Supporting Information

Transient Harman Measurement of the Cross-plane ZT of InGaAs/InGaAlAs Superlattices with Embedded ErAs Nanoparticles

Improvement of the Thermoelectric Properties of (Sr 0.9 La 0.1 ) 3 Ti 2 O 7 by Ag Addition

Using anodic aluminum oxide templates and electrochemical method to deposit BiSbTe-based thermoelectric nanowires

Clean Energy: Thermoelectrics and Photovoltaics. Akram Boukai Ph.D.

Recent Progress in Nanostructured Thermoelectric Materials

Exact Solution of a Constrained. Optimization Problem in Thermoelectric Cooling

Introduction of Nano Science and Tech. Thermal and Electric Conduction in Nanostructures. Nick Fang

Supporting information:

Very high thermoelectric power factor in a Fe 3 O 4 /SiO 2 /p-type Si(100) heterostructure

Semester Length Glass Courses and Glass Schools

status solidi Enhancement of thermoelectric performance in strongly correlated multilayered nanostructures

Supplementary Figure 1 Characterization of the synthesized BP crystal (a) Optical microscopic image of bulk BP (scale bar: 100 μm).

Device Testing and Characterization of Thermoelectric Nanocomposites

Predicting Thermoelectric Properties From First Principles

MOLECULAR DYNAMICS SIMULATION OF THERMAL CONDUCTIVITY OF NANOCRYSTALLINE COMPOSITE FILMS

TRANSIENT THERMAL CHARACTERIZATION OF ERAS/IN 0.53 GA 0.47 AS THERMOELECTRIC MODULE

Thermoelectric Oxide Materials For Electric Power Generation

Computational Study of the Electronic Performance of Cross-Plane Superlattice Peltier Devices

Ceramic Processing Research

Multiscale phonon blocking in Si phononic crystal nanostructures

Models for the electronic transport properties of thermoelectric materials. Lars Corbijn van Willenswaard

CHAPTER 3. OPTICAL STUDIES ON SnS NANOPARTICLES

Characterization of thermal conductivity of La 0.95 Sr 0.05 CoO 3 thermoelectric oxide nanofibers

n-type to p-type crossover in quaternary Bi x Sb y Pb z Se 3 single crystals

Challenges and Opportunities for Condensed Matter Physics of Thermoelectric Materials

Phonon Drag Effect in Nanocomposite FeSb 2

Thermal conductivity: An example of structure-property relations in crystals Ram Seshadri

Nanostructured thermoelectric materials: Current research and future challenge

Exact Solution of a Constrained. Optimization Problem in Thermoelectric Cooling

O 3. : Er nanoparticles prospective system for energy convertors

Substitution of Bi for Sb and its Role in the Thermoelectric Properties and Nanostructuring in Ag 1-x Pb 18 MTe 20 (M ) Bi, Sb) (x ) 0, 0.14, 0.

Quantum transport simulations for the thermoelectric power factor in 2D nanocomposites

Electronic Structures and Thermoelectric Properties of Two-Dimensional MoS 2 /MoSe 2 Heterostructures

Quantum Dot Superlattice Thermoelectric Materials and Devices. T. C. Harman, P. J. Taylor, M. P. Walsh, and B. E. LaForge. Supplementary Material

Modeling of thin-film solar thermoelectric generators

Thermoelectric properties of Sr 0.61 Ba 0.39 Nb 2 O 6-δ ceramics annealed in different oxygen-reduction conditions

Phonon Coherent Resonance and Its Effect on Thermal Transport In. Core-Shell Nanowires

A STUDY OF THERMOELECTRIC PROPERTIES OF GRAPHENE MATERIALS

Studies on Thermoelectric Properties of n-type Polycrystalline SnSe[subscript 1-x]S[subscript x] by Iodine Doping

Theoretical efficiency of solar thermoelectric energy generators

Segmented Power Generator Modules of Bi 2 Te 3 and ErAs:InGaAlAs Embedded with ErAs Nanoparticles

Anisotropy in Thermoelectric Properties of CsBi 4 Te 6

SUPPLEMENTARY INFORMATION

PREPARATION OF LUMINESCENT SILICON NANOPARTICLES BY PHOTOTHERMAL AEROSOL SYNTHESIS FOLLOWED BY ACID ETCHING

High Thermoelectric Figure of Merit by Resonant Dopant in Half-Heusler Alloys

Estimating Energy Conversion Efficiency of Thermoelectric Materials: Constant Property Versus Average Property Models

