Open access peer-reviewed chapter

Recent Advances in BiVO4- and Bi2Te3-Based Materials for High Efficiency-Energy Applications

Written By

Phuoc Huu Le, Nguyen Trung Kien and Chien Nguyen Van

Submitted: 22 December 2017 Reviewed: 19 February 2018 Published: 20 June 2018

DOI: 10.5772/intechopen.75613

From the Edited Volume

Bismuth - Advanced Applications and Defects Characterization

Edited by Ying Zhou, Fan Dong and Shengming Jin

Chapter metrics overview

1,894 Chapter Downloads

View Full Metrics

Abstract

This chapter provides recent progress in developments of BiVO4- and Bi2Te3-based materials for high efficiency photoelectrodes and thermoelectric applications. The self-assembling nanostructured BiVO4-based materials and their heterostructures (e.g., WO3/BiVO4) are developed and studied toward high efficiency photoelectrochemical (PEC) water splitting via engineering the crystal and band structures and charge transfer processes across the heteroconjunctions. In addition, crystal and electronic structures, optical properties, and strategies to enhance photoelectrochemical properties of BiVO4 are presented. The nanocrystalline, nanostructured Bi2Te3-based thin films with controlled structure, and morphology for enhanced thermoelectric properties are also reported and discussed in details. We demonstrate that BiVO4-based materials and Bi2Te3-based thin films play significant roles for the developing renewable energy.

Keywords

  • BiVO4
  • bismuth telluride (Bi2Te3)
  • photoelectrochemical (PEC) water splitting
  • thermoelectrics
  • pulsed laser deposition

1. Bismuth vanadate (BiVO4) material: a highly promising photoanode for use in solar water oxidation

The photoelectrochemical (PEC) water splitting, which uses semiconductors to directly harvest and convert the abundant solar energy into storable and friendly environmental energy in the form of hydrogen and oxygen from water, has been regarded as a promising approach to solve our current energy challenges [1, 2, 3, 4, 5, 6]. The PEC cell for energy conversion on a global scale requires the development of devices that are highly efficient, stable, and simple in design. The effective and economic energy transfer technologies are still under developing stage for more than 40 years. The PEC water splitting becomes more competitive as the cost continues to decrease with the development of nanotechnology, which offers the use of new nano-heterostructure photocatalysts. The current direction along this research field focuses on the development of high efficiency photoelectrodes based on metal oxides due to their high chemical and photostability, low cost, and ease of fabrication. Among these interesting and important photoelectrode materials for PEC, BiVO4 is demonstrated as a promising photocatalyst for PEC water splitting. Originally, the BiVO4was intensively studied as a ferroelasticity, acoustic-optical, and ionic conductivity materials [7]; however, recently most studies focus on its photocatalytic activity owing to its direct bandgap of ˜2.5 eV together with the desirable band edges and good stability [8, 9, 10, 11]. Additionally, the theoretical solar-to-hydrogen conversion efficiency of the monoclinic BiVO4 could reach 9.2% with a maximum photocurrent of 7.5 mA cm−2 under standard AM 1.5 solar light irradiation, which is significantly higher than those of the common metal oxides, and thus the monoclinic BiVO4 has been recognized as one of the most promising photoanode materials for PEC water splitting [1]. However, its photoactivity performance is limited by its poor electrical conductivity, slow hole transfer kinetics for water oxidation, and poor charge separation [12]. A concise overview of BiVO4 electronic and optical properties as well as the alternative strategies to improve its photocatalytic activities is discussed in this section.

Advertisement

2. Crystal and electronic structures of bismuth vanadate

The synthesis BiVO4 crystallizes in three main crystal structures of tetragonal scheelite, monoclinic phase, and tetragonal zircon structure. The photocatalytic activities of BiVO4 strongly depend on the optical properties of each crystal structure. Among these structures, the monoclinic scheelite BiVO4 was demonstrated to exhibit the highest photocatalytic water oxidation under visible light illumination [13]. The higher photocatalytic activity of the monoclinic scheelite was attributed to its smaller energy band gap and higher crystal distortion, which enhance the optical absorption and the electron-hole separation [14]. The conduction-band minimum of BiVO4 is formed mainly by V 3d, which splits into triplet bands V 3d x2−y2/z2, V 3d z2/zx, and V 3d zy/xy [15]. The valence band (VB) of tetragonal zircon structure BiVO4 is formed by O 2p orbital, but the top region of VB of monoclinic structure is formed by both Bi 6 s and O 2p orbitals. It was demonstrated that Bi 6 s is located above O 2p and the smaller bandgap of monoclinic BiVO4 attributes to the transition from the 6 s electrons of Bi to the 3d orbital of V [14, 16]. Recently, Zou et al. demonstrated that the crystal distortion in monoclinic BiVO4 improves the lone-pair impact of Bi 6 s states which raises up the O 2p states and reduces the bandgap [17]. In addition, the monoclinic scheelite structure possesses the better electron-hole separation owing to its higher degree of structural distortion than that in the tetragonal scheelite structure to result in enhanced photocatalytic activity for the monoclinic structure. In addition, the conduction-band edge of BiVO4 is only slightly below the H+/H2 redox potential, while its valence band is much below the oxidation potential of water resulting in low required external bias [15].

In order to investigate the experimental crystal and electronic structure of monoclinic BiVO4, the single crystal BiVO4, which possesses good crystallinity and eliminates the structural and electronic defects as well as impurity phases, is required. The vacuum deposition such as sputtering, molecular beam epitaxy, and pulsed laser deposition for the single crystal monoclinic BiVO4 has been developed recently [18, 19, 20]. A good lattice match (mismatch less than 1%) of cubic yttrium-stabilized zirconia (YSZ, a = 5.145 Å) was used as substrate for epitaxial growth of high-quality epitaxial monoclinic BiVO4 (a = 5.1935 Å, b = 5.0898 Å, c = 11.6972, γ = 90.3871) films. The X-ray diffraction θ–2θ scan (Figure 1a) shows only high intensity and sharp (00 l) reflections of monoclinic BiVO4 phase indicating that growth was c-axis oriented: BiVO4 (001) || YSZ (001). The high-resolution TEM image (the inset of Figure 1) reveals the in-plane and out-of-plane lattice parameters of the BiVO4 with d200 = 2.53 Å and d002 = 5.81 Å which are in good agreement with the lattice constants of the bulk monoclinic BiVO4. The orientation and growth direction of BiVO4 crystal were identified and constructed in the corresponding atomic model as shown in Figure 1b.

Figure 1.

(a) XRD 2θ/θ scans of BiVO4 film grown on YSZ substrate, and the inset shows HR-TEM image of BiVO4; (b) atomic model of BiVO4/ YSZ heteroconjunction.

The electronic structural properties of single crystal BiVO4 were investigated by the photoluminescence (PL) spectroscopy and UV-Visible spectroscopy. Figure 2 presents the PL and UV-Vis absorption spectra of a single crystal BiVO4 film. As presented in the UV-Vis absorption spectrum, the observed absorption in the range of 440–480 nm corresponding to the bandgap of ˜ 2.67 eV could attribute to the transition from the 6 s electrons of Bi (or the hybrid orbital of Bi 6 s and O 2p) to the empty 3d x2-y2/z2 orbital of V, which is consistent with emission peak of 2.65 eV in the PL spectrum. The second absorption band edge of approximately 3.15 eV and the second PL emission peak at 3.1 eV could be ascribed to the electron transition from O 2p valence band to V 3d z2/zx and/ or V 3d zy/xy orbitals.

