7
Seebeck and Figure of Merit Enhancement in Nanostructured Antimony Telluride by Antisite Defect Suppression through Sulfur Doping Rutvik J. Mehta, Yanliang Zhang, Hong Zhu, § David S. Parker, Matthew Belley, David J. Singh, Ramamurthy Ramprasad, § Theodorian Borca-Tasciuc, and Ganpati Ramanath* ,Materials Science and Engineering Department and Mechanical, Aerospace & Nuclear Engineering Department, Rensselaer Polytechnic Institute, Troy, New York 12180, United States § Chemical, Materials & Biomolecular Engineering Department, University of Connecticut, Storrs, Connecticut 06269, United States Materials Science & Technology Division, Oak Ridge National Laboratory, Oak Ridge, Tennessee 37831-6056, United States * S Supporting Information ABSTRACT: Antimony telluride has a low thermoelectric gure of merit (ZT < 0.3) because of a low Seebeck coecient α arising from high degenerate hole concentrations generated by antimony antisite defects. Here, we mitigate this key problem by suppressing antisite defect formation using subatomic percent sulfur doping. The resultant 1025% higher α in bulk nanocrystalline antimony telluride leads to ZT 0.95 at 423 K, which is superior to the best non-nanostructured antimony telluride alloys. Density functional theory calculations indicate that sulfur increases the antisite formation activation energy and presage further improvements leading to ZT 2 through optimized doping. Our ndings are promising for designing novel thermoelectric materials for refrigeration, waste heat recovery, and solar thermal applications. KEYWORDS: Nanobulk thermoelectrics, sulfur doping, antimony telluride, antisite defects, rst principle transport calculations, high gure of merit ZT T hermoelectric materials oer promise for realizing trans- formative solid-state refrigeration technologies that could replace extant ones based on liquid coolants 1 and allow the ecient conversion of waste heat to electricity. The fruition of this vision, however, requires bulk thermoelectric materials with gures of merit ZT > 1 in the 200 T 800 K range, necessitating high Seebeck coecient α, high electrical conductivity σ, and low thermal conductivity κ. Pnictogen chalcogenides (VVI compounds) and their alloys are well- suited for realizing high ZT due to their low band gaps, complex crystal structures, and the presence of heavy elements. 2 Antimony telluride is a p-type semiconductor with σ > 300 kΩ 1 ·m 1 , but the ZT < 0.3 due to low α arising from self- compensating antisite defects. 35 In particular, Sb atoms occupying Te sitesdenoted as SbTe create acceptor states resulting in high hole concentrations, 68 for example, h > 10 20 cm 3 , that in turn leads to low α. Self-alloying with >10 at. % Te has been shown to yield ZT 0.49 along the [0001] direction 8 of single crystals, but the direction-averaged ZT remained <0.3. Alloying with 25 at. % Bi 79 not only increases α but also decreases κ due to increased impurity scattering of phonons, leading to the highest room- temperature ZT 300K 1 reported for any bulk p-type material. However, such high impurity content decreases the bandgap and precludes the use of this material above 450 K. Lower alloying levels (e.g., <10 at. % of Tl or In) 3,10,11 have been shown to decrease h in single-crystal Sb 2 Te 3 but have not resulted in ZT increases. Dopant additions (e.g., <2 at. %) known to increase ZT in the PbTe system 12 have also been of limited utility for controlling the carrier concentration in Sb 2 Te 3 due to the self- compensating defect chemistry in this material. Here, we demonstrate that subatomic-percent sulfur doping of nanostructured Sb 2 Te 3 can result in 25% greater α than the highest reported for the non-nanostructured Sb 2 Te 3 . We show that sulfur doping increases the SbTe antisite defect formation activation energy, thus mitigating the key problem with this material. The resultant hole concentration decrease and possibly an increased valence band density of states near the Fermi level underpin the α enhancement mechanism. The combined eect of sulfur doping and nanostructuring leads to a three times higher ZT than non-nanostructured antimony telluride and holds promise for obtaining ZT 2 by optimized doping. Our ndings are of importance for designing high ZT thermoelectric materials Received: May 1, 2012 Revised: August 7, 2012 Published: August 14, 2012 Letter pubs.acs.org/NanoLett © 2012 American Chemical Society 4523 dx.doi.org/10.1021/nl301639t | Nano Lett. 2012, 12, 45234529

