Colloidal Synthesis of Te-Doped Bi Nanoparticles: Low-Temperature

May 16, 2017 - School of Materials Science and Engineering, Ulsan National ... *E-mail: [email protected] (J.E.L.)., *E-mail: [email protected] (J.S...
0 downloads 0 Views 2MB Size
Subscriber access provided by NEW YORK UNIV

Article

Colloidal Synthesis of Te-doped Bi Nanoparticles: LowTemperature Charge Transport and Thermoelectric Properties Da Hwi Gu, Seungki Jo, Hyewon Jeong, Hyeong Woo Ban, Sung Hoon Park, Seung Hwae Heo, Fredrick Kim, Jeong In Jang, Ji Eun Lee, and Jae Sung Son ACS Appl. Mater. Interfaces, Just Accepted Manuscript • Publication Date (Web): 16 May 2017 Downloaded from http://pubs.acs.org on May 18, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

ACS Applied Materials & Interfaces is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 34

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Colloidal Synthesis of Te-doped Bi Nanoparticles: Low-Temperature Charge Transport and Thermoelectric Properties Da Hwi Gu,† Seungki Jo,† Hyewon Jeong,† Hyeong Woo Ban,† Sung Hoon Park,† Seung Hwae Heo,† Fredrick Kim,† Jeong In Jang,‡ Ji Eun Lee,*,‡ and Jae Sung Son*,† †

School of Materials Science and Engineering, Ulsan National Institute of Science and Technology (UNIST), Ulsan 44919, Republic of Korea,



Thermoelectric Conversion Research Center, Korea Electrotechnology Research Institute (KERI), Changwon 51543, Republic of Korea.

KEYWORDS: doped nanoparticles, bismuth, colloidal synthesis, charge carrier transport, thermoelectric properties

ACS Paragon Plus Environment

1

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 34

ABSTRACT

Electronically doped nanoparticles formed by incorporation of impurities have been of great interest because of their controllable electrical properties. However, the development of a strategy for n-type or p-type doping on sub-10 nm-sized nanoparticles under the quantum confinement regime is very challenging using conventional processes owing to the difficulty in synthesis. Herein, we report the colloidal chemical synthesis of sub-10 nm-sized Te-doped Bi nanoparticles with precisely controlled Te content from 0% to 5% and systematically investigate their low-temperature charge transport and thermoelectric properties. Microstructural characterization of nanoparticles demonstrates that Te ions are successfully incorporated into Bi nanoparticles rather than attaching on the nanoparticle surfaces. Low-temperature Hall measurement results of the hot-pressed Te-doped Bi nanostructured materials, of which grain sizes are ranging from 30 nm to 60 nm, show that the charge transport properties are governed by the doping content, and the related impurity and nanoscale grain boundary scattering. Furthermore, the low-temperature thermoelectric properties reveal that the electrical conductivity and Seebeck coefficient expectedly change with the Te content, while the thermal conductivity is significantly reduced by Te doping because of phonon scattering at the sites arising from impurities and nanoscale grain boundaries. Accordingly, the 1% Te-doped Bi sample exhibit higher ZT by ~10% than that of the undoped sample. The synthetic strategy demonstrated in this study offers the possibility of electronic doping of various quantum confined nanoparticles for diverse applications.

ACS Paragon Plus Environment

2

Page 3 of 34

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

1. Introduction Doping, the intentional introduction of impurities into a material, lies at the heart of materials science because it provides unique tools to control the electronic, optical, and magnetic properties of materials. For example, an impurity with one more or one less valence electron can donate one more electron or hole carrier into a host material, which then has available charge carriers to carry current. With the recent advances in nanomaterials,1−3 doping of lowdimensional nanomaterials has also attracted tremendous attention. Since the potential applicability of nanoscale electronic materials such as semiconductors and semimetals under the quantum confinement regime ultimately depends on tailoring the electronic properties by doping as well as tuning sizes, the development of a nanoscale doping methodology is crucial for advanced electronic and energy applications.4-8 Bismuth has been recognized as an important electronic material because of the extremely small effective mass of electron (me*, me*=0.001m0), long electron mean free path at the L-point, and high electron mobility (35000 cm2/V·s).9−11 Furthermore, its relatively low thermal conductivity, arising from the heavy mass of a Bi atom that efficiently scatters phonons, makes it a great candidate for low-temperature n-type thermoelectric materials.12 The thermoelectric efficiency of materials can be evaluated by the figure-of-merit, ZT = σS2T/κ, where σ is the electrical conductivity, S is the Seebeck coefficient, κ is the thermal conductivity, and T is the absolute temperature.13 Recently, it has been reported that the semimetal-to-semiconductor transition can occur for low-dimensional Bi nanostructures such as quantum wells or quantum wires arising from the quantum confinement effect.14 Since hole carriers in semimetals like Bi contribute to compensate the Seebeck coefficient for electrons, a great enhancement in ZT has been predicted in Bi quantum nanostructures under the quantum confinement regime because of