Recap (so far) Low-Dimensional & Boundary Effects

Supplementary Figure 1. Selected area electron diffraction (SAED) of bilayer graphene and tblg. (a) AB

Transient Harman Measurement of the Cross-plane ZT of InGaAs/InGaAlAs Superlattices with Embedded ErAs Nanoparticles

A Universal Gauge for Thermal Conductivity of Silicon Nanowires. With Different Cross Sectional Geometries

STRUCTURE AND MAGNETIC PROPERTIES OF SiO 2 COATED Fe 2 NANOPARTICLES SYNTHESIZED BY CHEMICAL VAPOR CONDENSATION PROCESS

doi: /

Calculation of Confined Phonon Spectrum in Narrow Silicon Nanowires Using the Valence Force Field Method

Hole-doping effect on the thermoelectric properties and electronic structure of CoSi

Thermal Transport in Graphene and other Two-Dimensional Systems. Li Shi. Department of Mechanical Engineering & Texas Materials Institute

Comparison of solid-state thermionic refrigeration with thermoelectric refrigeration

Functional properties

Surface Transfer Doping of Diamond by Organic Molecules

SUPPORTING INFORMATION. Promoting Dual Electronic and Ionic Transport in PEDOT by Embedding Carbon Nanotubes for Large Thermoelectric Responses

QUANTUM SIMULATION OF NANOCRYSTALLINE COMPOSITE THERMOELECTRIC PROPERTIES

Supporting Information

Report on 7th US-Japan Joint Seminar on Nanoscale Transport Phenomena Science and Engineering

Transcription:

Increased Phonon Scattering by Nanograins and Point Defects in Nanostructured Silicon with a Low Concentration of Germanium The MIT Faculty has made this article openly available. Please share how this access benefits you. Your story matters. Citation As Published Publisher Zhu, G. H. et al. Increased Phonon Scattering by Nanograins and Point Defects in Nanostructured Silicon with a Low Concentration of Germanium. Physical Review Letters 102.19 (2009): 196803. http://dx.doi.org/10.1103/physrevlett.102.196803 American Physical Society Version Final published version Accessed Tue Nov 20 18:31:40 EST 2018 Citable Link Terms of Use Detailed Terms http://hdl.handle.net/1721.1/50996 Article is made available in accordance with the publisher's policy and may be subject to US copyright law. Please refer to the publisher's site for terms of use.

