Open access peer-reviewed chapter

Nanostructuring Bi2Te3-Based Thermoelectric Thin-Films Grown Using Pulsed Laser Deposition

Written By

Le Thi Cam Tuyen, Phuoc Huu Le and Sheng-Rui Jian

Submitted: 06 February 2021 Reviewed: 15 July 2021 Published: 09 August 2021

DOI: 10.5772/intechopen.99469

From the Edited Volume

Materials at the Nanoscale

Edited by Awadesh Kumar Mallik

Chapter metrics overview

662 Chapter Downloads

View Full Metrics

Abstract

This book chapter reports recent advances in nanostructured Bi2Te3-based thermoelectric (TE) thin-films fabricated by pulsed laser deposition (PLD). By controlling the processing conditions in PLD growths, various fascinating Bi2Te3-based nanostructured films with promising or enhanced TE properties have been successfully fabricated, including super-assembling of Bi2Te3 hierarchical nanostructures, self-assembled Bi2Te3 films with well-aligned 0D to 3D nanoblocks, polycrystalline-nanostructured Bi2Se3 and Bi2Te3 thin-films, etc. In addition, a PLD-growth mechanism for fabricating the super-assembling Bi2Te3 thin-films is presented. This book chapter provides fundamental understanding the relationship amongst processing condition, structure-morphology, and TE property of PLD-growths Bi2Te3-based thin-films. It also presents an overview of TE materials and applications with the challenges and perspectives.

Keywords

  • Bi2Te3
  • thermoelectrics
  • self-assembly nanostructures
  • thermoelectric power factor
  • pulsed laser deposition

1. Introduction

Thermoelectric materials are solid-state energy converters whose combination of thermal, electrical, and semiconducting properties allows them to be used to convert waste heat into electricity or electrical power directly into cooling and heating [1].

1.1 Thermoelectric effects

When an electric current flows through a pair of p-type and n-type semiconductors connected in series (Figure 1(a), the holes in the p-type material and the electrons in the n-type material carry heat away from the top metal–semiconductor junctions, which leads to a cooling at the junctions called the Peltier effect. When current flows within the module, one side is cooled and the other heated. If the current is reversed, the hot and cold sides reverse also. For each material, the cooling effect is gauged by the Peltier coefficient Π that relates the heat carried by the charges to the electrical current through [1, 2, 4]: Q = Π×I.

Figure 1.

Illustration of TE devices: (a) cooler (Peltier effect), (b) power generator (Seebeck effect). Redrawn after Ref. [2]. (c) Thermoelectric module showing the direction of charge flow on both cooling and power generation [3].

In Figure 1(b), when the two ends of the materials maintain a temperature difference, the higher thermal energy holes and electrons will diffuse from the hot side to the cold side, and consequently a potential difference is created. This is Seebeck effect and it is the principle for thermocouples. The power generation is measured by the Seebeck coefficient α, which relates the voltage generated to the temperature difference through ΔV = −αΔT. The Peltier and the Seebeck coefficients are related through the Kelvin relation [1, 2]: Π = αT.

Thermoelectric devices contain many thermoelectric couples (Figure 1c, bottom), which consist of p-type (containing free holes) and n-type (containing free electrons) thermoelectric elements connected electrically in series and thermally in parallel (Figure 1c, top). A thermoelectric generator uses heat flow across a temperature gradient to power an electric load through the external circuit.

1.2 The thermoelectric figure of merit (ZT)

The performance of the thermoelectric materials is often denoted as figure of merit Z whose unit is K−1, or ZT the dimensionless unit [5, 6].

ZT=α2σκT=α2σκE+κLTE1

where α, σ, T, and κ are the Seebeck coefficient, electrical conductivity, absolute temperature, and thermal conductivity, respectively. The total thermal conductivity can be split into electronic contribution (κE) and lattice contribution (κL). The thermoelectric power factor (PF) is calculated by the quantity of α2σ. The efficiency of a thermoelectric material is determined by its ZT. Meanwhile, the maximum efficiency (η) of a power generation is expressed by [3, 7]:

η=ThTcTh1+ZT¯11+ZT¯+TcThE2

and the coefficient of performance presents for the efficiency of air-conditioning and refrigeration [7]:

COP=TcThTc1+ZT¯ThTc1+ZT¯+1E3

where Th and Tc are the hot-end and cold-end temperature of the thermoelectric materials, respectively, and T¯ is the average temperature of Th and Tc. For practical applications, it is important to use high ZT thermoelectric materials.

The best materials so far are alloys of Bi2Te3 with Sb2Te3 and Bi2Te3 with Bi2Se3. ZT is of the order of 1 at room temperature. This value gives a COP of about 1 (Figure 2a), which is still far lower than the COP = 2–4 of household refrigerators and air conditioners. Similar situation is true for power generation (Figure 2b) [2, 8]. Thermoelectric cooling and power generation generally still not competitive with the other energy conversion methods.

Figure 2.

Comparison of thermoelectric technology with other energy conversion methods for (a) cooling and (b) power generation [2, 8].

1.3 Challenges in enhancing ZT

A concept of “phonon-glass electron-crystal” (or PGEC in short) was proposed for designing efficient thermoelectric materials. This is a controversial concept from the aspect of materials science that the materials should have a high electrical conductivity as in a crystal and a low lattice thermal conductivity as in a glass [9]. However, the TE parameters are strongly interdependent, which makes the enhancement efforts of ZT very challenging. A normal approach for the enhanced properties of TE materials is to increase the power factor α2σ by optimizing the carrier concentration n, and/or to reduce the lattice thermal conductivity κL by introducing the scattering centers. These parameters are the function of carrier effective mass m* and carrier mobility μ, scattering factor r, and their interconnectivity limit ZT to approximately 1 in large bulk materials [10].

The kinetic definition of α is the energy difference between the average energy of mobile carriers and the Fermi energy [11]. When carrier concentration (n) is increased, both the Fermi energy and the average energy increase, but the Fermi energy increases more rapidly than the average energy as n is increased. Consequently, α decreases and thus α2n is dragged down rapidly. Therefore, the carrier concentration (n) increases electrical conductivity (σ) but reduces the Seebeck coefficient (α) for most of the homogeneous materials. For this reason, in metals and degenerate semiconductors with energy-independent scattering approximation, the Seebeck coefficient can be expressed as [3, 12]:

α=8π2kB23eh2mTπ3n2/3E4

Where, the parameter m* is density of states effective mass, and an increase of m* can raise the Seebeck coefficient according to the Eq. (4). However, most high m* materials have generally low μ which limits the α by a weighted mobility with a factor proportional to (m*)3/2μ. Moreover, there is no such thing as an optimal effective mass. There are high mobility low effective mass semiconductors (SiGe, GaAs) as well as low mobility high effective mass polaron conductors (oxides, chalcogenides) [3].