Figure 2.

Photoluminescence taken under the irradiation of 325 nm laser at room temperature (green curve) and UV-Vis absorption spectrum (red curve) of BiVO4 epitaxial film.

Advertisement

3. Strategies to enhance photoelectrochemical properties of BiVO4

The development of suitable techniques to fabricate high efficiency BiVO4 photoelectrode for PEC water spitting is crucial. To date, there are many methods, which have been developed to prepare BiVO4 photoelectrodes for use in solar water oxidation. The representative synthesis methods are (1) metal organic decomposition (MOD) combining with the spin coating or spray pyrolysis deposition [21, 22, 23], (2) electrophoretic deposition (EPD) and chemical bath deposition (CBD) [24, 25, 26, 27], and (3) vacuum deposition methods such as sputtering deposition and pulsed laser deposition (PLD) [18, 19, 20, 28, 29]. The MOD is a facile method to synthesize BiVO4 materials for photoelectrodes. The main advantages of the MOD method are the easy composition tuning and morphology engineering. However, the existence of crystal and surface defect and low adhesion between BiVO4 and conducting substrate could limit its performance. The EPD and CBD methods are simple, have low cost, and are easy to scale up. The vacuum deposition is a powerful method to synthesize good crystalline BiVO4 photoelectrodes. The samples prepared by this method could easily control the doped composition; could facilely tune the crystalline in form of amorphous, polycrystalline, and single crystalline; and possess a good adhesion.

As described above, BiVO4 has many advantages features, however, the actual PEC performance of undoped BiVO4 is still far below its theoretical value, indicating that its advantages have not been fully developed. Many research groups have improved photoelectrochemical water oxidation by the introduction of (1) controlled morphologies, (2) forming nanocomposite structure, (3) doping, and (4) decorating with cocatalysts [30]. A summary of advantages and alternative strategies to enhance photoelectrochemical properties of BiVO4 was depicted in Figure 3. The best PEC performance of BiVO4 photoelectrodes that has been demonstrated to date is a combination of several above strategies to take advantages of high surface area, high electron mobility, high water oxidation kinetics, and low charge carrier recombination. Therefore, understanding the benefits and disadvantages of each strategy could provide an effective way to enhance the PEC properties.

Figure 3.

The advantages and strategies to enhance photoelectrochemical properties of BiVO4.

First, the morphologies of BiVO4 photoelectrode such as shape, size, and particle contact strongly effect on their interfacial energetics, kinetics, and charge transport properties as well as reactive sites. Because the electrochemical reactions only occur at the electrode-electrolyte interfacial area, the photogenerated electron-hole needs to be transferred to the surface for the reactions. Thus, the development of high surface area photoelectrodes is necessary. The porous structure and one-dimensional nanostructure such as nanorod/nanowire arrays reveal high surface areas per electrode volume, which directly improves the PEC efficiency. The high efficient PEC devices based on the BiVO4 porous structure and the one-dimensional nanostructures have been reported by many research groups [31, 32, 33, 34]. However, creating more porosity or smaller nanorod/nanowire structures to improve surface area could increase the defect site, increase grain boundaries, and reduce crystallinity leading reduction photocatalytic activity [35]. In addition, with hole diffusion length of ∼100 nm and carrier mobility of 0.044 cm2 V−2 s−1 [36], the length and thickness of nanostructure should be finely optimized to maximum light absorption and photogenerated charge carrier transport. Furthermore, researchers have recently demonstrated that the charge separation and photoactivity properties are closely related to the exposed crystal facet of BiVO4 photocatalysts [8, 9, 37, 38, 39]. Therefore, an optimized design and morphological control of crystal facets could improve the PEC performance.

Second, nanocomposites provide a powerful route to overcome limitations in the current studies of single material systems for water splitting, where the photoelectrochemical performance of photoelectrode can be significantly improved by the choice of proper interactions of constituents. The construction of nanocomposite consisting of BiVO4 and addition semiconductor/or conductor can improve the optical absorption, charge carrier separation, and charge transport processes. Combination of BiVO4 with additional metal oxide semiconductors could enhance overall photon absorption and improve electron-hole separation by the rapid photogenerated electron and hole injections at their heterojunctions and extension the optical absorption range. Many BiVO4-based heterostructures have been fabricated successfully such as TiO2/BiVO4, Ag3PO4/BiVO4, Bi2S3/BiVO4, g-C3N4/BiVO4, BiVO4/SnO2, WO3/BiVO4, BiOCl/BiVO4, Fe2O3/BiVO4, Cu2O/BiVO4, and ZnO/BiVO4 [40, 41, 42, 43]. Among them, the WO3 is a commonly coupled semiconductor with BiVO4 because it is a stable n-type semiconductor with better electron transport properties than the BiVO4 and a suitable conduction-band edge for facile electron injection from conduction band of BiVO4 [29, 44, 45, 46].

Recently, we reported a self-assembled nanocomposite photoanode composed of epitaxial BiVO4 matrix embedded with WO3 mesocrystal for photoelectrochemical application in the visible light regime using PLD [29]. By taking the advantage of the structural feature of this heterostructure, the well-defined crystal facet and interface between WO3 and BiVO4 phases provide a template for the fundamental understanding of photoactivity in the nanocomposite. The BiVO4-WO3 crystal structure details were investigated by XRD and TEM as shown in Figure 4. The XRD 2θ-θ scans, in-plane Φ-scans, and reciprocal space map results clearly show that BiVO4 and WO3 spontaneously separated into two single phases during the deposition with an in-plane orientation relationship as [100]BiVO4 //[110]WO3 //[001]YSZ. Both monoclinic BiVO4 and orthorhombic WO3 phases exhibit two sets of structural domains, which are separated by a 90° rotation around c-axis. The microstructure and the epitaxial relationship of BiVO4-WO3 heterostructure were further investigated as displayed in Figure 4(b–e), which are in an excellent agreement with the XRD analyses. The self-assembled mesocrystal-embedded heterostructure that composed WO3 and BiVO4 phases is illustrated in Figure 4g. The energy band alignment of BiVO4-WO3 heterojunction illustrated using X-ray photoelectron spectroscopy technique shows that the valence band maxima of WO3 lies 0.55 eV below that of BiVO4 and the conduction-band minimum of WO3 lies 0.25 eV below that of BiVO4. This energy band structure could favor the enhancement of photogenerated charge carrier separation, which was further confirmed by photoluminescence and ultrafast transient absorption spectroscopies. Figure 5 shows Nyquist electrochemical impedance spectroscopy and Mott-Schottky analysis for the BiVO4-WO3 photoelectrodes. The results demonstrated a much higher electron conductivity of WO3 with carrier densities of 9.68 × 1018 cm−3 under light illumination than BiVO4. Therefore, the BiVO4 matrix serves as the light absorber due to its narrow bandgap, and the WO3 mesocrystal acts as an electron conductor owing to its high electron conductivity resulting in a significantly enhanced photoelectrochemical performance.