Increased Phonon Scattering by Nanograins and Point Defects in Nanostructured Silicon with a Low Concentration of Germanium G. H. Zhu, 1 H. Lee, 2 Y. C. Lan, 1 X. W. Wang, 1 G. Joshi, 1 D. Z. Wang, 1 J. Yang, 1 D. Vashaee, 3 H. Guilbert, 1 A. Pillitteri, 1 M. S. Dresselhaus, 4 G. Chen, 2, * and Z. F. Ren 1, * 1 Department of Physics, Boston College, Chestnut Hill, Massachusetts 02467, USA 2 Department of Mechanical Engineering, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, USA 3 Department of Electrical and Computer Engineering, Oklahoma State University, Tulsa, Oklahoma 74106, USA 4 Department of Electrical Engineering and Computer Science, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, USA (Received 26 November 2008; published 14 May 2009) The mechanism for phonon scattering by nanostructures and by point defects in nanostructured silicon (Si) and the silicon germanium (Ge) alloy and their thermoelectric properties are investigated. We found that the thermal conductivity is reduced by a factor of 10 in nanostructured Si in comparison with bulk crystalline Si. However, nanosize interfaces are not as effective as point defects in scattering phonons with wavelengths shorter than 1 nm. We further found that a 5at: % Ge replacing Si is very efficient in scattering phonons shorter than 1 nm, resulting in a further thermal conductivity reduction by a factor of 2, thereby leading to a thermoelectric figure of merit 0.95 for Si 95 Ge 5, similar to that of large grained Si 80 Ge 20 alloys. DOI: 10.1103/PhysRevLett.102.196803 Solid state energy conversion between heat and electricity based on thermoelectric effects is attractive in waste heat recovery and environmentally friendly refrigeration [1]. The conversion efficiency depends on the dimensionless thermoelectric figure of merit ZT ¼ðS 2 =ÞT, where S,,, and T are the Seebeck coefficient, electrical conductivity, thermal conductivity, and absolute temperature, respectively [1]. Good thermoelectric materials behave as crystals for electrons and glasses for phonons [2]. However, such materials are rare in nature and are not easily engineered in the laboratory. After the 1950s, ZT did not significantly improve for almost another 40 years. In the early 1990s, Hicks and Dresselhaus proposed the possibility to enhance ZT with nanostructures [3]. Recently, a number of studies reported high ZT values using nanostructures [4 8]. In these studies, the ZT enhancement in nanostructures was mostly due to their low thermal conductivity, which is attributed to phonon scattering by their large density of interfaces [9,10]. Among the various nanostructured materials, the nanostructured composite (nanocomposite) approach [7,8,10] seems to be the best. For nanocomposites, when the grain size is smaller than the mean free path of a phonon, additional phonon scattering at boundaries will occur and the thermal conductivity is thereby reduced. The idea of reducing the thermal conductivity with smaller grains has been suggested since the 1980s [11], but most experimental efforts failed because the small grains also reduced the power factor S 2. Only recently, has a noticeable enhancement in ZT been achieved for the AgPbSbTe composite system [12], and for p-type BiSbTe [7,8], n-type Si 80 Ge 20 [13] and p-type Si 80 Ge 20 [14] systems. However, the reduction in from PACS numbers: 73.50.Lw, 84.60.Rb, 85.80.Fi the alloying comes through scattering by point defects, while for the nanosize effect it comes from the strong interface scattering of phonons. These two causes for phonon scattering could not be separated from one another in these studies. This distinction is the focus of the present Letter. In focusing on the mechanisms of the reduction in the thermal conductivity in nanograined materials by comparison of the phonon scattering processes in pure Si and in Si containing a low Ge concentration, we note that pure Si does not have point defect scattering from Ge, and hence offers an opportunity to study the scattering of grain boundaries, while the addition of Ge increases point defect scattering. In the present work we find experimentally that nanograins play a very important role in increasing the phonon scattering for phonons with wavelengths in the nanometer range. However, point defect scattering, caused by alloying Ge into Si, is more effective in scattering phonons than just using pure Si nanostructures especially for scattering phonons with wavelengths of less than 1 nm. In the present work we show that a combination of nanograins and a 5at: % Ge replacement of Si, that is Si 95 Ge 5, produces both a reduction in the thermal conductivity and a similar ZT value to that of bulk Si 80 Ge 20 alloys. In our work, chunks of Si and Ge (Alfa Aesar), phosphorus (P) and gallium phosphide (GaP) (Sigma Aldrich) are pulverized into a powder until the desired nanosize particles are obtained [7,8,10,13,14]. P, a typical n-type dopant, introduced into our samples is 2:5 at: %, which is larger than the maximum bulk solubility limit of 1at: %. With the addition of GaP and the introduction of nanosized particles, the solubility limit of P in Si or SiGe increases [15]. X-ray diffraction (XRD), scanning electron micros- 0031-9007=09=102(19)=196803(4) 196803-1 Ó 2009 The American Physical Society