ACS Paragon Plus Environment

3

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 34

this transition. For example, Dresselhaus et al. predicted that the ZT values of Te-doped Bi quantum wells or quantum wires can increase up to more than 13 by the increase in the density of states near the Fermi level.11,14−16 In addition to dimensionality, another important parameter for highly efficient n-type Bi thermoelectric materials is the electron concentration which can be controlled by doping because the electrical conductivity and the Seebeck coefficient are inversely proportional to the carrier concentration; therefore, the optimum doping content needs to be determined for the fabrication of highly efficient thermoelectric materials.17−19 However, despite the importance of doping on Bi nanostructures, studies on the properties of sub-10 nmsized doped Bi nanostructures under the quantum confinement regime have rarely been reported because of difficulties in synthesis. Moreover, most studies have focused on the thermoelectric properties of Bi nanostructures with grain sizes larger than hundreds of nanometer20,21 or Bi nanowires with diameter of > 40 nm,10,11,22 which restrict the realization of theoretically predicted highly efficient low-temperature thermoelectric materials. Colloidal chemistry routes are widely utilized for the synthesis of various nanoparticles such as metals, semiconductors, and ceramics because of the ease in controlling their sizes and shapes, scale-up, and the ability for surface functionalization.1,23−25 Although there have been recent advances in the synthesis of doped nanoparticles, most reports have focused on the introduction of magnetic ions such as Mn and Co into semiconductor quantum dots, while n-type or p-type electronic doping by the introduction of impurities has rarely been studied.4−8,26 In fact, nanoparticles with n-type or p-type dopants undergo self-purification and remove the dopants out of atomic lattices or exhibit electrochemical instability to trap the charge carriers at the surfaces.4,27 The recently suggested concept of remote doping, where extra electrons are injected into undoped nanoparticles by molecular attachment on nanoparticles or by an electrochemical

ACS Paragon Plus Environment

4

Page 5 of 34

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

route, can provide an alternative to the electronic doping of nanoparticles.6,7,28,29 However, the doping content in remotely doped nanoparticles is not maintained sustainably and is sensitive to environmental changes.6,7 As an alternative route, Geyer et al. proposed the chemical deposition of Cd layer on the surface of InAs nanocrystal to achieve the doping.30 Here, we report the development of a synthesis process for sub-10 nm-sized Te-doped Bi nanoparticles, where the Te content was precisely tuned from 0% to 5%. Te ion is a well-known n-type dopant that can donate one electron in a Bi crystal.17,18 We demonstrated that Te ions were successfully incorporated into Bi nanoparticles, rather than remaining on the surfaces. These Te-doped Bi nanoparticles can provide a great model system for studying the lowtemperature charge transport and thermoelectric properties of low-dimensional nanostructures with controlled doping content. Low-temperature Hall measurements on the hot-pressed Tedoped Bi nanostructured materials revealed that the charge transport properties were strongly dependent on the doping content and the related impurity and grain boundary scattering, and the electrical conductivity and Seebeck coefficient predictably changed with the doping content. Furthermore, the thermal conductivity of doped samples decreased significantly compared to that of the undoped samples because of the impurity and grain boundary scattering of phonons, which led to an increase of ZT in the Te-doped Bi nanoparticles by ~10%. This promising strategy for electronic doping in nanostructured electronic materials will provide an additional degree of control over charge transport and thermoelectric properties in addition to dimensionality.

ACS Paragon Plus Environment

5

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 34

2. Experimental Procedures Materials. Bismuth neodecanoate (technical grade), 1-dodecanethiol (DDT, ≥ 98%), 1octadecene (90%), tri-n-octylphosphine (TOP, 90%), and ethanethiol (97%) were purchased from Aldrich Chemical Co. Te powder (99.999%) was purchased from 5N Plus. All the materials were used without further purification. Synthesis of Te-doped Bi nanoparticles. All synthetic processes for Te-doped Bi nanoparticles were based on conventional air-free techniques using a Schlenk line and a nitrogen-filled glovebox. Typically, 10 mmol of bismuth neodecanoate was dissolved in 50 ml of 1-octadecene and was heated to 393 K under vacuum for 2 h. After that, the reaction mixture was cooled to 353 K. At this temperature, 2.4 ml of 1-dodecanethiol was added and the reaction was continued for 5 min. After the addition of 1-dodecanethiol, the colorless solution turned yellow, indicating the formation of a bismuth dodecanethiolate complex.31 For the synthesis of the Te precursor solution, the desired amount of the Te powder, depending on the doping concentration, was dissolved into 10 ml of TOP. After vigorous stirring for 12 h, the color of this solution turned yellow. This Te solution in TOP was injected into the reaction mixture at 353 K and it was immediately cooled down to 338 K. The yellow-colored reaction mixture turned black immediately after the injection of the Te solution in TOP, which indicated the formation of Bi nanoparticles. After aging it for 40 min at 338 K, the reaction mixture was centrifuged with excess acetone. The precipitate was redispersed in 10 ml hexane, followed by centrifugation with excess acetone. Typically, we obtained 1.5 g of powder-like Te-doped Bi nanoparticles in a single batch reaction. (Figure S1) Preparation of pellets. For measuring the electrical properties of Te-doped Bi nanoparticles, we fabricated disk-shaped pellets by hot-pressing. Typically, a dodecanethiol-capped Te-doped Bi