Noticeably, the defects scatter not only the phonons but also the electrons. When a thermoelectric material is designed for reducing lattice thermal conductivity, its carrier mobility is usually suppressed. Hence, the ratio of μ/κL determines the improvement of ZT [5, 10]. The ratio is observed to increase experimentally through a more reduction in κL rather than that in μ, but some fundamental issues in this mechanism are not understood well [10].

The electrical resistivity (ρ) and electrical conductivity (σ) are related to n through the carrier mobility μ:

1/ρ=σ=neμE5

The electronic contribution to the thermal conductivity is proportional to the electrical conductivity (σ) of the materials according to Wiedemann–Franz Law [3], and the relationship is expressed as follows:

κe=LσT=neμLTE6

where ‘e’ is electron charge, and L is Lorenz factor 2.48 × 10−8 J2/K2C2 for free electrons and this can vary particularly with carrier concentration [3, 13].

Figure 3 shows the compromise of σ, κ and α in thermoelectric materials that must be optimized to maximize the figure of merit ZT. Indeed, the lower carrier concentration will result in the lower σ and a decreasing ZT. Typically, the PF and ZT peaks occur at carrier concentrations of 1019–1021 cm−3 (depending on the material system), which falls in between common metals and heavily doped semiconductors [3]. High mobility carriers are most important for high value of electrical conductivity. Again from the Eq. (4), an increase of the carrier effective mass lead to increase the α but reduce the μ and hence the σ according to the Eq. (5). In case of the narrow semiconductor, the thermal excitation of carrier from valence band to conduction band creates holes and electrons. However, the concentration of the major carrier does not vary much. When two types of carriers are present, or bipolar effects takes place, and this is notorious to achieve effective thermoelectrics [4]. For example, the Seebeck coefficient for different carrier types is given by a weighted average of their electrical conductivity values (σe and σp) [13].

Figure 3.

Maximizing the efficiency (ZT) of a thermoelectric involves a compromise of thermal conductivity (κ; plotted on the y-axis from 0 to a top value of 10 Wm−1 K−1) and Seebeck coefficient (α; 0–500 μVK−1) with electrical conductivity (σ; 0–5000 Ω−1 cm−1) [3].

ααeσe+αpσpσe+σpE7

In short, any attempt to increase σ, will increase κe which contributes to thermal conductivity (κ). In order to counter the increment of κe, various approaches are reported to reduce κL. However, decreasing κL with phonon scattering by adding defects results in decrease in n and σ. These are the major conflicts in the properties of bulk thermoelectric materials which have been addressed in the researches for more than a half century [10].

1.4 Nanostructuring thermoelectric materials

In classical physics, the coefficients α, κe and σ are interrelated in such a way that it is impossible to increase one without affecting the others. Therefore, a compromise has to be achieved to find the maximum ZT value. Three different strategies have appeared to improve the ZT [14]:

  1. An approach for increasing α while keeping the values of σ and κe by looking for new materials with complex band structures, like heavy fermion compounds.

  2. Controlling the disorder in materials (such as Skutterudites or Clathrates) to present a rattling effect which causes, (↑) σ and decreases (↓) κL (see for instance ref. [15]).

  3. Developing nanostructured materials that could lead to (↑) α due to quantum confinement effects, while ↓κL due to the scattering of phonons at the interfaces. The latest improvements in the ZT of different materials has been achieved by this approach.

In 1993, Hicks and Dresselhaus pioneered the concept of nanostructuring in design of thermoelectric materials (i.e. Bi2Te3). The addition of the dimensionality and size of the system is added as a new parameter that affects the coupling of the electrical conductivity, Seebeck coefficient, and thermal conductivity, leading to substantially enhanced ZT [16, 17, 18]. Two ideas are dominant for the low-dimensional materials approach for improving ZT. Firstly, the presence of nanoscale constituents would introduce quantum confinement effects to enhance Seebeck coefficient and the power factor α2σ. Secondly, the numerous internal nanoinclusions and interfaces found in nanostructures would be designed so that the thermal conductivity would be reduced more than the electrical conductivity, based on differences in their respective scattering lengths [16].

As the dimensionality is decreased from 3D crystalline solids to 2D (quantum wells) to 1D (quantum wires) and finally to 0D (quantum dots), the spatial confinement are introduced that create the possibilities to tune the TE properties α, σ, and κ independently. When the system size decreases and approaches the scale comparable to the feature length of electron behavior (e.q. mean free path and wavelength) in any direction, the electronic density of states (D.O.S.) can split and become narrow as well as increase substantially (Figure 4a), resulting in the enhancement of α. Meanwhile, the thermal conductivity is also reduced because of the extensive phonon scattering at the surface, interfaces, and grain boundaries, as any dimension is less than the mean free path of phonons. Figure 4(b) illustrates examples of different nanostructuring with different dimensionalities [14]. A schematic diagram is shown in Figure 4(e) capturing these various phonon scattering mechanisms, along with the electrical transport within a thermoelectric material. For example, in material embedded nano-inclusions (nanoparticles), atomic defects are effective at scattering short wavelength phonons, but larger embedded nanoparticles are required to scatter mid- and long-wavelength phonons effectively. Grain boundaries can also play an effective role in scattering these longer-wavelength phonons [20].

Figure 4.

(a) Electronic density of states (D.O.S.) for a bulk 3D crystalline semiconductor, a 2D quantum well, a 1D nanowire or nanotube, and a 0D quantum dot [16]. (b) Examples of different nanostructuring with different dimensionalities [14]. (c) A spike in the density of states (solid line) above the bulk value (dashed line) occurs due to resonant states in Tl-doped PbTe [19]. (d) The measured ZT of Tl-PbTe and Na-PbTe samples for 300–800 K indicates an improvement due to the addition of Tl [19]. (e) Schematic diagram illustrating various phonon scattering mechanisms within a thermoelectric material, along with electronic transport of hot and cold electrons [20].

Figure 5 plots major milestones achieved for ZT over the past several decades as a function of both year and temperature [20]. In the 1950s, Bi2Te3 was first investigated as a material of great thermoelectric with ZT∼0.6 near room temperature [5, 6]. It was quickly realized that alloying with Sb2Te3 and Bi2Se3 allowed for the fine tuning of the carrier concentration alongside a reduction in lattice thermal conductivity. These compounds have played a dominant role in the field of thermoelectrics through today. The alloys of Bi2Te3 with Sb2Te3 (such as Bi0.5Sb1.5Te3; p type) and of Bi2Te3 with Bi2Se3 (such as Bi2Te2.7Se0.3; n type), with a ZT ∼ 1 at room temperature are traditional cooling materials [6]. In recent year, great enhancements in ZT owning to low dimension and nanostructure materials have been reported [19, 20, 21, 22, 23, 24, 25, 26, 27, 28, 29, 30, 31, 32] and achieved the highest ZT value of approximately 2.4.

Figure 5.

Thermoelectric figure-of-merit ZT as a function of temperature and year illustrating important milestones [20]. Although there have been several demonstrations of ZT > 1 in the past decade (2001–2010), no material has yet achieved the target goal of ZT ≥ 3. The material systems that have achieved ZT > 1 have all been based on some form of nanostructuring.