FIG. 1 (color online). XRD pattern (a), TEM image of ball milled Si 95 Ge 5 nanopowders (b). The insets to Fig. 1 (b) are the diffraction pattern and HRTEM image for the circled region. copy (SEM), and high resolution transmission electron microscopy (HRTEM) are used to characterize the structures of the samples. The nanopowder is pressed into disc specimens with a diameter of 12.7 mm and a thickness of a few mm by a quick dc hot press process [7,8,10,13,14]. The density of the hot pressed sample is measured using an Archimedes kit and the value is very close to the theoretical density [16]. The thermal conductivity is measured using a laser flash system (Netzsch LFA457), and the electrical conductivity and Seebeck coefficient are measured simultaneously in a multiprobe transport system (Ulvac ZEM-3). The carrier concentration is measured using the van der Pauw method [17]. Figure 1(a) shows the XRD spectrum of our Si 95 Ge 5 P 2:5 ðgapþ 1:5 nanopowders. It shows that the nanopowders have a single phase with broadened peaks indicating the nanosize of the grains in the range of 5 20 nm as obtained by the Williamson-Hall method [18]. Figure 1(b) shows the TEM image of the nanopowders around 20 150 nm, representing the size of the agglomeration of many smaller crystallites, indicated in the diffraction pattern [upper inset in Fig. 1(b)]. The HRTEM image [lower inset in Fig. 1(b)] confirms the small size of the constituent particles. Figure 2 shows the TEM images for the Si 95 Ge 5 P 2:5 ðgapþ 1:5 sample after hot press, where most of the nanoparticles are in the 10 30 nm range, larger than the 5 20 nm initial nanoparticle range due to some grain growth during the hot press process. The HRTEM image [Fig. 2(b)] shows that the nanoparticles are highly crystallized and randomly oriented after hot press. Figures 3(a) 3(f) show the comparative thermoelectric property measurement results for nanostructured Si, nanostructured Si 95 Ge 5, bulk Si, and bulk Si 80 Ge 20 alloy (RTG) that has been used by NASA for many years. Results plotted for the bulk Si are calculations using the Boltzmann transport equation within the relaxation time approximation, modified on the basis of the Vining model [19] but also considering nonparabolicity and a temperature-dependent band structure. Both the nanostructured Si and nanostructured Si 95 Ge 5 samples show a higher electrical conductivity [Fig. 3(a)] but a lower absolute Seebeck coefficient [Fig. 3(b)] than that of the bulk Si 80 Ge 20 RTG sample. This is mainly attributed to the FIG. 2. Low (a) and high (b) magnification TEM images of the hot pressed nanostructured Si 95 Ge 5 sample. 196803-2

FIG. 3 (color online). Temperature-dependent electrical conductivity (a), Seebeck coefficient (b), power factor (c), thermal conductivity (d), electron (K e ), phonon (K l ), and total (K) thermal conductivity by modeling (e), and ZT (f) of nanostructured Si (filled squares), nanostructured Si 95 Ge 5 (filled circles for experiment and solid line for model), bulk Si model (dashed line), and Si 80 Ge 20 RTG samples (open circles). higher solubility limit of P and the lower alloy scattering of charge carriers in nanostructured Si and nanostructured Si 95 Ge 5 samples in comparison with the bulk Si 80 Ge 20 RTG sample. The power factors for both nanostructured samples [Fig. 3(c)] are slightly lower than the values calculated for bulk materials with the same carrier concentration values as measured for the nanostructured samples. Also, due to the heavy doping in the nanostructured Si 95 Ge 5 and Si samples and the activation of excess dopant (P) at high temperature during the