Open access peer-reviewed chapter

Thermoelectric and Topological Insulator Bismuth Chalcogenide Thin Films Grown Using Pulsed Laser Deposition

Written By

Phuoc Huu Le and Chih Wei Luo

Submitted: 22 March 2016 Reviewed: 21 September 2016 Published: 21 December 2016

DOI: 10.5772/65898

Chapter metrics overview

2,106 Chapter Downloads

View Full Metrics

Abstract

Bismuth chalcogenides have been intensively studied for their high-performance thermoelectric properties and their novel topological surface states, which could significantly benefit novel applications in fields such as TE devices, spintronics, and quantum computing. This chapter reports recent advances in pulsed laser deposition (PLD) for the growth of bismuth chalcogenide (e.g., Bi2Te3, Bi2Se3, and Bi3Se2Te) thin films and their novel properties. It covers a wide range of fields such as thin film growth using PLD for fabricating polycrystalline and epitaxial films with different thermoelectric, nanomechanical, and magnetotransport properties as a function of the PLD processing conditions. Moreover, the proximity-induced superconductivities in Bi inclusions/bismuth chalcogenide thin films are also reported and discussed in detail.

Keywords

  • pulsed laser deposition
  • thermoelectrics
  • topological insulators
  • bismuth chalcogenides
  • superconductivity
  • magnetoresistance

1. Introduction

Bismuth chalcogenide thin films are of great interest because of the exciting properties of topological insulators (TIs) and their applications to thermoelectrics (TEs). These materials have been applied in integrated TE cooling devices working at near room temperature [1, 2]. TIs are exotic materials with an insulating bulk and topologically protected states on the surface which could be used in different applications, such as spintronics and quantum computing [36]. The topological surface states (TSSs) exhibit Dirac linear energy dispersion inside the bulk gap, spin-polarization by spin-momentum locking nature, and weak anti-localization (WAL) due to the strong spin-orbit coupling [36]. Thus, the WAL through magnetotransport studies has been widely used as a signature of TI materials [79].

For application purposes, thin film growth techniques for TE and TI materials are required. Among physical vapor deposition techniques, pulsed laser deposition (PLD) offers great versatility in growing polycrystalline and epitaxial thin films with high growth rates, multiple elements, and diverse structural morphologies for both fundamental studies and applications. The purpose of this chapter is to outline recent advances in the PLD growths of bismuth chalcogenide thin films with desired properties for TE/TI applications and fundamental studies.

Advertisement

2. Thin film growth using pulsed laser deposition (PLD)

Thin film growth consists of nucleation, growth, and coalescence (Figure 1a). In a physical vapor deposition, an extremely nonequilibrium process takes place at high supersaturations and at comparatively high concentrations of impure atoms [10]. Nucleation takes place at high supersaturations S (defined as S = p/pe, where p is the vapor pressure of the deposited material evaporated from the source at temperature T and pe is the equilibrium vapor pressure of the substrate material at temperature TS). The incident vapor arrives at the surface of substrates and then forms small but highly mobile clusters or islands with uniform distribution. In this stage, the impinging atoms and subcritical clusters are incorporated and consequently increase their sizes, while the island density rapidly saturates. In the following stage, the islands are emerged via a coalescence phenomenon which is liquid-like for some cases, especially at high substrate temperatures. Coalescence leads to a decrease in island density and forms local denuding positions on the surface of substrates where further nucleation can then occur (Figure 1a) [11]. The sequence of film nucleation and growth events can be well appreciated in the transmission electron microscopy (TEM) images in Figure 1bd [11].

2.1. Basic growth modes

For all phase transitions, the formation of thin films is characterized by the formation of nuclei and their growth. Depending on the interaction energies of substrate atoms and film atoms, any of three growth modes in Figure 2ac can occur:

  • 2D Frank-van der Merwe mode: layer-by-layer growth, in which the interaction between substrate and atoms of film is greater than that between adjacent atoms of film.

  • 3D Volmer-Weber mode: separated islands form on the surface of substrates, in which the interaction between atoms of film is greater than that between a substrate and the adjacent atoms of film.

  • Stranski-Krastanov mode: layer plus island, in which one or two monolayers form first and then grow individually.

Figure 1.

(a) Schematics of thin film growth processes: nucleation, growth, coalescence. Transmission electron microscope images of (b) nucleation, (c) growth, and (d) coalescence of Ag films on (1 1 1) NaCl substrates. Corresponding diffraction patterns are shown.

Figure 2.