Figure 4.

Structure characterization of BiVO4-WO3 nanocomposite. (a) X-ray diffraction 2θ-θ scans of pure WO3, BiVO4, and the BiVO4-WO3 nanocomposite showing only (00 l) type peaks of BiVO4 and WO3. Inset shows XRD ϕ-scans of YSZ (220), BiVO4 {013}, and WO3 {022} reflections for BiVO4-WO3 heterostructure. (b-e) Cross-sectional TEM images of BiVO4-WO3 (BVO-WO) nanocomposite taken along [010]YSZ zone axis. The enlarged images taken from the marked areas of (d) BiVO4 and (e) WO3 showing their out-of-plane and in-plane d-spacing parameters. (f) X-ray reciprocal space maps (RSMs) of the composite grown on ITO/YSZ. (g) Illustration of self-assembled BiVO4-WO3 nanocomposite [29].

Figure 5.

(a) The Nyquist plots and (b) the Mott-Schottky plots for the WO3, BiVO4, and BiVO4-WO3 photoelectrodes in 0.5 M Na2SO4 solution under AM 1.5 illumination [29].

Further example could be observed in the double-deck inverse opal WO3/BiVO4 structure that could reach a photocurrent density of 3.3 mA/cm2 at 1.23 V versus RHE and the core-shell WO3/BiVO4 helix nanostructure, in which BiVO4 was doped with Mo and naturally doped with W from the WO3 core, exhibits a high photocurrent density of 3.6 mA/cm2 at 1.23 V versus RHE [34, 47]. More recently, the highest photocurrent density of 6.72 mA/cm2 at 1.23 V versus RHE for BiVO4-based photoelectrode was obtained in the CoPi-coated WO3/BiVO4 core-shell nanorods [21]. The high efficiency photoelectrode was attributed to the superior charge carrier separation and efficient light scattering. In addition, by coupling with a conductor such as reduced graphene oxide and carbon nanotubes, the photocurrent density of the composite electrodes was improved significantly compared to those of the corresponding BiVO4 electrodes which could attribute to the enhancement of charge separation and transport properties [49, 50, 51, 52, 53, 54, 55, 56, 57]. However, due to the low crystallinity of BiVO4 as well as the poor BiVO4-conductor and conductor-substrate contact in these nanocomposites, it needs to be further developed to improve the PEC performance.

Third, doping with the metal (such as W, Fe, B, Cu, Zn, Ti, Nb, Sn, Co, Pd, Rb, Ru, Ag, Ga, Sr, and Ir) or nonmetal (such as P and N) materials can change the electrical and optical properties of BiVO4. Among them, only Mo- and W-doped BiVO4 show a significantly enhanced photocurrent and IPCE [21]. The intentional introduction of n-type conductivity W and/or Mo atoms does not change the bandgap energy and band edge positions of doped-BiVO4; however, it can increase the charge carrier density and the electrical conductivity. A suitable doping concentration of Mo and/or W results in significant improvement PEC performance of BiVO4 [54, 55, 56]. On the other hand, P doping forms an internal electric field which may improve the photogenerated electron–hole separation. However, it is noted that the doping could induce trap carrier which can enhance charge carrier recombination and reduce the charge transport [35]. Therefore, the optimum doping concentration is a key factor to improve the overall PEC performance of the photoelectrode.

Finally, decoration of BiVO4 with various oxygen evolution catalysts (OECs) such as Co–Pi, Co3O4, RhO2, Pt, CoO, and FeOOH improves the kinetic for oxygen evolution and provides unique active sites for catalytic reactions, thereby strongly enhanced the photocurrent density [3, 27, 40, 57, 58]. In addition, it decreases the bias potential and improves the stability of BiVO4 photoanodes. Perhaps the most successful route in the development of high efficient BiVO4 photoelectrode for water oxidation is the modification of the BiVO4-semiconductor nanocomposite surface with OECs. To date, the highest photocurrent density 6.72 mA/cm2 at 1.23 V versus RHE was obtained for the CoPi-coated WO3/BiVO4 core-shell nanorods [48].

Advertisement

4. Introduction to thermoelectric bismuth-based materials and applications

Thermoelectric (TE) materials are used in appliances such as heat pumps and power generators [59, 60, 61, 62, 63]. TE devices offer unique features of low environmental impact, no moving parts, quiet operation, and high reliability. The performance of TE materials is determined by a dimensionless figure of merit, ZT = α2σT/κ, in which α, σ, T, and κ are the Seebeck coefficient, the electrical conductivity, absolute temperature, and the thermal conductivity, respectively. A TE material must exhibit a high power factor (PF = α2σ) and low thermal conductivity (κ) to achieve a high ZT value. However, it is challenging for enhancing the ZT value due to the coupling among the TE parameters [62]: the relationship between α and the carrier concentration n (expressed by |α|˜ n−2/3 approximately [60]) limits the increase of the PF (=α2σ), while the proportional relationship between electrical conductivity and electronic thermal conductivity (the Wiedemann-Franz law) restricts the enhancement of the σ/κ ratio.

Bismuth (Bi)-based chalcogenide narrow-bandgap semiconductors such as Bi2Te3, Bi2Se3, Sb2Te3, Bi2(TexSe1-x)3, and (BixSb1-x)2Te3 are of high interest [64, 65, 66, 67, 68, 69]. They have been widely exploited for Peltier coolers and thermoelectric generators at low temperature regime (≤150°C) [70, 71]. Thanks to the extensive phonon scattering at grain boundaries, nanocrystalline and nanostructured Bi-based chalcogenide thin films have achieved reduced thermal conductivity [72, 73, 74, 75, 76, 77]. However, the lattice imperfections and grain-boundary defects impair the electrical transport properties of the films [72], which call for further investigations to improve PF or the electronic part of ZT. Currently, enhancing the PF of Bi-based thin films is challenging due to not only the coupling among TE material properties [60] but also the tendency of forming nonstoichiometric films at elevated substrate temperatures (Ts) [78]. Numerous charge carriers arising from vacancy defects of volatile elements can constrain the enhancement of ׀α׀; however, low carrier concentrations can suppress electrical conductivity if carrier mobility (μ) is poor.

The crystal structures of Bi2Se3 and Bi2Te3 are usually described by a hexagonal cell that consists of 15 layers of atoms stacking along the c-axis with a sequence shown below [60], as shown in Figure 6. ··· Se(1)–Bi–Se(2)–Bi–Se(1) ··· Se(1)–Bi–Se(2)–Bi–Se(1) ··· Se(1)–Bi–Se(2)–Bi–Se(1) ··· and ···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 Se or Te atoms. The five-atomic-layer thick lamellae of –(Se(1)–Bi–Se(2)–Bi–Se(1))– or –(Te(1)–Bi–Te(2)–Bi–Te(1))– is called quintuple layers, QLs. The Se(1)…Se(1) or Te(1)…Te(1) refers to van der Waals force between Se and Te atoms, whereas the Se(1)–Bi and Bi–Se(2) or Te(1)–Bi and Bi–Te(2) are ionic-covalent bonds. This weak binding between the Se(1) – Se(1) and Te(1)…Te(1) accounts for the easy cleavage of these materials perpendicular to the c-axis and the anisotropic thermal and electrical transport properties. For example, the thermal conductivity along the c-axis direction is ˜0.7 Wm−1K −1, while it is ˜1.5 Wm−1K −1 for the plane perpendicular to the c-axis [79].