ACS Paragon Plus Environment

6

Page 7 of 34

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

nanoparticle solution in hexane was mixed with ethanethiol (20% volume fraction of nanoparticle solution), followed by vigorous stirring for 1 h. After ligand exchange, the nanoparticles were precipitated at the bottom of the flask because of the agglomeration due to the shorter inter-particle distance. Approximately 0.8 g of dried nanoparticle powder was loaded into a graphite mold (12.7 mm diameter) and was pressed at 393 K under a pressure of 600 psi in an N2-filled glovebox with holding time of 10 min, producing a 1 mm-thick disk-shaped pellet. Materials characterization. The microstructural characterization of the Te-doped Bi nanoparticles was conducted using transmission electron microscopy (TEM, JEM-2100, JEOL) operated at 200 kV. Elemental mapping images of Te-doped Bi nanoparticles were obtained with energy dispersive X-ray spectroscopy (EDS) using a JEM-2100F, JEOL. Hot-pressed pellets were characterized using field emission scanning electron microscopy (Nano-FESEM, FEI Nova-NanoSEM230, FEI) operated at 10kV. Elemental mapping images of hot-pressed pellets were obtained using EDS (analysis by using a Nova-NanoSEM230, FEI. X-ray diffraction (XRD) patterns for Te-doped Bi nanoparticles were obtained using a high power XRD (HPXRD, D/MAX2500V/PC, Rigaku) with a Cu Kα X-ray source, which has a distinctive wavelength of 1.5418 Å, operating at 40 kV and 200mA. To estimate the grain sizes of the hot-pressed Tedoped Bi nanostructured materials, the Scherrer equation was used with the equipment full-width at half maximum of 0.087 o in 2θ, obtained using a Si bulk crystal reference. The molar ratio of Te in Te-doped Bi nanoparticles was characterized using inductively coupled plasma optical emission spectrometry (ICP-OES, 700-ES, Varian). The FT-IR spectrum was obtained using FTIR spectrometer (Varian 670). Low-temperature Hall effect and thermoelectric properties measurements. Carrier transport properties in the temperature range from 50K to 300K were measured by hall measurement using

ACS Paragon Plus Environment

7

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 34

physical property measurement system (PPMS, PPMS-EverCool-II, Quantum Design) in magnetic fields of ±1, 3 and 5 T and the averaged values are displayed. Thermoelectric properties in the temperature range from 50 K to 300 K were measured by physical property measurement system (PPMS) with Thermal Transport Option.

3. Results and discussion The synthesis of Te-doped Bi nanoparticles was conducted by modifying a method used for the synthesis of Bi nanoparticles reported by Son et al.12 In this reaction, a Bi-thiolate complex was reduced to form Bi nanoparticles with the mild reducing agent TOP. In addition, TOP can act as an efficient solvent for dissolving Te and form a TOP-Te complex, which is a widely used Te precursor for the synthesis of metal telluride nanoparticles.18,32 Accordingly, we utilized the TOP-Te solution as the Te precursor as well as the reducing agent for the synthesis of Te-doped Bi nanoparticles. The Te content in TOP-Te solution was controllably varied from 0% to 5% in molar ratio to Bi precursor. The Te doping content in the synthesized Bi nanoparticles, confirmed by the ICP-OES analyses (Table 1), was slightly higher than the initial content of the Te precursor. This can be attributed to the difference in the reactivity of the Bi and Te precursors. To confirm the Te doping inside the lattice of the Bi nanoparticles, we performed the surface treatment of the 5% Te-doped Bi nanoparticles with excess TOP, which effectively binds to the Te ions placed on the surface of Bi nanoparticles. The treated and purified Te-doped Bi nanoparticles exhibited the same Te content as the untreated nanoparticles, which demonstrated the doping of Te ions into atomic lattices in the Bi nanoparticles (Table 2).

ACS Paragon Plus Environment

8

Page 9 of 34

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Table 1. Comparison of Te precursor concentration for the synthesis of Te-doped Bi nanoparticles and the measured Te doping concentration by ICP-OES analysis. Te molar ratio Te precursor concentration (mol%)

Te concentration obtained by ICP-OES analysis (mol%)

1%

1.77%

3%

4.25%

5%

6.05%

ACS Paragon Plus Environment

9

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 34

Table 2. Te molar ratio to Bi nanoparticles before and after TOP treatment

Te molar ratio Before TOP treatment

After TOP treatment

6.05%

5.92%

ACS Paragon Plus Environment

10

Page 11 of 34

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

The structural characteristics of Te-doped Bi nanoparticles were investigated using TEM and XRD analysis. The TEM image of undoped Bi nanoparticles (Figure 1a) shows uniform-sized spherical nanostructures with an average size of 11 nm, which agrees with the result reported in a previous study by Son et al.12 Interestingly, we found larger nanostructures with a size of 10-30 nm mixed with the spherical 11 nm-sized structures, as shown in the TEM images of the Tedoped Bi nanoparticles (Figure 1b-d). The concentration of the former structures increased with increasing Te content, which indicates that there is a relationship between the large nanostructures and the Te doping content. To identify these structures, we conducted elemental mapping analysis on two separate regions for the 11 nm-sized nanoparticles and larger nanoparticles in 3% Te-doped Bi nanoparticles using EDS (Figure S2a and S2b). The elemental mapping image of 11 nm-sized nanoparticles indicates the presence of Bi with its content approaching 99%, whereas a significant Te content was observed in the larger nanostructures. Elemental mapping sum spectra of these two nanostructures exhibited 0.4 mol% of Te in the smaller one and 4.9 mol% in larger one (Table S1). These results indicate that the larger nanostructures are more heavily doped, which were mixed with the less doped Bi nanoparticles. The XRD patterns (Figure 1e) of all the nanoparticles, regardless of the Te content, correspond to that of bulk rhombohedral Bi crystals (JCPDS 85-1331), while the peaks become slightly sharper with increasing Te content because of an increase in the concentration of larger nanoparticles. In addition, the peaks related to secondary phases such as Bi-rich BiTe or Te were not observed. This indicates that Te ions are atomically doped in the Bi nanoparticles, rather than forming, secondary structures. More importantly, the diffraction peaks in the XRD patterns (Figure 1f and 1g) progressively shifted to higher 2θ angles with increasing Te content, demonstrating the integration of Te ions inside an atomic lattice in the Bi nanoparticles.

ACS Paragon Plus Environment

11

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 34

Figure 1. TEM images of Te-doped Bi nanoparticles with various Te concentrations of (a) 0%, (b) 1%, (c) 3%, and (d) 5%. (e) Full-scale XRD patterns and (f) enlarged patterns of Te-doped Bi nanoparticles with various Te content. (g) Plot for the (012) peak shift of Te-doped Bi nanoparticles in the XRD patterns.