Basic modes of thin film growth: (a) island in the Volmer-Weber mode, (b) layer by layer in the two-dimensional Frank-van der Merwe mode, (c) layer plus island in the Stranski-Krastanov mode. (d) Shadowing growth: a schematic of three-dimensional Monte Carlo simulations for oblique angle deposition [13].

After an initially random nucleation of islands on the surface of the substrates, the deposition on the top of the islands is faster than that in the valleys due to the oblique incident flux (the so-called shadowing effect) [12, 13]. Isolated columns are therefore formed on these islands during subsequent growths (Figure 2d) [13].

2.2. Epitaxy growth

Epitaxy refers to the growth of a single crystal film on top of a single crystal substrate. The deposited film is denoted as an epitaxial film or epitaxial layer. The growth is called homoepitaxy if the film and the substrate are the same material, and it is called heteroepitaxy if they are different materials. Epitaxial relationship is determined as: (HKL) || (hkl); [UVW] || [uvw], where (hkl) and (HKL) are the Miller indices of the overgrowth plane and substrate at the common interface. The corresponding parallel directions in the overgrowth and substrate planes, denoted by [wuw] and [UVW], respectively, must also be specified.

2.3. Factors governing the epitaxy growth

The key factors governing epitaxy growths are structural compatibility, chemical compatibility, and growth temperatures.

- Structural compatibility: The structures of a film and a substrate should have good lattice matching in terms of crystal structures (a0, sub) and lattice constants (a0, film), that is, small lattice misfit.

Lattice misfit f:

f=a0,suba0,filma0,filma0,suba0,filma0,sub2.E1

- Chemical compatibility: This includes chemical bonding and chemical diffusion.

- Growth temperatures: Good epitaxy growth is obtained at above or around the well-defined elevated substrate temperature (Te). Te depends on the deposition rate, particle energy, and surface contamination. Generally, a higher temperature is recommended to reduce surface contamination by desorption (or enhance surface mobility) of atoms to reach the favorable sites and also enhance the diffusivity in deposition for favoring re-crystallization and defect annihilation.

2.4. Preparation of bismuth chalcogenide films by PLD

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 then deposit thin films onto a substrate. Figure 3 shows a typical PLD system for preparing TE and TI thin films. A substrate is heated and maintained at a desired substrate temperature (TS) using a thermocouple and a proportional-integral-derivative temperature controller. The thermocouple was buried inside a stainless steel substrate holder, which is heated by a tungsten lamp just behind the holder. The pressure of ambient gas (He/Ar) can be fine-tuned by the needle valve. A KrF excimer laser beam (λ = 248 nm, pulsed duration of 15–20 ns, repetition rate in the range of 1–15 Hz, and fluence of 1–10 J/cm2) is guided by several UV mirrors and focused on a stoichiometric polycrystalline or a single crystal target (e.g., Bi2Te3, Bi2Se3, Bi2Se2Te) within the vacuum chamber by a UV lens. The target-to-substrate distance was 40 mm. During the deposition of Bi2Se3 films, pure (6N) He/Ar gas was introduced into the vacuum chamber, which was evacuated to a base pressure of 4 × 10−4 Pa (or 3 × 10−6 Torr) and maintained at a certain constant pressure (P), using a differential evacuation system.

Figure 3.

Schematic illustration of a pulsed laser deposition (PLD) system. G: gauge.

The surface of substrates should be atomically clean and free from impurities because the contaminants can interact with the thin films being deposited and substantially degrade its quality and adhesion to the substrates. The presence of unwanted surface contaminants can also influence the growth and orientation of the films in an undesired manner. In the depositions for TE thin films, an approximately 300-nm-thick SiO2 layer was thermally grown on the Si wafers (thickness 525 μm) for electrical isolation purpose. The wafers were cut into 1.5 cm × 1.5 cm substrates. The substrates were cleaned with acetone to dissolve any contaminants adhering to the surface of substrates such as grease and oils. This was followed by rinsing with methanol to remove any residues left behind after cleaning with acetone. Afterward, the substrates were rinsed in distilled water and dried with nitrogen flow. The substrates were then used for the deposition of TE thin films.