Figure 6.

The hexagonal structures of one unit cell of Bi2Se3 and Bi2Te3.

Thin-film technology is advantageous for obtaining nanocrystalline and nanostructured materials by adjusting deposition conditions and subsequent thermal treatments. The extensive phonon scattering at grain boundaries in the nanostructures causes a large reduction in thermal conductivity while maintaining reasonable electrical conductivity, leading to enhanced ZT. Among physical vapor deposition techniques, PLD offers a great versatility in the fabrication of films with multielement stoichiometry and with a variety of structures, from amorphous or nanostructured to polycrystalline or even epitaxial [66, 67, 68, 76, 77, 80, 81]. Thin-film TE devices offer some distinctive advantages.

First, a thin-film device has the natural advantage of a small volume (thickness of ≤10 μm, length and width of ˜100 μm) as compared to size of millimeters for a bulk TE couple. As a result, thin-film TE cooling can be integrated into microelectronic systems (Figure 7a). In principle, the bulk device can be scaled down to micro sizes; however, fabrication processes to do so are difficult.

Figure 7.

(a) An example for optoelectronics of the continuing reduction in package size. (b) A schematic of grain-boundary scattering for thin-film materials with (i) disoriented-small grains and (ii) highly oriented large grains. (c) The comparison of thermal/cooling time response of thin-film (˜5 μm) superlattice device and a bulk device [59]. (d) Thermoelectric module performance chart which presents the temperature drop ΔT versus pumping power [83]. (e) A commercial thin-film TE module. (f) Size comparison between a thin-film TE device and bulk TE device [83].

Second, thin-film devices have a much shorter response time than bulk devices, as shown in Figure 7c [59]. The thin-film device achieves the steady state in 15 μs, while the bulk device requires 0.35 s [59]. This is a result of the response time associated with the transport of heat through the thin film (micrometers) rather than through the millimeters associated with bulk devices.

Third, the thin-film device has the ability to handle much larger density of heat pumping power than does the bulk device. Typically, bulk devices are working with pumping power density lower than 10 W/cm2; however, the pumping power density in a thin-film device is on the order of hundreds of W/cm2 [82]. Figure 7d illustrates a comparison performance chart between thin-film and bulk TE modules, in which the load line represents the temperature difference (ΔT) between the top and bottom substrate and possible pumping power density (Q) for a given drive current of a TE module [83]. At the maximum drive current for a TE module, the load line is determined by the maximum power density (Qmax) and the ΔTmax. For a commercial example, under a given drive current, an eTEC™ Series thin-film TE module (Laird technology) possesses a larger pumping power up to ten times (Figure 7e) and a comparable maximum temperature difference ΔTmax than the conventional bulk TE modules [83].

Fourth, the granular structure-morphology of thin films increases the grain-boundary scattering which is helpful for suppressing κ and enhanced ZT values. Figure 7b represents the roles of grain orientation and grain size in carrier mobility. The grains with the same colors illustrate that they have the same/similar orientations. The grain-boundary scattering in Figure 7b(i) will be greater than that of the Figure 7b(ii) case because of its smaller grain size and the greater grain disorientation. The effective mobility is given by [84]

µg=Lq2πmkTexpφbkTE1

where q is the carrier charge, m* is the effective mass, k is the Boltzmann constant, T is the temperature, and φb is the grain-boundary potential barrier in the depletion region. In polycrystalline silicon, the potential barrier height is approximately twice as high at random boundaries as at low-energy coincidence boundaries.

Advertisement

5. Thermoelectric properties of nanocrystalline and nanostructured Bi2Te3-based thin films

Thermoelectric Bi2Te3-based materials have played a dominant role in the field of thermoelectrics. The traditional cooling materials are 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 at room temperature approximately equal to one [63]. Recently, nanocrystalline and nanostructured Bi2Te3-based thin films have recently attracted great interests because of their superior TE performance [72, 73, 74, 75, 76, 77, 80, 81, 85, 86]. The PF of Bi2Te3 films was 8.8 μWcm−1 K−2 for an (015)-oriented film with layered-hexagonal morphology [85] and 33.7 μWcm−1 K−2 for a highly (00l)-oriented layered film [86]. Furthermore, the hexagonal-Bi2Te3 film grown using MBE obtained a PF of 27 μWcm−1 K−2 [87], and the smooth-epitaxial- Bi2Te3 film prepared by co-evaporation achieved a PF of 39.9 μWcm−1 K−2 [88].

Despite fewer studies performed on thermoelectric Bi2Se3 as compared to Bi2Te3, a considerable amount of recent efforts to enhance TE performance has been devoted to the synthesis of Bi2Se3 nanostructures such as hexagonal flakes (PF ≈ 0.28 μWcm−1 K−2) [89] and nanoflakes (PF ≈ 0.97 μWcm−1 K−2) [90], using the solvo-thermal method and chemical bath deposition, respectively. Moreover, the Bi2Se3 film grown by metal organic-chemical vapor deposition has reached a PF = 5.8 μWcm−1 K−2 [91]. In PLD, tightly controlling substrate temperatures (Ts) and ambient pressures enable the morphologies and compositions of films to be manipulated extensively, which offer a new method for enhancing the TE properties of films. For example, self-assembled Bi2Te3 films featuring well-aligned zero- to three-dimensional nanoblocks have been fabricated (Figure 8), but the room temperature PFs of these films remain low (≤ 1.9 μWcm−1 K−2) due to the low electrical conductivity of the voided structures [77].

Figure 8.

The morphology and power factor (PF, unit μWcm−1 K−2) of nanostructured Bi2Te3 thin films grown by PLD at various substrate temperatures and ambient pressures, reported by Chang and Chen [77].

By contrast, the Bi2Te3-based thin films with compact and/or layered structures possess the high electrical conductivities and relatively high Seebeck values if the films obtain stoichiometry. For example, A. Li Bassi et al. reported a high PF of 50.6 μWcm−1 K−2 for the layered-smooth Bi2Te3 films (Figure 9a) [80]. Similarly, the Bi2Te3 films with layered-hexagonal grain structure exhibited a high PF of 33.7 μWcm−1 K−2 (Figure 9b) [86]. P.H. Le et al. found that the optimal PF films also present compact-layered structures, namely, PF = 24.3 μWcm−1 K−2 for the Bi2Te3 film (Figure 9c) and PF = 5.54 μWcm−1 K−2 for the hexagonal Bi2Se3 film (Figure 9d) [66, 67]. Usually, compact or layered structure films obtain high carrier mobility than those with the open/voided nanostructures.

Figure 9.

The morphology and power factor (PF, unit μWcm−1 K−2) of the optimal Bi2Te3 and Bi2Se3 thin films with layered structures grown by PLD, reported in Refs. [66, 67, 80, 86].