(2012 Nano Lett) Seebeck and Figure of Merit Enhancement in Nanostructued Antimony Telluride by Antisite Defect Suppression Through Sulfer Doping

Embed Size (px)

DESCRIPTION

an interesting work on thermoelectric research

Citation preview

  • Seebeck and Figure of Merit Enhancement in NanostructuredAntimony Telluride by Antisite Defect Suppression through SulfurDopingRutvik J. Mehta, Yanliang Zhang, Hong Zhu, David S. Parker, Matthew Belley, David J. Singh,

    Ramamurthy Ramprasad, Theodorian Borca-Tasciuc, and Ganpati Ramanath*,

    Materials Science and Engineering Department and Mechanical, Aerospace & Nuclear Engineering Department, RensselaerPolytechnic Institute, Troy, New York 12180, United StatesChemical, Materials & Biomolecular Engineering Department, University of Connecticut, Storrs, Connecticut 06269, United StatesMaterials Science & Technology Division, Oak Ridge National Laboratory, Oak Ridge, Tennessee 37831-6056, United States

    *S Supporting Information

    ABSTRACT: Antimony telluride has a low thermoelectric gure of merit(ZT < 0.3) because of a low Seebeck coecient arising from highdegenerate hole concentrations generated by antimony antisite defects. Here,we mitigate this key problem by suppressing antisite defect formation usingsubatomic percent sulfur doping. The resultant 1025% higher in bulknanocrystalline antimony telluride leads to ZT 0.95 at 423 K, which issuperior to the best non-nanostructured antimony telluride alloys. Densityfunctional theory calculations indicate that sulfur increases the antisiteformation activation energy and presage further improvements leading to ZT 2 through optimized doping. Our ndings are promising for designingnovel thermoelectric materials for refrigeration, waste heat recovery, andsolar thermal applications.

    KEYWORDS: Nanobulk thermoelectrics, sulfur doping, antimony telluride, antisite defects, rst principle transport calculations,high gure of merit ZT

    Thermoelectric materials oer promise for realizing trans-formative solid-state refrigeration technologies that couldreplace extant ones based on liquid coolants1 and allow theecient conversion of waste heat to electricity. The fruition ofthis vision, however, requires bulk thermoelectric materials withgures of merit ZT > 1 in the 200 T 800 K range,necessitating high Seebeck coecient , high electricalconductivity , and low thermal conductivity . Pnictogenchalcogenides (VVI compounds) and their alloys are well-suited for realizing high ZT due to their low band gaps, complexcrystal structures, and the presence of heavy elements.2

    Antimony telluride is a p-type semiconductor with > 300k1m1, but the ZT < 0.3 due to low arising from self-compensating antisite defects.35 In particular, Sb atomsoccupying Te sitesdenoted as SbTecreate acceptor statesresulting in high hole concentrations,68 for example, h > 1020

    cm3, that in turn leads to low .Self-alloying with >10 at. % Te has been shown to yield ZT

    0.49 along the [0001] direction8 of single crystals, but thedirection-averaged ZT remained

  • by using a combination of doping and nanostructuring forrefrigeration, waste heat recovery, and solar thermal applications.We synthesized sulfur-doped nanoplates of Sb2Te3 by a

    microwave-stimulated solvothermal approach13,14 (Figure 1aTEM inset). The nanocrystals were consolidated into nano-structured bulk pellets by cold compaction and sintering (Figure1a inset and b). X-ray diractograms of the nanobulk pelletsconrm the retention of the rhombohedral R3m structure, withno observable traces of extraneous phases (Figure 1a). Energyand wavelength dispersive X-ray spectroscopy (EDX and WDX)conrm 0.10.5 at. % sulfur and stoichiometric Sb2Te3 with 401 at. % Sb and 60 1 at. % Te. The sintered nanobulk pelletshave 92 3% of the theoretical density. We measured and of the nanobulk pellets by a steady state method described indetail elsewhere.14,15 A four-probe tool in the van der Pauwconguration was used to measure . Our measurements of , ,and along axial and radial directions yielded identical valueswithin instrumental uncertainty, conrming the isotropic