ACS Paragon Plus Environment

12

Page 13 of 34

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

To study the charge transport and thermoelectric behaviors of Te-doped Bi nanoparticles, we prepared disk-shaped pellets by pressing nanoparticle powders at 393 K under an inert atmosphere (Figure S3). The hot-pressing under even mild condition (T~393 K) was found to be advantageous to significantly improve the electrical properties of Te-doped Bi nanostructured materials. As shown in Figure S4, the electrical conductivity of the hot-pressed samples is an order of magnitude higher than those obtained from the cold-pressed samples, while their Seebeck coefficients give similar values. To further improve the electrical properties, long-hydrocarbon-chain dodecanethiol ligand capped Te-doped Bi nanoparticles were exchanged with short-chain ethanethiol before the hotpressing process. These short-chain alkylthiol ligands allow for efficient electronic communication as well as effective contact among the Bi nanograins, significantly improving the electrical properties. Son et al. reported that the electrical conductivity of the pressed Bi nanoparticles increased upon decreasing the chain-length of the alkylthiol capping ligands.12 Furthermore, the organic ethanethiol to the detriment of electrical properties can be further evaporated during drying process under vacuum due to its low boiling temperature (308 K) close to room temperature. The Fourier-transform infrared absorption (FT-IR) spectrum of both of dried and heat-treated Te-doped Bi nanoparticles at 393 K (Figure S5) shows no peaks related to the bands corresponding to S-H and C-H stretch modes of ethanethiol The structural characteristics of the prepared samples were characterized by SEM and XRD. The SEM images of the hot-pressed Te-doped Bi nanostructured materials (Figure 2a-d) show that the size of the grains ranges from several tens to several hundreds of nanometers, which is indicative of the grain growth of Te-doped Bi nanoparticles during the pressing process. Interestingly, the grain sizes apparently decrease with increasing Te content. Since the impurity

ACS Paragon Plus Environment

13

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 34

ions can sometimes either promote or disturb the grain growth by adjusting the atomic diffusion rate, this phenomenon can be attributed to the effect of Te ions on the diffusion of Bi atoms.33 The XRD patterns of all the samples (Figure 2e) were indexed to Bi rhombohedral structures, while the diffraction peaks broaden with increasing Te content. The grain sizes of the Te-doped Bi pellets, evaluated from the (012) peaks in the XRD patterns of the samples using the Scherrer equation (Figure 2f), ranged from 30 to 60 nm, and decreased with increasing Te content, which was in agreement with the results obtained by the SEM analysis. In addition, peaks related to the impurity or the secondary phase were not observed. This suggests that Te ions were integrated well into the Bi lattices, rather than forming a secondary phase. To further investigate the homogeneous doping of Te, the hot-pressed Te-doped Bi nanostructured materials were characterized by SEM-EDS. EDS mapping results (Figure S6a-d and Table S2) show that Te is detected on the entire surface of the specimen but the strong signals coming from a Te phase is not observed. Weak signals of Te on the undoped sample and the slightly higher Te content in the doped samples are attributed to artifacts, considering the composition of the materials. These results demonstrate homogeneous doping of Te ions in Bi nanostructured materials.

ACS Paragon Plus Environment

14

Page 15 of 34

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Figure 2. SEM images of Te-doped Bi pellets prepared from Te-doped Bi nanoparticles with various Te content of (a) 0%, (b) 1%, (c) 3%, and (d) 5%. (e) XRD patterns of Te-doped Bi pellets prepared from Te-doped Bi nanoparticles with various Te content of 0%, 1%, 3%, and 5%. (f) Mean grain sizes of the pellets estimated from the (012) peak in the XRD patterns using the Scherrer equation.

ACS Paragon Plus Environment

15

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 34

To study the charge transport properties of the hot-pressed Te-doped Bi nanostructured materials at low temperatures, we measured the carrier concentration (n) and mobility (µ) by Hall measurement, since doping of impurities to provide extra charge carriers in semiconductors or semimetals critically adjusts parameters that depend on temperature. The Hall coefficient (RH) (Figure S7) exhibited negative values over the entire temperature range from 50 to 300 K, indicating the n-type electrical behavior of the hot-pressed Te-doped Bi nanostructured materials. The electron concentration of an undoped Bi sample was 3.3 × 1019 cm-3 at room temperature and swiftly decreased with decreasing temperature (Figure 3a). On the other hand, Te-doped Bi samples exhibited almost one order of magnitude higher electron concentrations of 1.4 × 1020 cm-3 for 1%, 1.7 × 1020 cm-3 for 3%, and 1.9 × 1020 cm-3 for 5% Te-doped samples. In addition, the temperature dependence of the electron concentrations of Te-doped samples was much weaker than that of the undoped sample. This phenomenon can be attributed to the occupancy of thermally activated charge carriers in the total carrier concentration. The doped samples should have much more extra electrons generated from the shallow doping states near the conduction band edge, which should not be strongly temperature dependent. On the other hand, the number of electrons in the undoped sample can be governed by thermal activation of electrons across the entire bandgap. These results show the great possibility of precise control of the carrier concentration of nanostructured materials by chemical incorporation of dopants inside nanoparticles. The undoped Bi nanostructured material exhibited a mobility as high as 610 cm2/V·s at room temperature. (Figure 3b) Furthermore, it dramatically increased with decreasing temperature and reached 1400 cm2/V·s at 50 K. On the other hand, the electron mobilities of Te-doped Bi nanostructured materials were much lower than those of the undoped samples, between 120 and