Table 1 summarizes thermal transport properties (at room temperature) of nanocrystalline-nanostructured Bi2Te3-based thin films and bulk materials in the literature [62, 72, 73, 74, 75, 76, 87, 92]. Generally, the thermal conductivity κ value for polycrystalline films is expected to be smaller than that of bulk alloys because of the contribution of grain-boundary scattering [62, 72, 74]. Moreover, the κ of nanocrystalline Bi2Te3-based films will further decrease when the grain size of decreases (κ ≤ 0.81 W/mK) [73, 75]. 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 [93].

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[72]
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[73]
27 nm0.68540−138.110.30.46
60 nm0.80540−186.118.70.70
Nanocrystalline Bi-Sb-Te thin film, sputtering26 nm0.463.3[75]
45 nm0.656.7
84 nm0.8133.3
Nanocrystalline BiSbTe (8:30:62) thin film, flash evaporation150 nm0.6[74]
Single crystal BiSbTe bulk alloys0.75[62]
Bi2Te3/Sb2Te3 superlattices
(period ˜5 nm)
0.4[93]
Bi2Te3 + 0.63 bulk2.21000−240580.87[62]
Bi2(Te0.95Se0.05)3 bulk1.59901−223450.85[62]
Bi2Te3/Bi2(Te0.88Se0.12)3 superlattice film, MBE80 nm1.25639−204270.60[87]
Bi2Te3 films, laser ablation0.2–0.3[92]
BixSb2-xTe3 nanolayer film, PLD190 nm1.16270095250.65[76]
BixSb2-xTe3 nanodisk film, PLD100 nm1.001100132200.60
BixSb2-xTe3 nanocolumn film, PLD70 nm0.93280207120.39

Table 1.

Room temperature thermal transport properties of nanocrystalline-nanostructured Bi2Te3-based thin films and bulk materials in the literature, including 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).

For PLD Bi2Te3-based films, Walachova et al. [92] estimated κ starting from direct ZT measurement with the Harman method and found a value of about 0.2–0.3 W/mK for films with a thickness comparable to our films (>100 nm). Recently, Chang et al. [76] reported the κ between 0.93 and 1.16 W/mK for granular BixSb2-xTe3 films. The reported κ (at 300 K) of hexagonal flake Bi2Se3 was 0.75 W/mK.

Advertisement

6. Conclusion

We report the crystal, electronic structures, optical properties, and PEC activities of BiVO4-based materials. Morphology engineering, cocatalyst loading, and doping with metal and nonmetal composite structures are the approaches to fabricate high efficiency BiVO4 photoelectrode for PEC water spitting. In addition, the Bi2Te3-based thin films with nanocrystalline and nanostructured morphologies obtain the reduced thermal conductivity, meanwhile the close stoichiometric films with layered structure morphologies exhibit high thermoelectric power factors. BiVO4-based materials and Bi2Te3-based thin films play significant roles for the photoelectrode and thermoelectric applications.

Advertisement

Acknowledgments

Financial support from Vietnam National Foundation for Science and Technology Development (NAFOSTED) under grant numbers 103.99–2015.17, 103.99-2016.75 (PHL), and MOST106-2119-M-009-011-MY3, Taiwan (CNV), is gratefully acknowledged.