    properties of our nanobulk Sb2Te3. Hole concentrations h andmobility were determined from Hall measurements.We obtain high room-temperature electrical conductivity in

    the 90 150 k1m1 range (see Figure 1c), comparable tothe lower ranges reported for single crystal Sb2Te3, buthundred- to thousand-fold higher than that of other reports onnanocrystal assemblies.16,17 Our nanobulk Sb2Te3 show 110 135 VK1, which is more than 50% higher than that of non-nanostructured stoichiometric Sb2Te3 and approximately 1025% higher than the highest observed for o-stoichiometricnon-nanostructured Sb2Te3. Our isotropic nanobulk Sb2Te3 has10% higher than the highest in single crystals with similarcarrier concentration but lower than that of p-type Sb-rich SbBiTe alloys.7,8 Previous reports of nanostructured Sb2Te3 withunknown or nondeliberate doping have either not shownincreases in 18 (Figure 2b) or the increase compromises signicantly. For example, microwave synthesized nanostruc-tured Sb2Te3 with unknown doping chemistry

    19,20 reportssimilar or higher as exhibited by our S-doped Sb2Te3, but the

    Figure 1. (a) X-ray diractogram from nanostructured bulk pellets (see photograph inset) fabricated by cold-pressing and sintering S-doped Sb2Te3nanoplates (see TEM inset). Only peaks corresponding to the rhombohedral crystal structure of Sb2Te3 are seen; for brevity, only the major peaks areindexed. (b) TEMmicrograph showing nanogranular and nanoporous features in the sintered pellets. The thermoelectric properties (c) , , , and Land (d) ZT of our nanobulk S-doped Sb2Te3 plotted as a function of temperature. Color-coded arrows point to the ordinate to be used for reading thedierent plots. In d, the ZT values for undoped bulk Sb2Te3, and non-nanostructured Bi0.5Sb1.5Te3 alloys (non-nano alloydata from ref 33) are shownfor comparison.

    Nano Letters Letter

    dx.doi.org/10.1021/nl301639t | Nano Lett. 2012, 12, 452345294524

  • is more than 10- to 100-fold lower than both conventional non-nanostructured and our doped nanobulk Sb2Te3, whichtranslates to at least 23 fold lower ZT than ours.Our nanobulk Sb2Te3 exhibit room-temperature thermal

    conductivity in the 0.7 1.1 Wm1K1 range,14 which is4070% lower than that of the single- or polycrystallineSb2Te3. This decrease corresponds to about 6075%diminution in the lattice thermal conductivity L due to phononscattering by the 2050 nm nanograins14 and nanopores.21

    The high , ultralow , and enhanced in nanobulk Sb2Te3 yieldZT300K 0.75, which is more than 2- to 4-fold that of non-nanostructured poly- or single-crystalline antimony telluride7,8

    (Figure 1d).The ZT of our nanobulk Sb2Te3

    14 increases monotonicallybetween 300 T 423 K, reaching ZT423K = 0.95, which is35% higher than ZT423K 0.7 reported for non-nanostructuredSb1.5Bi0.5Te3 alloys.

    7,8 This ZT increase is mainly due to themonotonic increase with temperature, yielding values as high as423K 150 VK1 (Figure 1c). We expect to peak at highertemperatures than that of Sb1.5Bi0.5Te3 alloys and Bi2Te3, becauseof the higher temperature onset of bipolar conduction arisingfrom the larger bandgap of Sb2Te3. Both and remain relativelyunchanged within 20% in the 300 T 423 K range (seeFigure 1c). The weaker-than-typical temperature dependence of, reminiscent of degenerate doping implies signicant scatteringdue to grain boundaries and/or point defects, leaving open thescope for further ZT improvements by optimizing grain size and/or defect concentrations. The insensitivity of is expected abovethe Debye temperatureD (160 K for antimony telluride

    22) fornanostructured materials.14 These eects combined with thehigher endow our nanobulk Sb2Te3 with high ZT withoutalloying.

    Room-temperature Hall measurements on the nanobulkpellets reveal hole concentrations of 1019 h 1020 cm3 thatare about 10- to 100-fold smaller than that reported for bothpolycrystalline and single-crystal non-nanostructured Sb2Te3.