ACS Paragon Plus Environment

16

Page 17 of 34

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

240 cm2/V·s. As Te doping content increased, the mobility decreased over the entire temperature range and showed a much weaker temperature dependence. These results suggest that the charge transport properties of Te-doped Bi samples are primarily dominated by the Te doping content. The effects of Te doping on the temperature-dependent electrical behaviors can be summarized by two factors in the microstructures: 1) the increase in the impurity concentration arising from Te doping in Bi nanostructured materials, and 2) as aforementioned, the decrease in the grain sizes of Te-doped samples with increasing Te doping content. These complex microstructural characteristics affecting the charge transport properties can be explained by the carrier scattering mechanism. According to Matthiessen’s rule, the contribution of carrier scattering factors to the total mobility can be qualitatively estimated by the following equation.34,35 µT-1 = µI-1 + µL-1 + µD-1 + … , where µT is the total carrier mobility, µI is the impurity scattering factor, µL is the lattice scattering factor, and µD is the defect including grain boundary or alloy scattering factor. In this rule, the total scattering is the sum of each contribution from different electron scattering processes. Typically, the electron-phonon (lattice) scattering, which is generally observed in intrinsic semiconductors, strongly depends on the temperatures. However, the impurity and defect scattering processes are almost temperature independent.34−37 In the current study, the mobility of the undoped sample shows a strong dependence on temperature, which indicates that electron-phonon scattering is the dominant scattering process in its charge transport, being further supported by the low electron concentration. On the other hand, the Te-doped samples show an almost constant mobility regardless of the temperature. Since the Te-doped samples have much higher impurity concentrations and smaller nanoscale grain sizes, their temperaturedependent charge transport can be understood by impurity and grain boundary scattering of

ACS Paragon Plus Environment

17

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 34

charge carriers. The much lower mobility of Te-doped samples than the undoped sample should be attributed to these scattering processes.

ACS Paragon Plus Environment

18

Page 19 of 34

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Figure 3. Temperature dependences of (a) electron concentration (b) mobility of Te-doped Bi pellets with various Te content of 0% (black square), 1% (red circle), 3% (blue upward triangle), and 5% (green downward triangle).

ACS Paragon Plus Environment

19

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 34

In the current study, the Te doping content in Bi nanoparticles was precisely controlled by a chemical approach and it also provides an additional degree of control over the grain sizes at the nanoscale. This promising chemical approach may provide an effective route for realizing ideal Bi nanostructured thermoelectric materials with optimized doping content and grain sizes. In this regard, we measured the thermoelectric properties of the hot-pressed Te-doped Bi nanostructured materials with 0%, 1%, 3%, and 5% Te doping contents at temperatures ranging from 50 K to 300 K using PPMS. The electrical conductivity of the undoped sample was 3.2 × 105 S/m at room temperature and decreased with decreasing temperature. (Figure 4a) This positive temperature dependence can be observed in the electrical properties of an intrinsic semiconductor. This result was in agreement with the temperature dependence of electron concentration since the electrical conductivity (σ) of the materials can be determined by the equation σ = neµ for unipolar transport, where e is the electron charge. On the other hand, 1% and 3% Te-doped samples exhibited the highest electrical conductivity of 5.1 × 105 S/m at room temperature, which was around two times higher than that of the undoped sample and comparable to the value of bulk Bi (6.7 × 105 S/m).35 In addition, all Te-doped samples showed almost constant or negative temperature dependences. This phenomenon can be understood by considering that metals or degenerate semiconductors exhibit scattering dominant electrical properties, rather than the number of charge carriers generated by thermal energy. The Seebeck coefficients for all samples have negative values, indicating n-type characteristics. (Figure 4b) As predicted by the relationship that the Seebeck coefficient is inversely proportional to the carrier concentration, the samples show a decrease from -56.2 µV/K for the undoped sample to -29.2 µV/K for the 5% Te-doped sample with increasing Te doping

ACS Paragon Plus Environment

20

Page 21 of 34

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

content. In addition, considering the scattering probability of charge carriers at the sites of ionized impurity is reciprocally proportional to the energy of carriers, the low energy charge carriers can be more scattered than the high energy ones.38 Since the low energy charge carriers are known to contribute the Seebeck coefficient negatively, the ionized impurity scattering can enhance the Seebeck coefficient by the energy filtering effect. In the current samples, although the Seebeck coefficients were decreased with increasing the content of Te doping, the intensified ionized impurity scattering might somewhat compensate the reduction of the Seebeck coefficient. The power factors of all the samples increased with increasing temperatures because of the temperature dependence of the Seebeck coefficient (Figure 4c). The 1% Te-doped sample exhibits the highest value of 13.2 µW/cm·K2 at room temperature because of the highest electrical conductivity and moderately high Seebeck coefficient.

ACS Paragon Plus Environment

21

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 34

Figure 4. Temperature dependence of (a) electrical conductivity, (b) Seebeck coefficient, and (c) power factor of Te-doped Bi pellets with various Te content of 0% (black square), 1% (red circle), 3% (blue upward triangle), and 5% (green downward triangle).