1.5 Overview of thermoelectric applications

The solid-state devices based on TE effect have the inherent advantages of reliability, silent and vibration-free operation (no moving fluids or moving parts), a very high power density, and the ability to maintain their efficiency in small scale applications where only a moderate amount of power is needed [19].

Commercial use has been made mostly from Peltier’s thermoelectric cooling (TEC) effect in applications, as demonstrated in Figure 6 [35]:

  • Small refrigerator devices are used for camping and outdoor activities. For example, the cooler/warmer TE device (Engel Thermo 8) has volume 8 L and weighing just over 3 kg. Its features include cooling performance up to 22°C below ambient temperature and warming up to +65°C.

  • Gentherm designed and developed Automotive Climate Control Seat [36], which has TE heat pumps in the back and bottom cushions. The TE system makes conditioned flowing air through channels to the occupant for providing on-demand cooling or heating. As shown in the first panel in Figure 6, the seat has the heat pump consisting of a TE module (green box) and a fan (orange).

  • Thermal management of tiny laser diodes is used in fiber optic telecom, datacom backhaul networks. TEC can also be used for contact cooling of semiconductor lasers, infrared detectors, CCD- matrix, and miniconditioners for photomultipliers.

  • Localized cooling at hot spots of chips was created. For example, the Intel group is the first to demonstrate both concepts of applying the TE material only to a chip’s hottest spots (Figure 6) [33, 37]. On the substrate, the researchers grew a 100-μm-thick layered structure, called a superlattice, containing bismuth, tellurium, antimony, and selenium. The structure can pump 1300W/cm2 heat from the back side of the chip to the heat spreader. The superlattice induced an approximately 6°C temperature drop at the hot spot even before the device was powered up, because it conducts heat better than the grease that bonds the rest of the heat spreader to the chip. Yet, when a 3 A- current went through the thermoelectric cooler, the total temperature change was only of 15°C. Managing heat in electronics is a common issue, and TE coolers can improve electronic systems in thermal performance, cost, noise, weight, size or efficiency.

Figure 6.

Overview of potential thermoelectric cooling (TEC) applications [33, 34].

Figure 7 shows an overview of the present and potential applications of thermoelectric generators (TEGs) [34]. They include (1) heating systems and water boilers with TEG units which generate the electricity for the control units and pumping systems, (2) the long term perspective of waste heat recovery for medium-scale industrial facilities, (3) waste heat recovery in automobiles and other combustion-engine-powered vehicles for enhanced efficiency and electric current supply of the electronic system, (4) miniaturized autarkic sensor systems powered by an integrated TEG with a wireless data transmitter, (5) ventilated wood stove powered by a thermoelectric generator with enhanced oxygen supply, improves burning process.

Figure 7.

Overview of potential thermoelectric generator (TEG) applications [33, 34].

Advertisement

2. Nanostructured Bi2Te3-based thermoelectric thin films grown using pulsed laser deposition

2.1 PLD growths of nanostructured Bi2Te3-based thin films

PLD is one of the most convenient thin film growth techniques that uses a high intensity pulsed laser beam as an external energy source to ablate a target, form a plume, and deposit thin films onto a substrate. In practice, a large number of variables affect the properties of the film, such as substrate temperature (Ts), background gas pressure (P) and laser fluence. Figure 8 shows a PLD system for preparing thermoelectric thin films [38, 39]. The substrate was heated and maintained at desired Ts using a thermocouple and a proportional-integral-derivative temperature controller. The thermocouple was buried inside a substrate holder which was heated by a tungsten lamp or electrical resistance heating. The pressure of ambient gas (He, Ar) could be fine-tuned by the needle valve. Laser source can be KrF excimer laser beam (λ = 248 nm) and Q-switched Nd:YAG laser (λ = 355 nm) with properly selected laser fluence (e.g., 3.8, 6.2, or 8.3 J/cm2) pulsed duration of 5–20 ns, repetition rate of 5–10 Hz [38, 40, 41, 42]. The laser beam was guided by several UV mirrors and focused on a stoichiometric polycrystalline target (e.g., Bi2Se3, Bi2Te3, Bi0.5Sb1.5Te3, etc.) inside the vacuum chamber by the UV lens. The deposition chamber was evacuated to a base pressure of ∼10−6 Torr, and high-purity ambient gas (He or Ar) was then introduced until obtaining a target pressure (e.g., usually 10−5 – 3×10−1 Torr).

Figure 8.

A schematic illustration of a PLD system.

The enhancement of the PF of Bi2Te3-based thin films is challenging due to the coupling among TE material properties [3], and the difficulty in growing stoichiometric films [38]. Indeed, stoichiometry is a key factor for obtaining better TE properties [5, 38, 43, 44, 45]. Yet, both tendency for re-evaporation of volatile elements (i.e., Te, Se) at elevated Ts [45, 46, 47, 48] and the low sticking coefficient Te (< 0.6 for Bi2Te3) at Ts beyond 300°C [49, 50] constrain to grow stoichiometric Bi2Te3-based films (Figure 9a and b).

Figure 9.

(a) Vapor pressures of Bi, Sb, Te, Se, Bi2Se3, and Bi2Te3 as a function of temperature [46]. (b) The variation of sticking coefficient Ks (Bi, Te) as a function of substrate temperature Ts at fixed flux ratio FR = 4.5 [49].

2.2 Super-assembling of Bi2Te3 hierarchical nanostructured thin films

C.-H. Chen et al. [41] reported the PLD growths of super-assembling of Bi2Te3 hierarchical nanostructured thin films on the SiO2/Si substrates and their thermoelectric properties. Interesting Bi2Te3 super-assemblies were successfully grown using PLD with controlling the substrate temperatures from 350–600°C and at a fixed Ar ambient pressure of approximately 10−3 Torr. SEM images in Figure 10 clearly shows the morphological characteristics of the superassembling Bi2Te3 nanostructured thin films [41]. At lower deposition temperatures (< 450°C), the films are mainly composed of vertically aligned nanoscaled flakes, but flakes are horizontally stacked for 600°C-film (Figure 10ad). Moreover, the bottom of each of the deposited super-assemblies has a relatively continuous and dense layer, and this layer thickness increases with increasing substrate temperature from 350–450°C (Figure 10ad). The top-view SEM images confirm for the high uniformity and presents the unique super-assembling features of the repetitively and regularly assembled nano-flakes (Figure 10eh). These four films are uniformly composed of spindle-like (Figure 10e), worm-like (Figure 10f and g) and island-like (Figure 10h) hierarchical nanostructures. Magnified top-view SEM images (Figure 10ik) further show that the nanoflakes are composed of oriented and regular assemblies of numerous rice-like and elongated primitive nanoparticles [41]. At a higher substrate temperature, thin- and large-size nanoflakes are formed from packing of dense rice-like nanoparticles, driving by the relatively sufficient thermal energy for diffusion. In addition, the out-of- plane superassembly structure (600°C) has a limited column width, which is not always consistent along the out-of-plane direction (Figure 10d). Also, the parallel nano-flakes (at 600°C) are evidently formed by flake stacking along c-axis orientation or epitaxial-like growth. The special three-dimensional mesh-like structure of 600°C-film would also be an effective design for scattering phonons, and it’s extremely smooth top surface is certainly beneficial for subsequent analyses and applications [41].