Here are some examples of PLD growth of TE films. For Bi2Se3 thin films, the depositions were at TS of 200–350°C and helium ambient pressure (P) of 0.7–173 Pa. The number of laser pulses was 9000 and the deposition took 30 min. The average growth rate was approximately 0.46 Å/pulse [14]. For the growth of Bi2Te3 thin films, TS was varied from room temperature (30°C) to 380°C and the Ar ambient pressure (PAr) was at 80 Pa. The number of laser pulses was 12,000 and the deposition took 40 min. The average growth rate was approximately 0.52 Å/pulse [15]. For the growth of Bi-Se-Te thin films, the depositions were at TS of 200–350°C and a helium ambient pressure (PHe) of 0.027–86.7 Pa. The number of laser pulses was 9000 and the deposition took 15 min. The average growth rate was approximately 0.6 Å/pulse [16].

Advertisement

3. Thermoelectric bismuth chalcogenide thin films

3.1. Crystal structures of bismuth chalcogenides

The crystal structures of bismuth chalcogenides (e.g., Bi2Te3, Bi2Se3, and Bi3Se2Te) are usually described by a hexagonal cell consisting of 15 layers of atoms stacking along the c-axis with a sequence shown below [17], as shown in Figure 4a and b.

Figure 4.

Crystal structures of (a) Bi2Te3, (b) Bi2Se3, and (c) Bi3Se2Te in the form of (Bi2)m(Bi2Se2Te)n (m = 1, n = 2) homologous series. The unit cells are marked with black thick boxes. (d) X-ray diffraction patterns of the typical TE bismuth chalcogenide thin films grown at 300°C on SiO2/Si substrates.

A 5-atomic-layer-thick lamellae of ‐(Te(1)‐Bi‐Te(2)‐Bi‐Te(1))‐ or ‐(Se(1)‐Bi‐Se(2)‐Bi‐Se(1))‐ is called a quintuple layer, QL, in which the Te(1)‐Bi and Bi‐Te(2) or Se(1)‐Bi and Bi‐Se(2) are ionic-covalent bonds. Because of the weak binding (i.e., Van der Waals force) between Te or Se layers, bismuth chalcogenides could be cleaved easily along the plane perpendicular to the c-axis. It also induces the anisotropic thermal/electrical transport properties. For example, the thermal conductivity along the plane perpendicular to the c-axis (~1.5 W m−1 K−1) is nearly two times higher than that along the c-axis direction (~0.7 W m−1 K−1) [17, 18].

The crystal structures of Bi3Se2Te can be formed by ordered stacking of Bi2Se2Te and Bi2 building blocks, that is, (Bi2)m(Bi2Se2Te)n (m = 1, n = 2) [19, 20], in which the covalently connected double layers of bismuth (Bi–Bi) lie between two QL (Se‐Bi‐Te‐Bi‐Se) blocks (Figure 4c); the (Bi–Bi) strictly alternates with two QL (Se–Bi–Te–Bi–Se) blocks [9, 20]. Figure 4d shows the typical XRD patterns of Bi2Te3, Bi2Se3, and Bi3Se2Te thin films grown on SiO2/Si substrates at TS = 300°C. They exhibit the dominance of (0 0 1) family planes of the rhombohedral phases of Bi2Te3 (PDF#82-0358), Bi2Se3 (PDF#33-0214), and Bi3Se2Te (JCPDS 00-053-1190), indicating that the films are highly c-axis oriented (i.e., textured films).

3.2. Introduction to thermoelectrics and applications

Thermoelectric materials are solid-state energy converters in which the 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 [21].

3.2.1. 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 with the dimensionless unit [17, 22].

ZT=α2σκT=α2σκE+κLTE2

where σ, α, κ, and T are the electrical conductivity, Seebeck coefficient, thermal conductivity, and absolute temperature, respectively. The total thermal conductivity can be split into a lattice contribution (κL) and an electronic contribution (κE). The quantity α2σ is commonly used to represent the thermoelectric power factor (PF).

The efficiency of a thermoelectric device is directly related to ZT. For power generation, the maximum efficiency (η) is expressed by [23]

η=ThTcTh1+ZT¯11+ZT¯+TcThE3

and for air-conditioning or refrigeration, the coefficient of performance is [23]

COP=TcThTc1+ZT¯ThTc1+ZT¯+1E4

where Th and Tc are the hot-end and cold-end temperatures of the thermoelectric materials, respectively, and T¯ is the average temperature of Th and Tc. Therefore, the enhanced ZT value of TE materials is important to increase the COP for practical applications.

3.2.2. Conflicting properties in thermoelectric materials

Maximizing ZT is challenging due to the interdependence of the TE parameters. An increased power factor α2σ by optimizing the carrier concentration n and/or a reduced lattice thermal conductivity κL by introducing the scattering centers are necessary to enhance ZT value. The dependences of these parameters with scattering factor r, carrier effective mass m*, carrier mobility μ, and their interconnectivity limits the ZT to about 1 in large bulk materials [24].