    7,8

    The hole mobilities of 75 250 cm2V1s1 are lower thanthat of single crystals as expected, but higher than non-nanostructured Sb-rich SbBiTe alloys7,8 (Figure 2a) probablybecause the low

  • where VTe and VSb are vacancies.3,4,26 The consequent high hole

    concentrations decrease . We propose that doping with sulfur,which has a higher electronegativity than Te, decreases h bysuppressing SbTe formation, thereby oering a means forincreasing . This hypothesis is supported by the sulfur core-levelband structure of the nanoplates and the nanobulk pelletsmeasured by X-ray photoelectron spectroscopy (XPS) asdescribed below.The S 2p band at 162 eV indicates thioligation of Sb2Te3

    nanoplate surfaces with thioglycolic acid (TGA) used in oursynthesis;14,27 unfettered TGA has a higher binding energy peakat 164 eV (Figure 3a). Synchrotron XPS of the nanobulk pelletsreveals three S 1s sub-bands at 2477, 2471, and 2468 eV whoseintensities scale with the sulfur doping level measured by WDX(Figure 3b). All three sub-bands correspond to S2 states (higheroxidation numbers are expected28 only above 2480 eV) and canbe identied with sulfur occupancy of specic sites based on theinverse correlation between valence electron density and core-level binding energy. Sulfur at the Te 3a site, bound to two layersof electropositive Sb (see Figure 3d), would have the highest

    electron density and the lowest core-level binding energy. Sulfur

    at the Sb 6c site bound to two layers of isovalent Te would have a

    higher binding energy. Sulfur at the Te 6c site would have the

    highest binding energy because of a heterocovalent bond with

    one layer of Sb and weakly bound to Te atoms in the adjacent

    layer of the R3m structure via van der Waals bonding.We note that only the Te 6c sub-band intensity of S 1s strongly

    correlates with increasing (Figure 3b), suggesting that this state

    exerts the largest inuence on the hole concentration. The Te 3a

    and Sb 6c sites appear to be lled only after the Te 6c sites are rst

    lled to a threshold concentration. Such hierarchical site

    occupancy reported in selenium-alloyed bismuth telluride8 may

    be associated with the ease of interlayer sulfur incorporation in

    the R3m structure and lattice strain eects. Considering x moles

    of sulfur and y moles of decrease in SbTe, and hence h, per moleof Sb2Te3 at equilibrium, we can write

    Figure 3. (a) Core-level S 2p band structure obtained by XPS from as-synthesized Sb2Te3 nanoplate powders. A baseline spectrum from neat TGA isshown for reference. (b) Core-level S 1s band structure obtained by synchrotron XPS from sintered nanostructured bulk pellets of S-doped Sb2Te3 withdierent values, shown alongside the spectra. The sub-bands are labeled withWycko sites that specify sulfur occupancy; see text for details. (c) Core-level Sb 3d band structure with sub-bands from unoxidized and oxidized Sb2Te3 and the overlapping O 1s indicated. (d) Schematic atomic structures ofSb2Te3 with Wycko site labels depicting mixed ionic-covalent bonded quintet layers held by interlayer van der Waals bonding in the undoped, defect-free conguration (left), with an antisite SbTe defect (center), and with sulfur doping (right). The Wycko site notations are indicated on the left. Weused supercells of these structures for our DFT calculations.

    Nano Letters Letter

    dx.doi.org/10.1021/nl301639t | Nano Lett. 2012, 12, 452345294526

  • Figure 4. Density of states computed using DFT calculations for (a) undoped and (b) sulfur-doped Sb2Te3. The energy scale is referred to the valenceband maximum and EF, arbitrarily set to zero, connotes the Fermi level. Sulfur doping increases the density of states in the valence band as can be clearlyseen in (c) a magnied view of superimposed density of states of undoped (gray) and S-doped (red). (d) Valence band spectra obtained by synchrotronXPS from our nanobulk S-doped Sb2Te3 pellets.