Figure 10.

(a)–(d) The cross-sectional and (e)–(l) the corresponding top-view SEM images of the Bi2Te3 superassemblies deposited at 350°C, 400°C, 450°C, and 600°C, respectively [41].

Figure 11(a) shows the crystal structure of Bi2Te3, which is usually described by a hexagonal cell that consists of 15 layers of atoms stacking along the c-axis with a sequence [5], namely ···Te(1)–Bi–Te(2)–Bi–Te(1) ··· Te(1)–Bi–Te(2)–Bi–Te(1)···Te(1)–Bi–Te(2)–Bi–Te(1) ···. The superscripts refer to two different types of bonding for Te atoms. The 5-atomic-layer thick lamellae of–(Te(1)–Bi–Te(2)–Bi–Te(1))– is called quintuple layers, QLs. The Te(1)…Te(1) refers Van der Waals force between Te atoms, whereas the Te(1)–Bi and Bi–Te(2) are ionic-covalent bonds. This weak binding between the Te(1)…Te(1) accounts for the anisotropic thermal and electrical transport properties of Bi2Te3. For example, the thermal conductivity along the c-axis direction (∼0.7 Wm−1K −1) is approximately a haft of the value along the plane perpendicular to the c-axis (∼1.5 Wm−1K −1) [5, 6, 13]. The weak binding of Te(1)…Te(1) also make the ease of cleavage along the plane perpendicular to the c-axis.

Figure 11.

(a) Crystal structures of Bi2Te3. (b) X-ray diffraction (XRD) patterns of of the Bi2Te3 super-assemblies deposited at various deposition temperatures from 350–600°C and at an Ar ambient pressure ∼10−3 Torr [41].

Figure 11(b) shows XRD patterns of the Bi2Te3 super-assemblies deposited at various substrate temperatures from 350–600°C. Clearly, all the films exhibited rhombohedral Bi2Te3 (JCPDS no. 89–4302) without traceable impurities or oxides. When substrate temperature increases, the (00 l) preferential orientation gradually becomes stronger, the 600°C- film is highly (00 l)-preferred orientation, which is consistent with the SEM observation (Figure 10d). The gradually enhanced (00 l) peaks from 350–450°C mainly originate from the increased thickness of the bottom layer (Figure 10ad), which has similar lamellar morphology with (00 l)-preferential orientation of 600°C-film [41].

Figure 12 presents the proposed growth model of the super-assemling nanostructured Bi2Te3 films prepared at various TS. The growth mechanisms are layer-then-fake for TS = 350–450°C and layer-by-layer for higher TS of 600°C. We can only find a monotone morphology and single preferential orientation of (00 l) for 600°C-film, which lead to a fully lamellar morphology with the (00 l) preferential orientation. Meanwhile, drastic changes in morphologies from layer to flake and orientations from (00 l) to (015) are observed at lower temperatures (350–450°C). The (00 l)-preferred orientation should be attributed to the thin bottom layer of the films prepared at 350–450°C. The thickness of this layer increases with increasing TS. Since the bottom layer at 350°C is extremely thin, the required Ts for obtaining layer growth should be just below 350°C. The drastic change in the morphology and orientation at TS of 350–450°C, namely, the layer- then-flake growth can be induced by a temperature gradient along the growth direction that the temperature at top surface of the as-deposited film should be slightly lower than at the substrate [41].

Figure 12.

Schematic illustration of the layer-then-flake and layer-by-layer growth models and the resulting Bi2Te3 in-plane (350–450°C) and out- of-plane (600°C) super-assemblies. Inset is the optical image of the prepared super-assembled films with a size of 1.5 × 1.5 cm2 [41].

Table 1 summarizes the detailed properties of the super-assembling nanostructured Bi2Te3 thin-films. Due to such the voided structures, the films exhibited low electrical conductivity from 49 S.cm−1 for worm-like superassembly (450°C) to 160 S.cm−1 for 3D-layered super-assembly (600°C). Seebeck coefficient of the films was in range of 113–138 μV/K. As a result, the power factor (PF) is relatively low in range of 0.93 to 3.0 μW/cmK2, primarily due to the low electrical conductivity of the films with voided morphologies.

MaterialMorphologyMethodDeposition conditionsn (1019 cm−3)μ (cm2/Vs)σ (S/cm)α (μV/K)PF (μW/cmK2)Ref.
Bi2Te3Spindle-like super-assemblyPLD350°C0.13 Pa4.012.479−1131.01[41]
Bi2Te3Worm-like super-assemblyPLD400°C0.13 Pa1.925.973−1191.03[41]
Bi2Te3worm-like super-assemblyPLD450°C0.13 Pa1.229.449−1380.93[41]
Bi2Te33D-layered super-assemblyPLD600°C0.13 Pa5.120.3160−1373.0[41]
Bi2Te3NanoparticlesPLD300°C20 Pa9.714.8230−911.90[51]
Bi2Te3NanoparticlesPLD300°C1.0 Pa1058.31390−605.0[52]
Bi2Te3Layered StructureSputtering350°C1.0 Pa9512.11840−708.8[52]
Nanorods250°C0.9 Pa9.12.029−810.19[51]
Bi2Te3Columnar StructureSputtering350°C1.0 Pa2467.52990−466.4[52]
Bi2Se3Layered HPsPLD300°C40 Pa7.481.4963.8−75.85.5[38]
Bi2Te3Compact filmPLD300°C80 Pa5102814.3−172.824.3[40]
Bi2Te3Layered-smooth filmPLD250°C10 Pa10.190.61464−18650.6[43]
Bi3Se2TeNanocrystalline filmPLD250°C40 Pa35.534.41747.5−68.88.3[53]
Bi2Se3BulkMelting and hot-pressing251.9−1757.7[54]
Bi2Se0.3Te2.7BulkBall milling-hot pressing892−19032.2[55]
Bi2Se1.5Te1.5BulkZone melting1.2230441.6−19316.5[56]
Bi2Se1.8Te1.2Nano-platelet bulkPolyol method199.6−80.91.3[57]
Bi2Se2TeBulkBall milling- hot pressing1613−605.8[55]

Table 1.

Material, type, method, processing conditions, carrier concentration (n), mobility (μ), electrical conductivity (σ), Seebeck coefficient (α), power factor (PF = α2σ) of the Bi2Te3 films deposited by PLD and RF sputtering, as compared to properties of Bi2Se3, Bi2Se3, Bi2SexTe1-x bulk and films reported in the literature. All the selected values were recorded at room temperature.