ACS Paragon Plus Environment

22

Page 23 of 34

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

The thermal conductivity (κ) of the undoped sample was 7.32 W/m·K at room temperature and increases with decreasing temperatures (Figure 5a). On the other hand, the 1% Te-doped sample exhibits the highest thermal conductivity of 8.88 W/m·K at room temperature and decreased with decreasing temperatures. The 3% and 5% Te-doped samples show thermal conductivities of 7.37 W/m·K and 5.23 W/m·K, respectively, and show positive temperature dependences. The lattice thermal conductivity (κL) of the samples was evaluated based on the equation κ = κE + κL, where κE is the electronic contribution to total thermal conductivity, calculated by the WiedemannFranz law κE/σ = LT (Figure 5b). We used the Lorenz number (L) 2.4 × 10-8 V2/K2 for degenerate semiconductors.13,39The lattice thermal conductivity of undoped sample is 5.03 W/m·K at room temperature and increases with decreasing temperatures, whereas the thermal conductivities of the Te-doped samples are lower than that of the undoped sample over the entire temperature range and exhibit a weaker temperature dependence. The lattice thermal conductivity decreases with increasing Te doping content. In general, the phonon scattering for determining thermal conductivity can be understood using several mechanisms such as the Umklapp phonon-phonon scattering, phonon-impurity scattering, and phonon-boundary scattering.34−36 The decreased thermal conductivity with increasing Te doping content suggests that the phonon transport properties of these materials are dominated by Te impurity scattering and grain boundary scattering, which is independent of temperature. On the other hand, the Umklapp scattering is strongly dependent on temperature, which is responsible for the strong temperature dependence of the thermal conductivity of the undoped sample. Furthermore, the decrease in the thermal conductivity of the Te-doped samples can be attributed to the higher concentration of impurities and smaller grains available for phonon scattering. These results demonstrate that the Te doping on Bi nanoparticles can be

ACS Paragon Plus Environment

23

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 34

advantageous for reducing the thermal conductivity by phonon scattering at impurities and grain boundaries. The temperature-dependent ZT values of all the samples were calculated from the measured electrical conductivities, Seebeck coefficients, and thermal conductivity (Figure 5c). All the samples exhibit the highest ZT values at room temperature and decrease with decreasing temperatures. The highest value of 0.045 was achieved by the 1% Te-doped sample, which was ~10% higher than that of the undoped sample. In addition, the 1% doped Te sample exhibited higher ZT values than the undoped sample over a wide range of temperatures from 170-300 K. Although the absolute value of ZT in the current study is not so high as to compete with those of traditional bulk Bi (ZT=0.16, 300K) and recently developed nanostructured materials,15,40−43 the currently developed promising chemical route will offer an additional degree of control over the doping content in nanostructured thermoelectric materials with more optimized and enhanced efficiency.

ACS Paragon Plus Environment

24

Page 25 of 34

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Figure 5. Temperature dependence of (a) thermal conductivity, (b) lattice thermal conductivity, and (c) figure of merit ZT of Te-doped Bi pellets with various Te contents of 0% (black square), 1% (red circle), 3% (blue upward triangle), and 5% (green downward triangle).

ACS Paragon Plus Environment

25

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 34

4. Conclusion In summary, we successfully developed a method for the synthesis of Te-doped Bi nanoparticles and systematically investigated the charge transport and thermoelectric properties of the hot-pressed samples at low temperatures. The structural characterization of the synthesized Te-doped Bi nanoparticles clearly demonstrated the doping of Te ions inside the Bi atomic lattices rather than their attachment on their surfaces. The homogeneously doped Te ions in the hot-pressed samples were confirmed by elemental analysis. In addition, we found that the Te ion impurity significantly disturbed the grain growth of Bi, so that the grain sizes were controllable on the nanoscale by the Te doping content. The low-temperature charge transport properties of the hot-pressed Te-doped Bi nanostructured materials were somewhat different from those of the undoped sample, which exhibited typical semiconducting properties, and the developed doped materials were arguably heavily doped degenerate semiconductors. We found that the properties of the doped samples were governed by the impurity and grain boundary scattering of charge carriers, as demonstrated by the weak temperature dependences of the carrier mobility and concentration. These interesting properties of Te-doped Bi nanostructured materials were directly reflected in the thermoelectric properties, where the electrical conductivity increased and the Seebeck coefficient decreased with increasing Te content. The thermal conductivity decreased with increasing Te content because of phonon scattering at the sites of impurities and grain boundaries. The 1% Te-doped sample exhibited a higher ZT value than the undoped sample over a wide temperature range of 170-300 K. These low-temperature charge transport and thermoelectric properties of Te-doped Bi nanostructures will extend our understanding of the fundamental charge transport and thermoelectric properties in nanostructured electronic materials. Furthermore, the currently developed synthetic route for Te-doped Bi nanoparticles

ACS Paragon Plus Environment

26

Page 27 of 34

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

shows a new method for the precise control of extra charge carriers in nanostructures without performing electrochemical reactions at the surfaces. We believe that this methodology will be widely applicable for the production of doped semiconductor nanoparticles and nanostructured thermoelectric materials.

■ ASSOCIATED CONTENT Supporting Information The supporting information is available free of charge via the Internet at http://pubs.acs.org. More characterization results using EDS, Hall coefficient, ICP-OES, and photographs to show nanoparticle powder and a hot-pressed sample.

■ AUTHOR INFORMATION

Corresponding Author E-mail: [email protected] (J. S. Son), [email protected] (J. E. Lee) Notes The authors declare no competing financial interest.

■ ACKNOWLEDGMENT We acknowledge the Center for Advanced Meta-Materials (CAMM) funded by the Ministry of Science, ICT and Future Planning as Global Frontier Project (NRF-2014M3A6B3063704) in

ACS Paragon Plus Environment

27

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 28 of 34

Republic of Korea, and Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Education (2015R1C1A1A01053599 (J.S. Son), 2014R1A1A3053206 (J.E. LEE) ).