References

  1. 1. Li R, Han H, Zhang F, Wang D, Li C. Highly efficient photocatalysts constructed by rational assembly of dual-cocatalysts separately on different facets of BiVO4. Energy & Environmental Science. 2014;7:1369. DOI: 10.1039/c3ee43304h
  2. 2. Zhao X, Luo W, Feng J, Li M, Li Z, Yu T, Zou Z. Quantitative analysis and visualized evidence for high charge separation efficiency in a solid-liquid bulk heterojunction. Advanced Energy Materials. 2014;4:1301785. DOI: 10.1002/aenm.201301785
  3. 3. Kim TW, Choi K-S. Nanoporous BiVO4 photoanodes with dual-layer oxygen evolution catalysts for solar water splitting. Science. 2014;343:990-995. DOI: 10.1126/science.1246913
  4. 4. Prévot MS, Sivula K. Photoelectrochemical tandem cells for solar water splitting. Journal of Physical Chemistry C. 2013;117:17879-17893. DOI: 10.1021/jp405291g
  5. 5. Grätzel M. Photoelectrochemical cells. Nature. 2001;414:338-344. DOI: 10.1016/S0140-6736(06)68542-5
  6. 6. Walter MG, Warren EL, James R. McKone JR, Boettcher SW, Mi Q, Santori EA, Lewis NS. Solar water splitting cells. Chemical Reviews. 2010;110:6446-6473. DOI: 10.1021/cr1002326
  7. 7. Tan HL, Amal R. Alternative strategies in improving the photocatalytic and photoelectrochemical activities of visible light-driven BiVO4: A review. Journal of Materials Chemistry A. 2017;5:16498-16521. DOI: 10.1039/C7TA04441K
  8. 8. Xi G, Ye J. Synthesis of bismuth vanadate nanoplates with exposed {001} facets and enhanced visible-light photocatalytic properties. Chemical Communications. 2010;46:1893-1895. DOI: 10.1039/b923435g
  9. 9. Li R, Han H, Zhang F, Wang D, Li C. Highly efficient photocatalysts constructed by rational assembly of dual-cocatalysts separately on different facets of BiVO4. Energy & Environmental Science. 2014;7:1369. DOI: 10.1039/c3ee43304h
  10. 10. Yu J, Kudo A. Effects of structural variation on the photocatalytic performance of hydrothermally synthesized BiVO4. Advanced Functional Materials. 2006;16:2163-2169. DOI: 10.1002/adfm.200500799
  11. 11. Huang Z-F, Pan L, Zou J-J, Zhang X, Wang L. Nanostructured bismuth vanadate-based materials for solar-energy-driven water oxidation: A review on recent progress. Nanoscale. 2014;(23):14044-14063. DOI: 10.1039/c4nr05245e
  12. 12. Liang Y, Tsubota T, Mooij LPA, Krol R. Highly improved quantum efficiencies for thin film BiVO4 photoanodes. Journal of Physical Chemistry C. 2011;115(35):17594-17598. DOI: 10.1021/jp203004v
  13. 13. Tokunaga S, Kato H, Kudo A. Selective preparation of monoclinic and tetragonal BiVO4 with Scheelite structure and their photocatalytic properties. Chemistry of Materials. 2001;13:4624-4628. DOI: 10.1021/cm0103390
  14. 14. Kudo A, Omori K, Kato H. A novel aqueous process for preparation of crystal form-controlled and highly crystalline BiVO4 powder from layered vanadates at room temperature and its photocatalytic and photophysical properties. Journal of the American Chemical Society. 1999;121:11459-11467. DOI: 10.1021/ja992541y
  15. 15. Cooper JK, Gul S, Toma FM, Chen L, Glans PA, Guo J, Ager JW, Yano J, Sharp ID. Electronic structure of monoclinic BiVO4. Chemistry of Materials. 2014;26(18):5365-5373. DOI: 10.1021/cm5025074
  16. 16. Walsh A, Yan Y, Huda MN, Al-Jassim MM, Wei S-H. Band edge electronic structure of BiVO4: Elucidating the role of the Bi s and V d orbitals. Chemistry of Materials. 2009;21(3):547-551. DOI: 10.1021/cm802894z
  17. 17. Zhao Z, Li Z, Zou Z. Electronic structure and optical properties of monoclinic clinobisvanite BiVO4. Physical Chemistry Chemical Physics. 2011;13:4746-4753. DOI: 10.1039/c0cp01871f
  18. 18. Van CN, Chang WS, Chen J-W, Tsai K-A, Tzeng W-Y, Lin Y-C, Kuo H-H, Liu H-J, Chang K-D, Chou W-C, Wu C-L, Chen Y-C, Luo CW, Hsu Y-J, Chu Y-H. Heteroepitaxial approach to explore charge dynamics across Au/ BiVO4 interface for photoactivity enhancement. Nano Energy. 2015;15:625-633. DOI: 10.1016/j.nanoen.2015.05.024
  19. 19. Rettie AJE, Mozaffari S, Mcdaniel MD, Pearson KN, Ekerdt JG, Markert JT, Mullins CB. Pulsed laser deposition of epitaxial and polycrystalline bismuth vanadate thin films. Journal of Physical Chemistry C. 2014;118(46):26543-26550. DOI: 10.1021/jp5082824
  20. 20. Stoughton S, Showak M, Mao Q, Koirala P, Hillsberry DA, Sallis S, Kourkoutis LF, Nguyen K, Piper LFJ, Tenne DA, Podraza NJ, Muller DA, Adamo C, Schlom DG. Adsorption-controlled growth of BiVO4 by molecular-beam epitaxy. APL Materials. 2013;1:042112. DOI: 10.1063/1.4824041
  21. 21. Luo W, Yang Z, Li Z, Zhang J, Liu J, Zhao Z, Wang Z, Yan S, Yu T, Zou Z. Solar hydrogen generation from seawater with a modified BiVO4 photoanode. Energy & Environmental Science. 2011;4:4046-4051. DOI: 10.1039/c1ee01812d
  22. 22. Sayama K, Nomura A, Arai T, Sugita T, Abe R, Yanagida M, Oi T, Iwasaki Y, Abe Y, Sugihara H. Photoelectrochemical decomposition of water into H2 and O2 on porous BiVO4 thin-film electrodes under visible light and significant effect of Ag ion treatment. The Journal of Physical Chemistry. B. 2006;3:11352-11360. DOI: 10.1021/jp057539+
  23. 23. Su J, Guo L, Bao N, Grimes CA. Nanostructured WO3/BiVO4 heterojunction films for efficient photoelectrochemical water splitting. Nano Letters. 2011;11:1928-1933. DOI: 10.1021/nl2000743
  24. 24. Sui M, Han C, Wang Y, Li J, Gu X. Photoelectrochemical studies on BiVO4 membranes deposition on transparent conductive substrates by a facile electrophoresis route. Journal of Materials Science: Materials in Electronics. 2016;27:4290-4296. DOI: 10.1007/s10854-016-4295-x
  25. 25. Zhang Y, Guo Y, Duan H, Li H, Sun C, Liu H. Facile synthesis of V4+ self-doped, [010] oriented BiVO4 nanorods with highly efficient visible light- induced photocatalytic activity. Physical Chemistry Chemical Physics. 2014;16:24519-24526. DOI: 10.1039/C4CP03795B
  26. 26. Neves MC, Trindade T. Chemical bath deposition of BiVO4. Thin Solid Films. 2002;406:93-97. DOI: 10.1016/S0040-6090(01)01787-4
  27. 27. Seabold JA, Choi K. Efficient and stable photo-oxidation of water by a bismuth vanadate photoanode coupled with an iron oxyhydroxide oxygen evolution catalyst. Journal of the American Chemical Society. 2012;134:2186-2192. DOI: 10.1021/ja209001d
  28. 28. Song J, Cha J, Lee MG, Jeong HW, Seo S, Yoo JA, Kim TL, Lee J, No H, Kim DH, Jeong SY, An H, Lee BH, Bark CW, Park H, Jang HW, Lee S. Template-engineered epitaxial BiVO4 photoanodes for efficient solar water splitting. Journal of Materials Chemistry A. 2017;5:18831-18838. DOI: 10.1039/C7TA04695B
  29. 29. Van CN, Do TH, Chen J-W, Tzeng W-Y, Tsai K-A, Song H, Liu H-J, Lin Y-C, Chen Y-C, Wu C-L, Luo CW, Chou W-C, Huang R, Hsu Y-J, Chu Y-H. WO3 mesocrystal-assisted photoelectrochemical activity of BiVO4. NPG Asia Materials. 2017;9:e357. DOI: 10.1038/am.2017.15
  30. 30. Kim JH, Lee JS. BiVO4-based Heterostructured Photocatalysts for solar water splitting: A review. Energy and Environment Focus. 2014;3:339-353. DOI: 10.1166/eef.2014.1121