    Figure 5. (a) Planar [112 0] Seebeck coecient for three dierent carrier concentrations (in cm3), and peak ZT, calculated as a function oftemperature. The black circles represent experimentally measured in our isotropic nanobulk. (b) Pisarenko plot of the planar [112 0] and (c)estimated ZT of nanobulk Sb2Te3, both calculated by DFT as a function of hole concentration for various temperatures. Experimental data from sulfur-doped nanobulk Sb2Te3 at 300 K (black circles) and 400 K (red square) are also shown for comparison, along with 300 K data for bulk single-crystalSb2Te3 from ref 8 for two crystallographic directions [0001] (magenta diamonds) and planar [112 0] (purple triangles).

    Nano Letters Letter

    dx.doi.org/10.1021/nl301639t | Nano Lett. 2012, 12, 452345294527

  • + + + + +

    + + + +

    x

    y x y y h

    x y y

    S 2Sb 3Te

    Sb S (2 )Sb (2 )

    (1 )V (2 )V 3/2Te (g)

    Sb Te

    Sb Te Te

    Te Sb 2

    Although we expect x = y, our experimental result shows that theh decrease exceeds sulfur doping by 1520%; that is, y > x. Weattribute this greater-than-anticipated eectiveness of sulfurdoping to a higher activation energy3,10,29 for creating SbTedefects in sulfur-doped Sb2Te3, than its undoped counterpart.The increase in activation energy for SbTe formationES due tosulfur doping can be estimated by recognizing that

    + c c E E kT[Sb ] ( )exp [( )/( )]Te Sb S 0 s mwhere cSb and cS are Sb and S concentrations, respectively, Tm themelting point, and E0 the activation energy

    3,10,29 in undopedSb2Te3. Setting [SbTe] = h fromHall measurements of nanobulkpellets with dierent cSb and cS quantied by WDX, we obtainES 0.06 eV connoting a 17% increase over E0 0.35 eV. Thisresult is corroborated by density functional theory (DFT)calculations showing ES 0.05 eV for uncharged SbTe at theTe 6c site (see Figure 3d). Calculations with singly chargedantisite defects result in ES 0.17 eV, suggesting that theantisite defects are predominantly uncharged. This result is alsoconsistent with all of the Sb 3d5/2 core-level bands beingattributable to SbTe bonds14 (Figure 3c) and residual surfaceoxide and the lack of charged state signatures expected at >530eV. We are thus persuaded to conclude that subatomic percentdoping of substitutional sulfursupposed to be electronicallyinactive in pnictogen chalcogenides26is more eective intuning carrier concentration than adding tens of atomic percentof alloying elements11,23,29 which degrade the charge carriermobility.Our DFT supercell calculations also reveal that sulfur doping

    increases the density of states close to the valence bandmaximum(Figures 4ac) and creates an additional peak in the 12.5 at. %doping range. This result suggests the possibility of resonantstates30 that can lead to increased , although this could also bedue to limitations of the supercell approximation. The valenceband spectra obtained by synchrotron XPS31 (see Figure 4d)agree with envelope proles of the discrete structure obtained bythe DFT calculations, but no resonant states are discernible (seeFigure 4d). The absence of an O 2p peak at 5 eV in the valenceband points to inhibited oxidation32 and absence of oxygendoping. The above results conrm sulfur doping is responsiblefor the observed increase, but the question of whether or notthere are resonant states remains open.The high ZT 0.95 at 420 K despite a comparatively low

    150 VK1 suggests the possibility of accessing even higher ZTthrough optimized doping. To estimate optimal doping, resultsof rst principles band structure calculations carried out using thelinearized augmented plane wave method (see SupportingInformation) were used to determine as a function of dopingand temperature based on Boltzmann transport theory. Ourcalculations indicate a substantial increase in at h < 1019 cm3

    for a wide temperature range of 300 T 800 K (Figures 5a,b)with maxima 200 VK1, without entering the bipolarregime at 300 K. The theoretical calculations agree well withexperimentally determined in the 300450 K range (Figure5a,b), conrming the applicability of our approach.We calculated ZT as a function of temperature and hole