■ REFERENCES (1) Alivisatos, A. P. Semiconductor Clusters, Nanocrystals, and Quantum Dots. Science 1996, 271, 933-937. (2) Takagahara, T.; Takeda, K. Theory of the Quantum Confinement Effect on Excitons in Quantum Dots of Indirect-Gap Materials. Phys. Rev. B 1992, 46, 15578-15581. (3) Xia, Y.; Yang, P.; Sun, Y.; Wu, Y.; Mayers, B.; Gates, B.; Yin, Y.; Kim, F.; Yan, H. OneDimensional Nanostructures: Synthesis, Characterization, and Applications. Adv. Mater. 2003, 15, 353-389. (4) Norris, D. J.; Efros, A. L.; Erwin, S. C. Doped Nanocrystals. Science 2008, 319, 17761779. (5) Yu, J. H.; Liu, X.; Kweon, K. E.; Joo, J.; Park, J.; Ko, K.-T.; Lee, D. W.; Shen, S.; Tivakornsasithorn, K.; Son, J. S.; Park, J.-H.; Kim, Y.-W.; Hwang, G. S.; Dobrowolska, M.; Furdyna, J. K.; Hyeon, T. Giant Zeeman Splitting in Nucleation-Controlled Doped CdSe:Mn2+ Quantum Nanoribbons. Nat. Mater. 2010, 9, 47-53. (6) Shim, M.; Guyot-Sionnest, P. n-Type Colloidal Semiconductor Nanocrystals. Nature 2000, 407, 981-983.

ACS Paragon Plus Environment

28

Page 29 of 34

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

(7) Yu, D.; Wang, C.; Guyot-Sionnest, P. n-Type Conducting CdSe Nanocrystal Solids. Science 2003, 300, 1277-1280. (8) Schimpf, A. M.; Knowles, K. E.; Carroll, G. M.; & Gamelin, D. R. Electronic Doping and Redox-Potential Tuning in Colloidal Semiconductor Nanocrystals. Acc. Chem. Res. 2015, 48, 1929-1937. (9) Heremans, J. P.; Thrush, C. M.; Morelli, D. T.; Wu, M.-C. Thermoelectric Power of Bismuth Nanocomposites. Phys. Rev. Lett. 2002, 88, 216801. (10) Heremans, J.; Thrush, C. M. Thermoelectric Power of Bismuth Nanowires. Phys. Rev. B 1999, 59, 12579-12583. (11) Zhang, Z.; Sun, X.; Dresselhaus, M. S.; Ying, J. Y.; Heremans, J. Electronic Transport Properties of Single-Crystal Bismuth Nanowire Arrays. Phys. Rev. B 2000, 61, 4850-4861. (12) Son, J. S.; Park, K.; Han, M.-K.; Kang, C.; Park, S.-G.; Kim, J.-H.; Kim, W.; Kim, S.-J.; Hyeon, T. Large-Scale Synthesis and Characterization of the Size-Dependent Thermoelectric Properties of Uniformly Sized Bismuth Nanocrystals. Angew. Chem. 2011, 123, 1399-1402. (13) Synder, G. J.; Toberer, E. S. Complex Thermoelectric Materials. Nat. Mater. 2008, 7, 105-114. (14) Hicks, L. D.; Harman, T. C.; Dresselhaus, M. S. Use of Quantum-Well Superlattices to Obtain a High Figure of Merit from Nonconventional Thermoelectric Materials. Appl. Phys. Lett. 1993, 63, 3230-3232.

ACS Paragon Plus Environment

29

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 30 of 34

(15) Dresselhaus, M. S.; Chen, G.; Tang, M. Y.; Yang, R.; Lee, H.; Wang, D.; Ren, Z.; Fleurial, J.-P.; Gogna, P. New Directions for Low-Dimensional Thermoelectric Materials. Adv. Mater. 2007, 19, 1043-1053. (16) Hicks, L. D.; Dresselhaus, M. S. Effect of Quantum-Well Structures on the Thermoelectric Figure of Merit. Phys. Rev. B 1993, 47, 12727-12731. (17) Sumithra, S.; Takas, N. J.; Misra, D. K.; Nolting, W. M.; Poudeu, P. F. P.; Stokes, K. L. Enhancement in Thermoelectric Figure of Merit in Nanostructured Bi2Te3 with Semimetal Nanoinclusions. Adv. Energy Mater. 2011, 1, 1141-1147. (18) Son, J. S.; Choi, M. K.; Han, M.-K.; Park, K.; Kim, J.-Y.; Lim, S. J.; Oh, M.; Kuk, Y.; Park, C.; Kim, S.-J.; Hyeon, T. n-Type Nanostructured Thermoelectric Materials Prepared from Chemically Synthesized Ultrathin Bi2Te3 Nanoplates. Nano Lett. 2012, 12, 640-647. (19) Poudel, B.; Hao, Q.; Ma, Y.; Lan, Y.; Minnich, A.; Yu, B.; Yan, X.; Wang, D.; Muto, A.; Vashaee, D.; Chen, X.; Liu, J.; Dresselhaus, M. S.; Chen, G.; Ren, Z. High-Thermoelectric Performance of Nanostructured Bismuth Antimony Telluride Bulk Alloys. Science 2008, 320, 634-638. (20) Puneet, P.; Podila, R.; Zhu, S.; Skove, M. J.; Tritt, T. M.; He, J.; Rao, A. M. Enhancement of Thermoelectric Performance of Ball-Milled Bismuth due to Spark-Plasma-Sintering-Induced Interface Modifications. Adv. Mater. 2013, 25, 1033-1037. (21) Zhu, T.; Liu, Y.; Fu, C.; Heremans, J. P.; Synder, J. G.; Zhao, X. Compromise and Synergy in High-Efficiency Thermoelectric Materials. Adv. Mater. 2017, 1605884.