  31. 31. Qiu Y, Liu W, Chen W, Chen W, Zhou G, Hsu P-C, Zhang R, Liang Z, Fan S, Zhang Y, Cui Y. Efficient solar-driven water splitting by nanocone BiVO4-perovskite tandem cells. Science Advances. 2016;2:e1501764. DOI: 10.1126/sciadv.1501764
  32. 32. Zhao J, Guo Y, Cai L, Li H, Wang KX, Cho IS, Lee CH, Fan S, Zheng X. High-performance ultrathin BiVO4 photoanode on textured polydimethylsiloxane substrates for solar water splitting. ACS Energy Letters. 2016;1:68-75. DOI: 10.1021/acsenergylett.6b00032
  33. 33. Kim TW, Choi K-S. Nanoporous BiVO4 photoanodes with dual-layer oxygen evolution catalysts for solar water splitting. Science. 2014;343:990-995. DOI: 10.1126/science.1246913
  34. 34. Shi X, Choi IY, Zhang K, Kwon J, Kim DY, Lee JK, Oh SH, Kim JK, Park JH. Efficient photoelectrochemical hydrogen production from bismuth vanadate-decorated tungsten trioxide helix nanostructures. Nature Communications. 2014;5:4775. DOI: 10.1038/ncomms5775
  35. 35. Park Y, McDonald KJ, Choi K-S. Progress in bismuth vanadate photoanodes for use in solar water oxidation. Chemical Society Reviews. 2013;42:2321-2337. DOI: 10.1039/c2cs35260e
  36. 36. Rettie AJE, Lee HC, Marshall LG, Lin J-F, Capan C, Lindemuth J, McCloy JS, Zhou J, Bard AJ, Mullins CB. Combined charge carrier transport and photoelectrochemical characterization of BiVO4 single crystals: Intrinsic behavior of a complex metal oxide. Journal of the American Chemical Society. 2013;135(30):11389-11396. DOI: 10.1021/ja405550k
  37. 37. Li R, Zhang F, Wang D, Yang J, Li M, Zhu J, Zhou X, Han H, Li C. Spatial separation of photogenerated electrons and holes among {010} and {110} crystal facets of BiVO4. Nature Communications. 2013;4:1432. DOI: 10.1038/ncomms2401
  38. 38. Yang J, Wang D, Zhou X, Li C. A theoretical study on the mechanism of photocatalytic oxygen evolution on BiVO4 in aqueous solution. Chemistry–A European Journal. 2013;(4):1320-1326. DOI: 10.1002/chem.201202365
  39. 39. Li R, Han H, Zhang F, Wang D, Li C. Highly efficient photocatalysts constructed by rational assembly of dual-cocatalysts separately on different facets of BiVO4. Energy & Environmental Science. 2014;7:1369-1376. DOI: 10.1039/c3ee43304h
  40. 40. Tolod KR, Hernández S, Russo N. Recent advances in the BiVO4 photocatalyst for sun-driven water oxidation: Top-performing photoanodes and scale-up challenges. Catalysts. 2017;7:13. DOI: 10.3390/catal7010013
  41. 41. Li C, Zhang P, Lv R, Lu J, Wang T, Wang S, Wang H, Gong J. Selective deposition of Ag3PO4 on monoclinic BiVO4(040) for highly efficient photocatalysis. Small. 2013;(23):3951-3956. DOI: 10.1002/smll.201301276
  42. 42. Balachandran S, Prakash N, Thirumalai K, Muruganandham M, Sillanpa M, Swaminathan M. Facile construction of heterostructured BiVO4−ZnO and its dual application of greater solar photocatalytic activity and self-cleaning property. Industrial and Engineering Chemistry Research. 2014;53:8346-8356. DOI: 10.1021/ie404287m
  43. 43. Gao X, Wu HB, Zheng L, Zhong Y, Hu Y, Lou XWD. Formation of mesoporous heterostructured BiVO4/Bi2S3 hollow discoids with enhanced photoactivity. Angewandte Chemie, International Edition. 2014;53:5917-5921. DOI: 10.1002/anie.201403611
  44. 44. Grigioni I, Stamplecoskie KG, Selli E, Kamat PV. Dynamics of photogenerated charge carriers in WO3/BiVO4 heterojunction photoanodes. Journal of Physical Chemistry C. 2015;119:20792-20800. DOI: 10.1021/acs.jpcc.5b05128
  45. 45. Rao PM, Cai L, Liu C, Cho IS, Lee CH, Weisse JM, Yang P, Zheng X. Simultaneously efficient light absorption and charge separation in WO3/BiVO4 core/shell nanowire photoanode for photoelectrochemical water oxidation. Nano Letters. 2014;14(2):1099-1105. DOI: 10.1021/nl500022z
  46. 46. Hong SJ, Lee S, Jang JS, Lee JS. Heterojunction BiVO4/WO3 electrodes for enhanced photoactivity of water oxidation. Energy & Environmental Science. 2011;4:1781-1787. DOI: 10.1039/c0ee00743a
  47. 47. Ma M, Kim JK, Zhang K, Shi X, Kim SJ, Moon JH, Park J. Double-deck inverse opal photoanodes: Efficient light absorption and charge separation in heterojunction. Chemistry of Materials. 2014;26:5592-5597. DOI: 10.1021/cm502073d
  48. 48. Pihosh Y, Turkevych I, Mawatari K, Uemura J, Kazoe Y, Kosar S, Makita K, Sugaya T, Matsui T, Fujita D, Tosa M, Kondo M, Kitamori T. Photocatalytic generation of hydrogen by core-shell WO3/ BiVO4 nanorods with ultimate water splitting efficiency. Scientific Reports. 2015;5:11141. DOI: 10.1038/srep11141
  49. 49. Fu Y, Sun X, Wang X. BiVO4–graphene catalyst and its high photocatalytic performance under visible light irradiation. Materials Chemistry and Physics. 2011;131:325-330. DOI: 10.1016/j.matchemphys.2011.09.049
  50. 50. Ng YH, Iwase A, Kudo A, Amal R. Reducing graphene oxide on a visible-light BiVO4 photocatalyst for an enhanced photoelectrochemical water splitting. Journal of Physical Chemistry Letters. 2010;1:2607-2612. DOI: 10.1021/jz100978u
  51. 51. Wu X, Zhao J, Guo S, Wang L, Shi W, Huang H, Liu Y, Kang Z. Carbon dot and BiVO4 quantum dot composites for overall water splitting via a two-electron pathway. Nanoscale. 2016;8:17314-17321. DOI: 10.1039/c6nr05864g
  52. 52. Wang Y, Wang W, Mao H, Lu Y, Lu J, Huang J, Ye Z, Lu B. Electrostatic self-assembly of BiVO4−reduced graphene oxide nanocomposites for highly efficient visible light photocatalytic activities. ACS Applied Materials & Interfaces. 2014;6:12698-12706. DOI: 10.1021/am502700p
  53. 53. Zhao D, Zong W, Fan Z, Xiong S, Du M, Wu T, Fang Y-W, Ji F, Xu X. Synthesis of carbon-doped BiVO4@multi-walled carbon nanotubes with high visible-light absorption behavior, and evaluation of their photocatalytic properties. CrystEngComm. 2016;18:9007-9015. DOI: 10.1039/C6CE01642A
  54. 54. Rao PM, Cai L, Liu C, Cho IS, Lee CH, Weisse JM, Yang P, Zheng X. Simultaneously efficient light absorption and charge separation in WO3/BiVO4 Core/Shell nanowire photoanode for photoelectrochemical water oxidation. Nano Letters. 2014;14(2):1099-1105. DOI: 10.1021/nl500022z
  55. 55. Berglund SP, Rettie AJE, Hoang S, Mullins CB. Incorporation of Mo and W into nanostructured BiVO4 films for efficient photoelectrochemical water oxidation. Physical Chemistry Chemical Physics. 2012;14:7065-7075. DOI: 10.1039/c2cp40807d
  56. 56. Park HS, Kweon KE, Ye H, Paek E, Hwang GS, Bard AJ. Factors in the metal doping of BiVO4 for improved photoelectrocatalytic activity as studied by scanning electrochemical microscopy and first-principles density-functional calculation. Journal of Physical Chemistry C. 2011;115:17870-17879. DOI: 10.1021/jp204492r
  57. 57. Han L, Abdi FF, Krol RVD, Liu R, Huang Z, Lewerenz H-J, Dam B, Zeman M, Smets AHM. Efficient water-splitting device based on a bismuth vanadate photoanode and thin-film silicon solar cells. ChemSusChem. 2014;7:2832-2838. DOI: 10.1002/cssc.201402456