    concentration from experimentally measured L, and as

    function of h, and theoretically determined , Lorenz numberand e using rst principles transport calculations. We nd 1.7 peak ZT 2 for 400 T 600 K for h < 3 1019 cm3 (Figure5b,c). For example, peak ZT 2 at 500 K for h 1.4 1019 cm3and decreases to 1.2 at 800 K due to bipolar conduction. Fornanobulk Sb2Te3 with higher h, the ZT peaks at highertemperatures, and for a factor of 10 increase in h the peak ZTshifts from 300 to 800 K. These results predict unprecedentedZT 2 without alloying if the antisite defects can be suppressedfurther by optimal sulfur doping. Furthermore, as mentionedearlier, the weak temperature-dependence of suggests that thepower factor 2 and ZT can be improved further viaoptimization of the grain size and nanostructure.In summary, we have demonstrated a novel approach for

    producing high ZT thermoelectrics by the use of subatomicpercent sulfur doping. Sulfur substitution of isovalent Te 6c sitessuppresses antisite defect formation. The resultant 10- to 100-fold decrease in the hole concentration increases by 1025%over the highest observed in nondoped counterparts. Subatomicsulfur doping also fosters high due to the diminished eect ofimpurity scattering of holes, while nanostructuring results inultralow . Cumulatively, we obtain 3-fold higher ZT0.95 thannon-nanostructured analogues at 423 K and the state-of-artalloys. Our experimental results and theoretical predictionspresage that further gains in , and hence ZT, are possible bymanipulating dopant type, level, and site occupancy. Ourapproach of doping-induced increase and nanostructuring-induced diminution, without compromising , opens up newopportunities for obtaining ZT > 2 and transforming thermo-electric refrigeration and energy-harvesting.

    ASSOCIATED CONTENT*S Supporting InformationExperimental methods and materials: nanoplate synthesis,nanobulk fabrication, materials characterization, thermoelectrictransport characterization, and theoretical modeling as well as aPisarenko and Weidemann Franz plots (Figure S1). Thismaterial is available free of charge via the Internet at http://pubs.acs.org.

    AUTHOR INFORMATIONCorresponding Author*E-mail: [email protected] authors declare no competing nancial interest.

    ACKNOWLEDGMENTSWe gratefully acknowledge funding from the U.S. Department ofEnergy through the S3TEC Energy Frontier Research Center(G.R. and T.B.T.), and partial support from National ScienceFoundation grant ECCS 1002282/301 (R.R. and G.R.), and theKaiteki Institute, Japan (G.R. and T.B.T.). We thank Dr. JoeWoicik, Dr. Barry Karlin, and Dr. Dan Fischer for help withsetting up the photoemission experiments carried out at theNational Synchrotron Light Source at Brookhaven NationalLaboratory, supported under the U.S. Department of EnergyContract DE-AC02-98CH10886.

    REFERENCES(1) Minnich, A. J.; Dresselhaus, M. S.; Ren, Z. F.; Chen, G. EnergyEnviron. Sci. 2009, 2, 466479.(2) Nolas, G. S.; Sharp, J.; Goldsmid, H. J. Thermoelectrics: Basicprinciples and new materials developments; Springer: New York, 2001.