measurement process, our nanostructured Si and nanostructured Si 95 Ge 5 samples show different trends for the temperature-dependent electrical conductivity and Seebeck coefficient from those of the bulk Si. In fact, the power factor of the nanostructured Si 95 Ge 5 sample is much higher than that of the bulk Si 80 Ge 20 RTG sample [Fig. 3(c)], especially at temperatures above 300 C. The main advantage of the nanostructure approach for Si 95 Ge 5 is that we can maintain the high electrical conductivity and power factor as shown in Figs. 3(a) and 3(c) and, at the same time, we can reduce the phonon thermal conductivity significantly. Such joint behavior does not occur in bulk samples. Figure 3(d) shows the temperature-dependent thermal conductivity of the nanostructured Si and nanostructured Si 95 Ge 5 samples in comparison with bulk Si and bulk Si 80 Ge 20 RTG samples. The thermal conductivity of the nanostructured Si shows a significant reduction (by about a factor of 10) compared with that of the heavily doped bulk Si, which is around 196803-3 100 W=m K at room temperature, a clear demonstration of the nanosize effect on phonon scattering. Moreover, with a 5at: % replacement of Si by Ge, the thermal conductivity value of the nanostructured Si 95 Ge 5 is even lower, close to that of the bulk Si 80 Ge 20 RTG sample, caused by both the nanosize and point defect scattering effects in nanostructured Si 95 Ge 5. Since the bulk Si 80 Ge 20 RTG sample has 20 at: % Ge, and our nano Si 95 Ge 5 sample has only 5at: % Ge, a weaker alloy phonon scattering effect is expected in Si 95 Ge 5. When the Ge concentration is increased from 5 to 20 at: %, the thermal conductivity is decreased by another factor of 2 to about 2 3 W=m K, but the power factor is also decreased [13] accordingly because of the reduced charge mobility due to the alloy scattering of charge carriers. The thermal conductivity of nanostructured Si 95 Ge 5 has also been investigated by modeling based on Callaway s model [20] in combination with a modified effective medium theory [21] to consider the effect of nanosized grains. Figure 4(a) shows the mean free path of phonons vs phonon wavelength. For bulk Si, only three-phonon scattering and electron-phonon scattering are the dominant scattering mechanisms. As a 5at: % of Ge is added, the scattering rate increases significantly due to point defect scattering. For pure Si, the mean free paths for most of the phonons will be limited by the small grain size. For Si 95 Ge 5, the small grain size significantly reduces the mean free path of phonons at long wavelengths. Figure 4(b) shows the accumulative thermal conductivity normalized to the thermal conductivity of bulk Si as the phonon wavelength is increased. The thermal conductivity of pure Si is reduced almost by an order of magnitude using nanograins, since a 20 nm grain can reduce the mean free path of phonons at almost all wavelengths. By adding a 5at: % Ge, alloy scattering can reduce the thermal conductivity more significantly than using nanograins in Si. As shown in Fig. 4(a), the mean free path by alloy scattering is even smaller than the grain size (20 nm) for phonon wavelengths less than 1 nm, and the contribution of short wavelength phonons is large. Nanosized grains in Si 95 Ge 5 can further reduce the thermal conductivity by limiting the mean free path of phonons with wavelengths larger than 1 nm. As FIG. 4 (color online). Modeling results for the thermal conductivity at room temperature: (a) mean free path vs phonon wavelength and (b) accumulative thermal conductivity ratio vs phonon wavelength for different Ge ratios and grain sizes.