In PLD, tightly controlling substrate temperatures (Ts) and ambient pressures (P) enables the morphologies and compositions of films to be manipulated extensively, which offers a new method for enhancing the TE properties of films [38, 43, 51, 58, 59]. For example, self-assembled Bi2Te3 films featuring well-aligned zero- to three-dimensional nanoblocks have been fabricated (Figure 13ad), but the room-temperature PFs of these films remain low (≤ 1.9 μWcm−1 K−2) [51]. By contrast, A. Li Bassi et al. [43] obtained several microstructured Bi2Te3 films (Figure 13eh) with high PFs for morphologies: layered-smooth (50.6 μWcm−1 K−2, Figure 13e,e1), and compact-smooth (21.2 μWcm−1 K−2, Figure 13f,f1) at room-temperature; whereas the PFs remained low values of 8.8 μWcm−1 K−2 for 3D crystallite shapes (Figure 13g,g1) and 0.08 μWcm−1 K−2 for 3D-voided platelets (Figure 13h,h1).

Figure 13.

The morphology and power factor (unit μWcm−1 K−2) of nano/micro-structured Bi2Te3 thin-films grown by PLD at various substrate temperatures and ambient pressures, reported by (i) Chang and Chen [51] and (ii) Li Bassi et al. [43].

Table 1 summaries the morphology and properties of Bi2Te3-based thin-films deposited by PLD, sputtering, in comparison with the properties of TE bulks. Usually, TE nanomaterials possess low σ values due to the separating or voided structure-morphology, but bulk and thin films have superior σ. For example, the compact-polycrystalline Bi3Se2Te achieved σ = 1747.5 S/cm [53] or even higher for some other films [52]. Unfortunately, the κ of films are missed in many published works to calculate ZT of the films. Thermal conductivity (κ) of nanocrystalline and nanostructured thermoelectrics is expected to achieve low values thanks to the extensive phonons scattering at interfaces, surfaces and grain boundaries. Indeed, reduced κ values have been noted for the monocrystalline Bi2Se0.3Te2.7 films (κ = 0.8 W/m K for an average grain size of 60 nm) [60], and for Bi-Sb-Te films [61, 62].

2.3 Thermal conductivity κ of Bi2Se3 and Bi2Te3 and Bi-Te-Se compounds

A transient 3ω technique is usually employed in measuring thermal conductivity of thermoelectric films. The detail of this technique can be found in refs. [62, 63, 64]. Table 2 summarizes thermal transport properties (at room–temperature) of nanocrystalline–nanostructured Bi2Te3-based thin films and bulk materials in the literature. Generally, the thermal conductivity κ value for polycrystalline films is expected to be smaller than that of bulk alloys because of the extensive phonons scattering at interfaces, surfaces and grain boundaries [5, 60, 66]. Moreover, the κ of nanocrystalline Bi2Te3-based films will further decrease when the grain size of decreases (κ ≤ 0.81 W/mK, Figure 14A) [62, 65]. For Bi2Te3/Sb2Te3 superlattice films, the coherent backscattering of phonon waves at the superlattice interfaces is outlined for the reduction of lattice thermal conductivity, resulting in the low κ ≤ 0.4 W/mK [67, 69].

Sample, fabrication methodAvg. grain sizeκ (W/m K)σ (S/cm)α (μV/K)PF = σα2 (μW/cmK2)ZT (300 K)Ref.
Bi2Te2.7Se0.3 nanocrystalline thin film, flash evaporation60 nm0.8 (cross-plane)540−186.1 (in-plane)18.7 (in-plane)0.7[60]
Sintered bulk Bi2Te3-xSex material, hot-pressing30 μm1.6930−177.529.30.6
Nanocrystalline bismuth-telluride-based (Bi2Te3-xSex) thin film10 nm0.61550−84.03.90.19[65]
27 nm0.68540−138.110.30.46
60 nm0.80540−186.118.70.70
Nanocrystalline Bi-Sb-Te thin film, sputtering26 nm0.463.3[62]
45 nm0.656.7
84 nm0.8133.3
Nanocrystalline BiSbTe (8:30:62) thin film, flash evaporation150 nm0.6[66]
Single crystal BiSbTe bulk alloys0.75[5]
Bi2Te3/Sb2Te3 superlattices (period∼5 nm)0.4[67]
Bi2Te3+0.63 bulk2.21000−240580.87[5]
Bi2(Te0.95Se0.05)3 bulk1.59901−223450.85[5]
Bi2Te3/Bi2(Te0.88Se0.12)3 superlattice film, MBE80 nm1.25639−204270.60[68]
Bi2Te3 film, PLD1.1[69]
Bi2Te3/Sb2Te3 superlattices film (layered thickness ∼ 6 nm), PLD.0.11
Bi2Te3 films, laser ablation0.2–0.3[70]
BixSb2-xTe3 nanolayer film, PLD190 nm1.16270095250.65[71]
BixSb2-xTe3 nanodisc film, PLD100 nm1.001100132200.60
BixSb2-xTe3 nanocolumn film, PLD70 nm0.93280207120.39

Table 2.

Room–temperature thermal transport properties of nanocrystalline–nanostructured Bi2Te3-based thin films and bulk materials in the literature, included: sample and fabrication method, average grain size, thermal conductivity κ, electrical conductivity σ, Seebeck coefficient α, power factor PF (= α2σ), and ZT (at 300 K).

Figure 14.

The morphology and thermal conductivity of Bi2Te3-based films with different grain sizes: (A) nanocrystalline Bi2Te3-xSex films [65], (B) the BixSb2-xTe3 films [71].

For PLD Bi2Te3-based films, Yamasaki et al. [69] measured thermal conductivity with an ac calorimetric method in the direction across the film, obtaining k∼1.1 W/m K for a Bi2Te3 film deposited by PLD in vacuum (Table 2). In addition, Walachova et al. [70] used direct ZT measurement with the Harman method to estimate the κ value, and it is about 0.2–0.3 W/mK for the Bi2Te3 films. Recently, Chang et al. [71] reported the κ values of 0.93–1.16 W/mK for BixSb2-xTe3 films with the granular-layered morphologies (Figure 14B).

Advertisement

3. Conclusions

In this book chapter, we present an overview of thermoelectric materials and applications, challenging of enhancing TE properties, and the nanostructuring approach in development TE materials. Various interesting nanostructured Bi2Te3- based thin films have been grown successfully by PLD with properly controlled substrate temperatures ambient gas pressures. For example, super-assembling of Bi2Te3 hierarchical nanostructures were grown at TS from 350 to 600°C, and the films possessed relative high Seebeck coefficient of 113–138 μV/K, but exhibited low electrical conductivities of 49–160 S.cm−1, and thus they had relatively low PF in range of 0.93 to 3.0 μW/cmK2. Our intensive literature review on Bi2Te3-based TE materials can make general conclusion that TE nanomaterials possess low σ values when their structure-morphology are separating or voided, meanwhile, bulk and compact-smooth thin films can achieve high σ values. The PF values of Bi2Te3-based thermoelectrics varied in a wide range, i.e. below 5.0 μW/cmK2 for voided structure-morphology, and reaching intermediate-high PF values of 5.0–50.6 μW/cmK2 for compact-smooth or compact-layered structures. An advantage of nanocrystalline and nanostructuring thermoelectrics is the reduced thermal conductivity (possibly below 1 W.m−1 K−1). This book chapter provides fundamental understanding the relationship amongst processing condition in PLD growths, structure-morphology, and TE properties of Bi2Te3-based thin films.