    Nano Letters Letter

    dx.doi.org/10.1021/nl301639t | Nano Lett. 2012, 12, 452345294528

  • (3) Horak, J.; Koudelka, L.; Klikorka, J. React. Solids 1988, 5, 351359.(4) Horak, J.; Drasar, C.; Novotny, R.; Karamazov, S.; Lostak, P. Phys.Status Solidi A 1995, 149, 549556.(5) Drasar, C.; Lostak, P.; Uher, C. J. Electron. Mater. 2010, 39, 21622164.(6) Zhitinskaya, M. K.; Nemov, S. A.; Ivanova, L. D. Phys. Solid State2002, 44, 4247.(7) Rowe, D. M. Thermoelectrics Handbook: Macro to Nano; CRCPress: Boca Raton, FL, 2005.(8) Rowe, D. M. CRC Handbook of Thermoelectrics; CRC Press: BocaRaton, FL, 1995.(9) Stary, Z.; Horak, J.; Stordeur, M.; Stolzer, M. J. Phys. Chem. Solids1988, 49, 2934.(10) Lostak, P.; Karamazov, S.; Horak, J. Phys. Status Solidi A 1994,143, 271276.(11) Horak, J.; Stary, Z.; Lostak, P.; Pancir, J. J. Phys. Chem. Solids 1988,49, 191198.(12) Kanatzidis, M. G. Chem. Mater. 2010, 22, 648659.(13) Mehta, R. J.; Karthik, C.; Singh, B.; Teki, R.; Borca-Tasciuc, T.;Ramanath, G. ACS Nano 2010, 4, 50555060.(14) Mehta, R. J.; Zhang, Y. L.; Karthik, C.; Singh, B.; Siegel, R. W.;Borca-Tasciuc, T.; Ramanath, G. Nat. Mater. 2012, 11, 233240.(15) Pal, S. K.; Son, Y.; Borca-Tasciuc, T.; Borca-Tasciuc, D. A.; Kar, S.;Vajtai, R.; Ajayan, P. A. J. Mater. Res. 2008, 23, 20992105.(16) Zhao, Y. X.; Dyck, J. S.; Hernandez, B. M.; Burda, C. J. Am. Chem.Soc. 2010, 132, 49824983.(17) Scheele, M.; Oeschler, N.; Veremchuk, I.; Reinsberg, K. G.;Kreuziger, A. M.; Kornowski, A.; Broekaert, J.; Klinke, C.; Weller, H.ACS Nano 2010, 4, 42834291.(18) Pelz, U.; Kaspar, K.; Schmidt, S.; Dold, M.; Jagle, M.; Pfaadt, A.;Hillebrecht, H. J. Electron. Mater. 2012, 41, 18511857.(19) Dong, G. H.; Zhu, Y. J.; Chen, L. D.Cryst. Eng. Commun. 2011, 13,68116816.(20) Dong, G. H.; Zhu, Y. J.; Chen, L. D. J. Mater. Chem. 2010, 20,19761981.(21) Zhang, Y. L.; Mehta, R. J.; Belley, M.; Han, L.; Ramanath, G.;Borca-Tasciuc, T. Appl. Phys. Lett. 2012, 100, 193113193113-4.(22) Antimony telluride (Sb2Te3) Debye temperature, heat capacity.In Springer Materials - The Landolt-Bornstein Database; Madelung, O.,Rossler, U., Schulz, M., Eds.; Springer: New York, 1998; Vol. 41C; DOI:10.1007/10681727_1053.(23) Horak, J.; Lostak, P.; Benes, L. Philos. Mag. B 1984, 50, 665671.(24) Simon, G.; Eichler, W. Phys. Status Solidi B 1981, 107, 201206.(25) Zhang, Y. Thermal and thermoelectric transport in nano-structured materials from pnictogen chalcogenide nanoplate crystals.Ph.D. Dissertation, Rensselaer Polytechnic Institute, Troy, NY, 2012.(26) Chizhevskaya, S. N.; Shelimova, L. E. Inorg. Mater. 1995, 31,10831095.(27) Purkayastha, A.; Kim, S.; Gandhi, D. D.; Ganesan, P. G.; Borca-Tasciuc, T.; Ramanath, G. Adv. Mater. 2006, 18, 29582963.(28) Carroll, T. X.; De, J.; Maclaren, D. C.; Thomas, T. D.; Saethre, L. J.J. Electron Spectrosc. Relat. Phenom. 1987, 42, 281284.(29) Horak, J.; Cermak, K.; Koudelka, L. J. Phys. Chem. Solids 1986, 47,805809.(30) Heremans, J. P.; Jovovic, V.; Toberer, E. S.; Saramat, A.; Kurosaki,K.; Charoenphakdee, A.; Yamanaka, S.; Snyder, G. J. Science 2008, 321,554557.(31) Chasse, T.; Berg, U. Cryst. Res. Technol. 1985, 20, 14751479.(32) Fang, F.; Opila, R. L.; Venkatasubramanian, R.; Colpitts, T. J. Vac.Sci. Technol. A 2011, 2935.(33) Poudel, B.; Hao, Q.; Ma, Y.; Lan, Y. C.; Minnich, A.; Yu, B.; Yan,X.; Wang, D. Z.; Muto, A.; Vashaee, D.; Chen, X. Y.; Liu, J. M.;Dresselhaus, M. S.; Chen, G.; Ren, Z. Science 2008, 320, 634638.

    Nano Letters Letter

    dx.doi.org/10.1021/nl301639t | Nano Lett. 2012, 12, 452345294529