ACS Paragon Plus Environment

30

Page 31 of 34

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

(22) Kim, J.; Shim, W.; Lee, W. Bismuth Nanowire Thermoelectrics. J. Mater. Chem. C 2015, 3, 11999-12013. (23) Talapin, D. V.; Lee, J.-S.; Kovalenko, M. V.; Shevchenko, E. V. Prospects of Colloidal Nanocrystals for Electronic and Optoelectronic Applications. Chem. Rev. 2010, 110, 389-458. (24) Yin, Y.; Alivisatos, A. P. Colloidal Nanocrystal Synthesis and the Organic-Inorganic Interface. Nature 2005, 437, 664-670. (25) Kovalenko, M. V.; Scheele, M.; Talapin, D. V. Collodial Nanocrystals with Molecular Metal Chalcogenide Surface Ligands. Science 2009, 324, 1417-1420. (26) Archer, P. I.; Santangelo, S. A.; Gamelin, D. R. Direct Observation of sp-d Exchange Interactions in Colloidal Mn2+- and Co2+-Doped CdSe Quantum Dots. Nano Lett. 2007, 7, 10371043. (27) Mocatta, D.; Cohen, G.; Schattner, J.; Millo, O.; Rabani, E.; Banin, U. Heavily Doped Semiconductor Nanocrystal Quantum Dots. Science 2011, 332, 77-81. (28) Sahu, A.; Kang, M. S.; Kompch, A.; Notthoff, C.; Wills, A. W.; Deng, D.; Winterer, M.; Frisbie, C. D.; Norris, D. J. Electronic Impurity Doping in CdSe Nanocrystals. Nano Lett. 2012, 12, 2587-2594. (29) Oh, S. J.; Kim, D. K.; Kagan, C. R. Remote Doping and Schottky Barrier Formation in Strongly Quantum Confined Single PbSe Nanowire Field-Effect Transistors. ACS Nano 2012, 6, 4328-4334.

ACS Paragon Plus Environment

31

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 32 of 34

(30) Geyer, S. M.; Allen, P. M.; Chang, L.-Y.; Wong, C. R.; Osedach, T. P.; Zhao, N.; Bulovic, V.; & Bawendi, M. G. Control of the Carrier Type in InAs Nanocrystal Films by Predeposition Incorporation of Cd. ACS Nano 2010, 4, 7373-7378. (31) Romann, T.; Grozovski, V.; Lust, E. Formation of the Bismuth Thiolate Compound Layer on Bismuth Surface. Electrochem. Commun. 2007, 9, 2507-2513. (32) Murphy, J. E.; Beard, M. C.; Norman, A. G.; Ahrenkiel, S. P.; Johnson, J. C.; Yu, P.; Mic´ic´, O. I.; Ellingson, R. J.; Nozik, A. J. PbTe Colloidal Nanocrystals: Synthesis, Characterization, and Multiple Exciton Generation. J. Am. Chem. Soc. 2006, 128, 3241-3247. (33) Malow, T. R.; Koch, C. C. Grain Growth in Nanocrystalline Iron Prepared by Mechanical Attrition. Acta mater. 1997, 45, 2177-2186. (34) Kim, W.; Zide, J.; Gossard, A.; Klenov, D.; Stemmer, S.; Shakouri, A.; Majumdar, A. Thermal Conductivity Reduction and Thermoelectric Figure of Merit Increase by Embedding Nanoparticles in Crystalline Semiconductors. Phys. Rev. Lett. 2006, 96, 045901. (35) Toberer, E. S.; Zevalkink, A.; Synder, G. J. Phonon Engineering through Crystal Chemistry. J. Mater. Chem. 2011, 21, 15843-15852. (36) Callaway, J. Model for Lattice Thermal Conductivity at Low Temperatures. Phys. Rev. 1959, 113, 1046-1051. (37) Hartman, R. Temperature Dependence of the Low-Field Galvanomagnetic Coefficients of Bismuth. Phys. Rev. 1969, 181, 1070-1086.

ACS Paragon Plus Environment

32

Page 33 of 34

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

(38) Pan, L.; Mitra, S.; Zhao, L-D.; Shen, Y.; Wang, Y.; Felser, C.; & Berardan, D. The Role of Ionized Impurity Scattering on the Thermoelectric Performances of Rock Salt AgPbm¬SnSe2+m. Adv. Funct. Mater. 2016, 26, 5149-5157. (39) Gallo, C. F.; Chandrasekhar, B. S.; Sutter, P. H. Transport Properties of Bismuth Single Crystals. J. Appl. Phys. 1963, 34, 144-152. (40) Hostler, S. R.; Qu, Y. Q.; Demko, M. T.; Abramson, A. R.; Qiu, X.; Burda, C. Thermoelectric Properties of Pressed Bismuth Nanoparticles. Superlattices Microstruct. 2008, 43, 195-207. (41) Minnich, A. J.; Dresselhaus, M. S.; Ren, Z. F.; Chen, G. Bulk Nanostructured Thermoelectric Materials: Current Research and Future Prospects. Energy Environ. Sci. 2009, 2, 466-479. (42) Dresselhaus, M. S.; Dresselhaus, G.; Sun, X.; Zhang, Z.; Cronin, S. B.; Koga, T. LowDimensional Thermoelectric Materials. Phys. Solid state 1999, 41, 679-682. (43) Sootsman, J. R.; Chung, D. Y.; Kanatzidis, M. G. New and Old Concepts in Thermoelectric Materials. Angew. Chem. Int. Ed. 2009, 48, 8616-8639.

ACS Paragon Plus Environment

33

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 34 of 34

TOC

ACS Paragon Plus Environment

34