shown in Fig. 4(b), the thermal conductivity of nanograined Si 95 Ge 5 is reduced by a factor of 2 from bulk Si 95 Ge 5. Thus, by using the nanostructures and also adding only a small amount of Ge, the thermal conductivity can be reduced to as low a value as for a SiGe alloy with a much higher Ge ratio. Figure 3(e) shows that the calculated thermal conductivity of nanostructured Si 95 Ge 5 matches well with the experimental results [Fig. 3(d)]. The electron contribution to the thermal conductivity is calculated from the electrical conductivity measurement results using the Wiedemann- Franz Law. The Lorenz number is calculated from the bulk model. Our modeling results show that the Lorenz number in bulk SiGe alloy varies from 1.3 to 2.2 from 25 C to 1000 C, and that variation within any specific temperature is 0.2 for the range of the doping concentration in our samples. The calculated phonon thermal conductivity dropped below 4W=mK at room temperature and reached 3 W=m K at 900 C [Fig. 3(e)]. The low thermal conductivity for the nanostructured Si 95 Ge 5 system is mainly attributed to both the enhanced boundary phonon scattering and the alloy effect. Thus, due to the significant thermal conductivity reduction without reducing the power factor, ZT of the nanostructured Si 95 Ge 5 shows a maximum value of 0.95 at 900 C, which is about the same as that of the bulk Si 80 Ge 20 RTG sample [Fig. 3(f)]. While phonon scattering at the grain boundary can be explained by a modified effective medium theory, electron scattering due to a grain boundary with nanosize particles has not yet been well investigated. Our measurement results show that the electrical conductivity is slightly lower than the value expected for the given carrier concentration. There may be different explanations for the additional carrier scattering caused by grain boundaries, but the most plausible reason might be the electron potential variation at the grain boundaries due to defect sites and/ or dopant precipitation at boundaries. Since our doping concentration is higher than the solubility limit, excess amount of dopants must be precipitated somewhere. A previous study suggested that P is likely to form a compound such as SiP which precipitates at the grain boundary [22]. Since the composition is different between the grain boundary region and the grain region, an electron potential difference will occur. These two effects can also happen in bulk Si, but the effect is greater in nanostructured Si 95 Ge 5 due to a higher boundary density. In summary, we have achieved an enhancement in ZT by a factor of 2 in nanostructured Si and of almost a factor of 4 in nanostructured Si 95 Ge 5 in comparison with bulk Si. The enhancement is mainly due to the reduction in the thermal conductivity by the increased scattering of intermediate wavelength phonons at the nanosized grains and by the point defect scattering of short wavelength phonons. It is clearly demonstrated that phonons with different wavelengths need to be matched with similar size scatterers so that effective phonon scattering can take place to achieve the lowest possible thermal conductivity. 196803-4 We would like to thank Drs. Jean-Pierre Fleurial and Pawan Gogna for providing the data for the RTG sample. The work was sponsored by DE-FG02-08ER46516 (M. D., G. C., and Z. F. R.), NSF NIRT 0506830 (M. D., G. C., and Z. F. R.), DOE DE-FG02-00ER45805 (Z. F. R.), and NSF CMMI 0833084 (Z. F. R. and G. C.). *To whom correspondence should be addressed. gchen2@mit.edu (G. C.); renzh@bc.edu (Z. F. R.) [1] C. Wood, Rep. Prog. Phys. 51, 459 (1988). [2] G. Slack, in CRC Handbook of Thermoelectrics, edited by D. M. Rowe (CRC Press, Boca Raton, FL 1996). [3] L. D. Hicks and M. S. Dresselhaus, Phys. Rev. B 47, 16 631 (1993). [4] R. Venkatasubramanian, E. Siivola, T. Colpitts, and B. O Quinn, Nature (London) 413, 597 (2001). [5] T. C. Harman, P. J. Taylor, M. P. Walsh, and B. E. LaForge, Science 303, 818 (2004). [6] A. I. Hochbaum, R. Chen, R. D. Delgado, W. Liang, E. C. Garnett, M. Najarian, A. Majumdar, and P. D. Yang, Nature (London) 451, 163 (2008); A. I. Boukai, Y. Bunimovich, J. Tahir-Kheli, J. Yu, W. A. Goddard, and J. R. Heath, Nature (London) 451, 168 (2008). [7] B. Poudel, Q. Hao, Y. Ma, Y. C. Lan, A. Minnich, B. Yu, X. Yan, D. Z. Wang, A. Muto, D. Vashaee, X. Chen, J. M. Liu, M. S. Dresselhaus, G. Chen, and Z. F. Ren, Science 320, 634 (2008). [8] Y. Ma, Q. Hao, B. Poudel, Y. C. Lan, B. Yu, D. Z. Wang, G. Chen, and Z. F. Ren, Nano Lett. 8, 2580 (2008). [9] R. G. Yang and G. Chen, Phys. Rev. B 69, 195316 (2004). [10] M. S. Dresselhaus, G. Chen, Z. F. Ren, J. P. Fleurial, and P. Gogna, Adv. Mater. 19, 1043 (2007). [11] N. Savvides and H. J. Goldsmid, J. Phys. C 13, 4671 (1980). [12] F. K. Hsu, S. Loo, F. Guo, W. Chen, J. S. Dyck, C. Uher, T. Hogan, E. K. Polychroniadis, and M. G. Kanatzidis, Science 303, 818 (2004). [13] X. W. Wang, H. Lee, Y. C. Lan, G. H. Zhu, G. Joshi, D. Z. Wang, J. Yang, A. J. Muto, M. Y. Tang, J. Klatsky, S. Song, M. S. Dresselhuas, G. Chen, and Z. F. Ren, Appl. Phys. Lett. 93, 193121 (2008). [14] G. Joshi, H. Lee, Y. C. Lan, X. W. Wang, G. H. Zhu, D. Z. Wang, A. J. Muto, M. Y. Tang, M. S. Dresselhuas, G. Chen, and Z. F. Ren, Nano Lett. 8, 4670 (2008). [15] O. Yamashita, J. Appl. Phys. 89, 6241 (2001). [16] J. P. Dismukes, L. Ekstrom, and R. J. Paff, J. Phys. Chem. 68, 3021 (1964). [17] C. Wood, A. Lockwood, A. Chmielewski, J. Parker, and A. Zoltan, Rev. Sci. Instrum. 55, 110 (1984). [18] G. K. Williamson and W. H. Hall, Acta Metall. 1, 22 (1953). [19] C. B. Vining, J. Appl. Phys. 69, 331 (1991). [20] J. Callaway and H. C. von Baeyer, Phys. Rev. 120, 1149 (1960). [21] A. Minnich and G. Chen, Appl. Phys. Lett. 91, 073105 (2007). [22] B. A. Cook, J. L. Harringa, S. H. Han, and C. B. Vining, J. Appl. Phys. 78, 5474 (1995).