Advertisement

Acknowledgments

This research is funded by Vietnam National Foundation for Science and Technology Development (NAFOSTED) under grant number 103.02-2019.374, and the Ministry of Science and Technology, Taiwan under Contract Nos. MOST 109-2221-E-214-016 and MOST 110-2221-E-214-013.

References

  1. 1. Bell LE. Cooling, heating , generating power, and recovering waste heat with thermoelectric systems. Science. 2008; 321: 1457–1461. DOI: 10.1126/science.1158899
  2. 2. Chen G., Shakouri A. Heat Transfer in Nanostructures for Solid-State Energy Conversion. J. Heat Transfer. 2002; 124: 242. DOI: 10.1115/1.1448331.
  3. 3. Snyder GJ, Toberer ES. Complex thermoelectric materials. Nat. Mater. 2008; 7:105–114. DOI: 10.1038/nmat2090
  4. 4. Li J-F, Liu W-S, Zhao L-D, Zhou M. High-performance nanostructured thermoelectric materials. NPG Asia Mater. 2010; 2: 152. DOI: 10.1038/asiamat.2010.138
  5. 5. Rowe DM. Thermoelectrics handbook: macro to nano. D.M. Rowe Ed. Boca Raton: CRC Press; 2006. 1014 p. DOI: 10.1201/9781420038903
  6. 6. Nolas GS, Sharp J, Goldsmid HJ, Thermoelectrics: Basic Principles and New Materials Developments. New York: Springer; 2001.
  7. 7. Lan Y, Minnich AJ, Chen G, Ren Z. Enhancement of thermoelectric figure-of-merit by a bulk nanostructuring approach. Adv. Funct. Mater. 2010; 20: 357–376. DOI: 10.1002/adfm.200901512
  8. 8. http://www.electronics-cooling.com/2005/11/advances-in-high-performance-cooling- for-electronics/
  9. 9. Nolas GS, Morelli DT, Tritt TM. KUTTERUDITES: A Phonon-Glass-Electron Crystal Approach to Advanced Thermoelectric Energy Conversion Applications. Annu. Rev. Mater. Sci. 1999; 29: 89. DOI: 10.1146/annurev.matsci.29.1.89.
  10. 10. Alam H and Ramakrishna S. A review on the enhancement of figure of merit from bulk to nano-thermoelectric materials. Nano Energy. 2013; 2: 190–212. DOI: 10.1016/j.nanoen.2012.10.005
  11. 11. Pichanusakorn P, Bandaru P. Nanostructured thermoelectrics. Mater. Sci. Eng. R. 2010; 67:19–63. DOI: 10.1016/j.mser.2009.10.001
  12. 12. Cutler M, Leavy JF, Fitzpatrick RL. Electronic transport in semimetallic cerium sulfid. Phys. Rev. 1964; 133: A1143.
  13. 13. Tritt TM. Thermoelectric Phenomena, Materials, and Applications. Annu. Rev. Mater. Res. 2011; 41: 433. DOI: 10.1146/annurev-matsci-062910-100453.
  14. 14. Martín-González M, Caballero-Calero O, Díaz-Chao P. Nanoengineering thermoelectrics for 21st century: Energy harvesting and other trends in the field. Renew. Sustain. Energy Rev. 2013; 24: 288. DOI: 10.1016/j.rser.2013.03.008.
  15. 15. Eilertsen J, Li J, Rouvimov S, Subramanian MA. Thermoelectric properties of indium-filled InxRh4Sb12 skutterudites. J. Alloys Compd. 2011; 509: 6289. DOI: 10.1016/j.jallcom.2011.03.057.
  16. 16. Dresselhaus MS, Chen G, Tang MY, Yang RG, Lee H, Wang DZ, Ren ZF, Fleurial J-P, Gogna P. New Directions for Low-Dimensional Thermoelectric Materials. Adv. Mater. 2007; 19: 1043. DOI: 10.1002/adma.200600527.
  17. 17. Hicks LD, Dresselhaus MS. Effect of quantum-well structures on the thermoelectric figure of merit. Phys. Rev. B 1993; 47: 12727. DOI: 10.1103/PhysRevB.47.12727
  18. 18. Hicks LD, Dresselhaus MS. Thermoelectric figure of merit of a one-dimensional conductor. Phys. Rev. B 1993; 47: 16631. DOI: 10.1103/PhysRevB.47.16631
  19. 19. Heremans JP, Jovovic V, Toberer ES, Saramat A, Kurosaki K, Charoenphakdee A, Yamanaka S, Snyder GJ. Enhancement of thermoelectric efficiency in PbTe by distortion of the electronic density of states. Science. 2008; 321: 554–557. DOI: 10.1126/science.1159725
  20. 20. Vineis CJ, Shakouri A, Majumdar A, Kanatzidis MG. Nanostructured thermoelectrics: big efficiency gains from small features. Adv. Mater. 2010; 22: 3970. DOI: 10.1002/adma.201000839.
  21. 21. Androulakis J, Lin C-H, Kong H-J, Uher C, Wu C-I, Hogan T, Cook BA, Caillat T, Paraskevopoulos KM, Kanatzidis MG. Spinodal decomposition and nucleation and growth as a means to bulk nanostructured thermoelectrics: enhanced performance in Pb1-xSnxTe-PbS. J. Am. Chem. Soc. 2007; 129: 9780. DOI: 10.1021/ja071875h.
  22. 22. Rhyee J-S, Lee KH, Lee SM, Cho E, Kim SI, Lee E, Kwon YS, Shim JH, Kotliar G. Peierls distortion as a route to high thermoelectric performance in In(4)Se(3-delta) crystals. Nature 2009; 459: 965. DOI: 10.1038/nature08088.
  23. 23. Saramat A, Svensson G, Palmqvist AEC, Stiewe C, Mueller E, Platzek D, Williams SGK, Rowe DM, Bryan JD, Stucky GD. Large thermoelectric figure of merit at high temperature in Czochralski-grown clathrate Ba8Ga16Ge30. J. Appl. Phys. 2006; 99: 023708. DOI: 10.1063/1.2163979.
  24. 24. Wang XW, Lee H, Lan YC, Zhu GH, Joshi G, Wang DZ, Yang J, Muto AJ, Tang MY, Klatsky J, Song S, Dresselhaus MS, Chen G, Ren ZF. Enhanced thermoelectric figure of merit in nanostructured n-type silicon germanium bulk alloy. Appl. Phys. Lett. 2008; 93: 193121. DOI: 10.1063/1.3027060.
  25. 25. Venkatasubramanian R, Siivola E, Colpitts T, O’Quinn B. Thin-Film thermoelectric devices with high room-temperature figures of merit. Nature. 2001; 413: 597–602. DOI: 10.1038/35098012