  58. 58. Abdi FF, Han L, Smets AHM, Zeman M, Dam B, Krol RVD. Efficient solar water splitting by enhanced charge separation in a bismuth vanadate-silicon tandem photoelectrode. Nature Communications. 2013;4:2195. DOI: 10.1038/ncomms3195
  59. 59. 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
  60. 60. Snyder GJ, Toberer ES. Complex thermoelectric materials. Nature Materials. 2008;7:105-114. DOI: 10.1038/nmat2090
  61. 61. Dongfang Y, editor. Applications of Laser Ablation: Thin Film Deposition, Nanomaterial Synthesis and Surface Modification. Croatia: Intech; 2016. 55-84. doi: dx.doi.org/10.5772/65898
  62. 62. Rowe DM, editor. Thermoelectrics Handbook: Macro to Nano. FL, Boca Raton: CRC/Taylor & Francis; 2006
  63. 63. Nolas GS, Sharp J, Goldsmid HJ. Thermoelectrics: Basic Principles and New Materials Developments. New York: Springer; 2001
  64. 64. Schumacher C, Reinsberg KG, Rostek R, Akinsinde L, Baessler S, Zastrow S, Rampelberg G, Woias P, Detavernier C, Broekaert JAC, Bachmann J, Nielsch K. Optimizations of pulsed plated p and n-type Bi2Te3-based ternary compounds by annealing in different ambient atmospheres. Advanced Energy Materials. 2013;3:95-104. DOI: 10.1002/aenm.201200417
  65. 65. Soni A, Yanyuan Z, Ligen Y, Aik MKK, Dresselhaus MS. Enhanced thermoelectric properties of solution grown Bi2Te3-xSex nanoplatelet composites. NanoLetters. 2012;12:1203-1209. DOI: 10.1021/nl2034859
  66. 66. 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. Applied Surface Science. 2013;285P:657-663. DOI: 10.1016/j.apsusc.2013.08.107
  67. 67. Le PH, Liao C-N, Luo CW, Lin J-Y, Leu J. Thermoelectric properties of nanostructured bismuth–telluride thin films grown using pulsed laser deposition. Journal of Alloys and Compounds. 2014;615:546-552. DOI: 10.1016/j.jallcom.2014.07.018
  68. 68. Tuyen LTC, Le PH, Luo CW, Leu J. Thermoelectric properties of nanocrystalline Bi3Se2Te thin films grown using pulsed laser deposition. Journal of Alloys and Compounds. 2016;673:107-114. DOI: 10.1016/j.jallcom.2016.03.006
  69. 69. Huang B, Lawrence C, Gross A, Hwang G-S, Ghafouri N, Lee S-W, Kim H, Li C-P, Uher C, Najafi K, Kaviany M. Low-temperature characterization and micropatterning of coevaporated Bi2Te3 and Sb2Te3 films. Journal of Applied Physics. 2008;104:113710. DOI: 10.1063/1.3033381
  70. 70. Böttner H, Nurnus J, Gavrikov A, Kühner G, Jägle M, Künzel C, Eberhard D, Plescher G, Schubert A, Schlereth K-H. New thermoelectric components using microsystem technologies. International Journal of Energy & Environment. 2004;13:414-420. DOI: 10.1109/JMEMS.2004.828740
  71. 71. Takashiri M, Shirakawa T, Miyazaki K, Tsukamoto H. Fabrication and characterization of bismuth-telluride-based alloy thin film thermoelectric generators by flash evaporation method. Sensors and Actuators A. 2007;138:329-334. DOI: 10.1016/j.sna.2007.05.030
  72. 72. 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. Journal of Applied Physics. 2007;101:74301. DOI: 10.1063/1.2717867
  73. 73. 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. Journal of Applied Physics. 2008;104:84302. DOI: 10.1063/1.2990774
  74. 74. Takashiri M, Tanaka S, Miyazaki K, Tsukamoto H. Cross-plane thermal conductivity of highly oriented nanocrystalline bismuth antimony telluride thin films. Journal of Alloys and Compounds. 2010;490:L44-L47. DOI: 10.1016/j.jallcom.2009.10.117
  75. 75. Liao C-N, Wang Y-C, Chu H-S. Thermal transport properties of nanocrystalline Bi-Sb-Te thin films prepared by sputter deposition. Journal of Applied Physics. 2008;104:104312. DOI: 10.1063/1.3026728
  76. 76. 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
  77. 77. 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
  78. 78. Noro H, Sato K, Kagechika H. The thermoelectric properties and crystallography of bi-Sb-Te-se thin films grown by ion beam sputtering. Journal of Applied Physics. 1993;73:1252-1260. DOI: 10.1063/1.353266
  79. 79. Tritt TM. Thermoelectric phenomena, materials, and applications. Annual Review of Materials Research. 2011;41:433-448. DOI: 10.1146/annurev-matsci-062910-100453
  80. 80. 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. Journal of Applied Physics. 2009;105:124307. DOI: 10.1063/1.3147870
  81. 81. Chang H-C, Chen T-H, Whang W-T, Chen C-H. Superassembling of Bi2Te3 hierarchical nanostructures for enhanced thermoelectric performance. Journal of Materials Chemistry A. 2015;3:10459-10465. DOI: 10.1039/c5ta00911a
  82. 82. Wang G, Endicott L, Uher C. Recent advances in the growth of Bi-Sb-Te-Se thin films. Science of Advanced Materials. 2011;3:539-560. DOI: 10.1166/sam.2011.1182
  83. 83. https://www.lairdtech.com/
  84. 84. Martin J, Wang L, Chen L, Nolas G. Enhanced seebeck coefficient through energy-barrier scattering in PbTe nanocomposites. Physical Review B. 2009;79:115311. DOI: 10.1103/PhysRevB.79.115311
  85. 85. Deng Y, Liang H-M, Wang Y, Zhang Z-W, Tan M, Cui JL. Growth and transport properties of oriented bismuth telluride films. Journal of Alloys and Compounds. 2011;509:5683-5687. DOI: 10.1016/j.jallcom.2011.02.123
  86. 86. Zhang Z, Wang Y, Deng Y, Xu Y. The effect of (00l) crystal plane orientation on the thermoelectric properties of Bi2Te3 thin film. Solid State Communications. 2011;151:1520-1523. DOI: 10.1016/j.ssc.2011.07.036
  87. 87. Peranio N, Eibl O, Nurnus J. Structural and thermoelectric properties of epitaxially grown Bi2Te3 thin films and superlattices. Journal of Applied Physics. 2006;100:114306. DOI: 10.1063/1.2375016
  88. 88. Zou H, Rowe DM, Min G. Growth of p- and n-type bismuth telluride thin films by co-evaporation. Journal of Crystal Growth. 2001;222:82-87. DOI: 10.1016/S0022-0248(00)00922-2
  89. 89. Kadel K, Kumari L, Li WZ, Huang JY, Provencio PP. Synthesis and thermoelectric properties of Bi2Se3 nanostructures. Nanoscale Research Letters. 2010;6:57. DOI: 10.1007/s11671-010-9795-7
  90. 90. Sun Z, Liufu S, Chen L. Synthesis and characterization of nanostructured bismuth selenide thin films. Dalton Transactions. 2010;39:10883-10887. DOI: 10.1039/c0dt00840k
  91. 91. Al Bayaz A, Giani A, Artaud MC, Foucaran A, Delannoy FP, Boyer A. Growth parameters effect on the electric and thermoelectric characteristics of Bi2Se3 thin films grown by MOCVD. Journal of Crystal Growth. 2002;241:463-470. DOI: 10.1016/S0022-0248(03)01511-2
  92. 92. 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. Applied Physics Letters. 2005;87:081902. DOI: 10.1063/1.2001755
  93. 93. Venkatasubramanian R. Lattice thermal conductivity reduction and phonon localization like behavior in superlattice structures. Physical Review B. 2000;61:3091-3097. DOI: 10.1103/PhysRevB.61.3091

Written By

Phuoc Huu Le, Nguyen Trung Kien and Chien Nguyen Van

Submitted: 22 December 2017 Reviewed: 19 February 2018 Published: 20 June 2018