  26. 26. Hsu KF, Loo S, Guo F, Chen W, Dyck JS, Uher C, Hogan T, Polychroniadis EK, Kanatzidis MG. Cubic AgPbmSbTe2+m: bulk thermoelectric materials with high figure of merit. Science. 2004; 303: 818. DOI: 10.1126/science.1092963.
  27. 27. Chung DY, Hogan T, Brazis P, Rocci-Lane M, Kannewurf C, Bastea M, Uher C, Kanatzidis MG. CsBi4Te6: A High-Performance Thermoelectric Material for Low-Temperature Applications. Science 2000; 287: 1024. DOI: 10.1126/science.287.5455.1024.
  28. 28. Harman TC, Taylor PJ, Walsh MP, LaForge BE. Quantum dot superlattice thermoelectric materials and devices. Science. 2002; 297: 2229. DOI: 10.1126/science.1072886.
  29. 29. Boukai AI, Bunimovich Y, Tahir-Kheli J, Yu J-K, Goddard WA, Heath JR. Silicon nanowires as efficient thermoelectric materials. Nature. 2008; 451: 168. DOI: 10.1038/nature06458.
  30. 30. Hochbaum AI, Chen R, Delgado RD, Liang W, Garnett EC, Najarian M, Majumdar A, Yang P. Enhanced thermoelectric performance of rough silicon nanowires. Nature 2008; 451: 163. DOI: 10.1038/nature06381.
  31. 31. 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 MS, Chen G, Ren Z. High-thermoelectric performance of nanostructured bismuth antimony telluride bulk alloys. Science. 2008: 320: 634. DOI: 10.1126/science.1156446.
  32. 32. Poudeu PFP, Angelo JD’, Downey AD, Short JL, Hogan TP, Kanatzidis MG. High thermoelectric figure of merit and nanostructuring in bulk p-type Na1-xPbmSbyTem+2. Angew. Chem. Int. Ed. Engl. 2006; 45: 3835. DOI: 10.1002/anie.200600865.
  33. 33. Chowdhury I, Prasher R, Lofgreen K, Chrysler G, Narasimhan S, Mahajan R, Koester D, Alley R, Venkatasubramanian R. On-chip cooling by superlattice-based thin-film thermoelectrics. Nat. Nanotechnol. 2009; 4: 235. DOI: 10.1038/NNANO.2008.417.
  34. 34. Nielsch K, Bachmann J, Kimling J, Böttner H. Thermoelectric Nanostructures: From Physical Model Systems towards Nanograined Composites. Adv. Energy Mater. 2011; 1: 713–731. DOI: 10.1002/aenm.201100207
  35. 35. http://www.electronics-cooling.com/2011/09/thin-film-thermoelectrics-today-and- tomorrow/
  36. 36. Heremans JP, Dresselhaus MS, Bell LE, Morelli DT. When thermoelectrics reached the nanoscale. Nat. Nanotechnol. 2013; 8: 471. DOI: 10.1038/nnano.2013.129.
  37. 37. http://spectrum.ieee.org/computing/hardware/intelled-team-demonstrates-first-chipscale-thermoelectric-refrigerator
  38. 38. Le PH, Liao C-N, Luo CW, Lin J-Y, Leu J. Thermoelectric properties of bismuth-selenide films with controlled morphology and texture grown using pulsed laser deposition. Appl. Surf. Sci. 2013; 285P: 657–663. DOI: 10.1016/j.apsusc.2013.08.107
  39. 39. Le PH, Luo CW. Thermoelectric and Topological Insulator Bismuth Chalcogenide Thin Films Grown Using Pulsed Laser Deposition. INTECH, UK. 2016; 55–84 DOI: 10.5772/65898
  40. 40. Le PH, Liao C-N, Luo CW, Leu J. Thermoelectric properties of nanostructured bismuth–telluride thin films grown using pulsed laser deposition. J. Alloy. Compd. 2014; 615: 546–552. DOI: 10.1016/j.jallcom.2014.07.018
  41. 41. Chang H-C, Chen T-H, Whang W-T, Chen C-H. Superassembling of Bi2Te3 hierarchical nanostructures for enhanced thermoelectric performance. J. Mater. Chem. A 2015; 3: 10459–10465. DOI: 10.1039/c5ta00911a.
  42. 42. Chen T-H, Lin P-Y, Chang H-C, Chen C-H. Enhanced thermoelectricity of three- dimensionally mesostructured BixSb2-xTe3 nanoassemblies: from micro-scaled open gaps to isolated sealed mesopores, Nanoscale. 2017; 9: 3283. DOI: 10.1039/c7nr00132k.
  43. 43. Li Bassi A, Bailini A, Casari CS, Donati F, Mantegazza A, Passoni M, Russo V, Bottani CE. Thermoelectric properties of Bi–Te films with controlled structure and morphology. J. Appl. Phys. 2009; 105: 124307. DOI: 10.1063/1.3147870
  44. 44. Makala RS, Jagannadham K, Sales BC. Pulsed laser deposition of Bi2Te3-based thermoelectric thin films. J. Appl. Phys. 2003; 94: 3907. DOI: 10.1063/1.1600524.
  45. 45. Goncalves LM, Couto C, Alpuim P, Rolo AG, Völklein F, Correia JH. Optimization of thermoelectric properties on Bi2Te3 thin films deposited by thermal co-evaporation. Thin Solid Films. 2010; 518: 2816. DOI: 10.1016/j.tsf.2009.08.038.
  46. 46. Noro H, Sato K, Kagechika H. The thermoelectric properties and crystallography of Bi-Sb-Te-Se thin films grown by ion beam sputtering. J. Appl. Phys. 1993; 73: 1252–1260. DOI: 10.1063/1.353266
  47. 47. Navrátil J, Horák J, Plecháček T, Kamba S, Lošt’ák P, Dyck JS, Chen W, Uher C. Conduction band splitting and transport properties of Bi2Se3. J. Solid State Chem. 2004; 177: 1704. DOI: 10.1016/j.jssc.2003.12.031.
  48. 48. Hor Y, Richardella A, Roushan P, Xia Y, Checkelsky J, Yazdani A, Hasan M, Ong N, Cava R. p-type Bi2Se3 for topological insulator and low-temperature thermoelectric applications. Phys. Rev. B. 2009; 79: 195208. DOI: 10.1103/PhysRevB.79.195208.
  49. 49. Mzerd A, Sayah D, Tedenac JC, Boyer A. Crystal growth and sticking coefficient of Bi2Te3 thin films on Si(1 1 1) substrate. J. Mater. Sci. Lett. 1995; 14: 194. DOI: 10.1007/BF00318254
  50. 50. Peranio N, Winkler M, Bessas D, Aabdin Z, König J, Böttner H, Hermann RP, Eibl O. Room-temperature MBE deposition, thermoelectric properties, and advanced structural characterization of binary Bi2Te3 and Sb2Te3 thin films. J. Alloys Compd. 2012; 521: 163. DOI: 10.1016/j.jallcom.2012.01.108.
  51. 51. Chang H-C, Chen C-H. Self-assembled bismuth telluride films with well-aligned zero- to three-dimensional nanoblocks for thermoelectric applications. CrystEngComm. 2011; 13: 5956. DOI: 10.1039/c1ce05350g
  52. 52. Deng Y, Liang H-M, Wang Y, Zhang Z-W, Tan M, Cui J-L. Growth and transport properties of oriented bismuth telluride films. J. Alloys Compd. 2011; 509: 5683–4587. DOI: 10.1016/j.jallcom.2011.02.123
  53. 53. Tuyen LTC, Le PH, Luo CW, Leu J. Thermoelectric properties of nanocrystalline Bi3Se2Te thin films grown using pulsed laser deposition. J. Alloy. Compd. 2016; 673: 107–114. DOI: 10.1016/j.jallcom.2016.03.006
  54. 54. Sun G, Qin X, Li D, Zhang J, Ren B, Zou T, Xin H, Paschen SB, Yan X. Enhanced thermoelectric performance of n-type Bi2Se3 doped with Cu. J. Alloys Compd. 2015; 639: 9–14. DOI: 10.1016/j.jallcom.2015.03.124.
  55. 55. Liu W, Lukas KC, McEnaney K, Lee S, Zhang Q, Opeil CP, Chen G, Ren Z. Studies on the Bi2Te3–Bi2Se3– Bi2S3 system for mid-temperature thermoelectric energy conversion. Energy Environ. Sci. 2013; 6: 552–560. DOI: 10.1039/c2ee23549h
  56. 56. Wang S, Tan G, Xie W, Zheng G, Li H, Yang J, Tang X. Enhanced thermoelectric properties of Bi2(Te1-xSex)3- based compounds as n-type legs for low-temperature power generation. J. Mater. Chem. 2012; 22: 20943. DOI: 10.1039/c2jm34608
  57. 57. Soni A, Yanyuan Z, Ligen Y, Aik MKK, Dresselhaus MS, Xiong Q. Enhanced thermoelectric properties of solution grown Bi2Te3-xSex nanoplatelet composites. Nano Lett. 2012; 12: 1203–1209. DOI: 10.1021/nl2034859
  58. 58. Bailini A, Donati F, Zamboni M, Russo V, Passoni M, Casari CS, Li Bassi A, Bottani CE. Pulsed laser deposition of Bi2Te3 thermoelectric films. Appl. Surf. Sci. 2007; 254: 1249. DOI: 10.1016/j.apsusc.2007.09.039.
  59. 59. Obara H, Higomo S, Ohta M, Yamamoto A, Ueno K, Iida T. Thermoelectric Properties of Bi2Te3-Based Thin Films with Fine Grains Fabricated by Pulsed Laser Deposition. Jpn. J. Appl. Phys. 2009; 48: 085506. DOI: 10.1143/JJAP.48.085506.
  60. 60. Takashiri M, Takiishi M, Tanaka S, Miyazaki K, Tsukamoto H. Thermoelectric properties of n-type nanocrystalline bismuth-telluride-based thin films deposited by flash evaporation. J. Appl. Phys. 2007; 101: 074301. DOI: 10.1063/1.2717867.
  61. 61. Liao C-N, Su X-W, Liou K-M, Chu H-S. Electrical and thermal transport properties of electrically stressed Bi–Sb–Te nanocrystalline thin film. Thin Solid Films 2011; 519: 4394. DOI: 10.1016/j.tsf.2011.02.066.
  62. 62. Liao C-N., Wang Y-C., Chu H-S. Thermal transport properties of nanocrystalline Bi–Sb–Te thin films prepared by sputter deposition. J. Appl. Phys. 2008; 104: 104312. DOI: 10.1063/1.3026728
  63. 63. Wang G, Endicott L, Uher C. Recent advances in the growth of Bi–Sb–Te–Se thin films. Sci. Adv. Mater. 2011; 3: 539–560. DOI: 10.1166/sam.2011.1182
  64. 64. Cahill DG. Thermal conductivity measurement from 30 to 750 K: the 3ω method. Rev. Sci. Instrum. 1990; 61: 802. DOI: 10.1063/1.1141498.
  65. 65. Takashiri M, Miyazaki K, Tanaka S, Kurosaki J, Nagai D, Tsukamoto H. Effect of grain size on thermoelectric properties of n-type nanocrystalline bismuth-telluride based thin films. J. Appl. Phys. 2008; 104: 084302. DOI: 10.1063/1.2990774.
  66. 66. Takashiri M, Tanaka S, Miyazaki K, Tsukamoto H. Cross-plane thermal conductivity of highly oriented nanocrystalline bismuth antimony telluride thin films. J. Alloys Compd. 2010; 490: L44–L47. DOI: 10.1016/j.jallcom.2009.10.117
  67. 67. Venkatasubramanian R. Lattice thermal conductivity reduction and phonon localizationlike behavior in superlattice structures. Phys. Rev. B. 2000; 61: 3091–3097. DOI: 10.1103/PhysRevB.61.3091
  68. 68. Peranio N, Eibl O, Nurnus J. Structural and thermoelectric properties of epitaxially grown Bi2Te3 thin films and superlattices. J. Appl. Phys. 2006; 100: 114306. DOI: 10.1063/1.2375016.
  69. 69. N. F. Hinsche, B. Yu. Yavorsky, M. Gradhand, M. Czerner, M. Winkler, J. König, H. Böttner, I. Mertig, and P. Zahn. Thermoelectric Properties of Bi2Te3/ Sb2Te3 Superlattices. Phys. Rev. B. 2012; 86: 085323. DOI: 10.1103/PhysRevB.86.085323
  70. 70. Walachová J, Zeipl R, Zelinka J, Malina V, Pavelka M, Jelínek M, Studnička V, Lošt’ák P. High room-temperature figure of merit of thin layers prepared by laser ablation from Bi2Te3 target. Appl. Phys. Lett. 2005; 87: 081902. DOI: 10.1063/1.2001755.
  71. 71. Chang H-C., Chen C-H., Kuo Y-K. Great enhancements in the thermoelectric power factor of BiSbTe nanostructured films with well-ordered interfaces. Nanoscale. 2013; 5: 7017–7025. DOI: 10.1039/c3nr01499a

Written By

Le Thi Cam Tuyen, Phuoc Huu Le and Sheng-Rui Jian

Submitted: 06 February 2021 Reviewed: 15 July 2021 Published: 09 August 2021