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

1/ρ=σ=neμ.E5

The Wiedemann-Franz Law [2] states that the electronic contribution to the thermal conductivity is proportional to the electrical conductivity (σ) of the materials, with the relationship being

κe=LσT=neμLTE6

where e is electron charge, and L is the Lorenz factor of 2.48 × 10−8 J2/K2 C2 for free electrons and this can vary particularly with carrier concentration [2].

The kinetic definition of α is the energy difference between the average energy of mobile carriers and the Fermi energy [25]. An increase in n leads to the increase in both the Fermi energy and the average energy, but the former increases more rapidly than the latter and thus results in a decrease in α value and a reduction factor of α2n. Thus, in attempting to increase ZT for most of the homogeneous materials, the carrier concentration (n) increases electrical conductivity (σ) but reduces the Seebeck coefficient (α). For this reason, in metals and degenerate semiconductors (energy-independent scattering approximation), the Seebeck coefficient can be expressed as [2]:

α=8π2kB23eh2m*T(π3n)2/3.E7

The high m* causes the Seebeck coefficient to rise according to Eq. (7). High m* materials generally possess low μ which limits the enhancement of power factor by (m*)3/2μ. Noticeably, the defect scatters are not only the phonons but also the electrons which lead to reduce κL as well as μ. Therefore, the ratio of μ/κL determines the improvement in ZT [17, 24]. Although the increase in the ratio is usually experimentally achieved through a greater reduction in κL rather than that in μ, some fundamental issues in this mechanism are not understood well [24].

Figure 5 shows the compromise between large α and high σ in thermoelectric materials that must be struck to maximize the figure of merit ZT. Meanwhile, the low carrier concentration will result in lower electrical conductivity with decreasing ZT. The ZT and PF peaks typically occur at carrier concentrations between 1019 and 1021 carriers per cm3 (depending on the material system), which fall in between common metals and semiconductors, that is, the concentrations found in heavily doped semiconductors [2].

Figure 5.

Maximizing the efficiency (ZT) of a thermoelectric device involves a compromise of thermal conductivity (κ, plotted on the y-axis from 0 to a top value of 10 W m−1 K−1) and Seebeck coefficient (α, 0–500 μV K−1) with electrical conductivity (σ, 0–5000 Ω−1 cm−1) [2].

3.2.3. Overview of thermoelectric applications

TE devices have unique features: no moving parts, substantially less maintenance, quiet operation, high power density, low environmental impact, and high reliability [26]. Commercial use has been made mostly from Peltier thermoelectric cooling (TEC) effect, such as in small refrigerator devices used for camping and outdoor activities, automotive climate control seats, and localized cooling at the hot spots of chips. Figure 6 gives an overview of the present and potential applications of thermoelectric generators (TEGs) [27]. Indeed, TEGs have been used for the power in miniaturized autarkic sensor systems, automotive waste heat recovery systems, ventilated wood stove, heating systems, water boilers, and heat recovery in industry.

Figure 6.

Overview of the potential applications of thermoelectric generators [27].

3.3. Thermoelectric properties of polycrystalline Bi2Te3, Bi2Se3, and Bi3Se2Te thin films with controlled structure morphology

Some typical HRTEM images of Bi2Te3, Bi2Se3, and Bi3Se2Te grown using PLD are shown in Figure 7 [1416]. HRTEM images performed on a high μ Bi2Te3 film with nanodisk-like morphology grown at 220°C are shown in Figure 7a. Clearly, the lower inset in Figure 7a shows the film with uniform thickness of approximately 295 nm and a SiO2 layer with a thickness of 300 nm. It shows that projected period of 0.508 nm along the c-axis corresponds to the lattice spacing of the (0 0 6) planes. The highly (0 0 1)-orientated and crystallized structures of the film should facilitate the transport of charge carriers. The c-axis lattice constant of the Bi2Te3 film is 30.48 Å, which agrees closely with the value (30.44 Å) presented in JCPDS 82-0358. The other Bi2Te3 films grown at TS ≥ 220°C also display similar HRTEM results.

For a Bi2Se3 film deposited at 300°C and 40 Pa, an HRTEM image taken at the boundary of three platelets (P1, P2, and P3) revealed the granular-polycrystalline structure of the films (Figure 7b). Moreover, P1 and P2 partly overlapped and the corresponding fast Fourier transform (FFT) of this overlapping region indexed by { 003 } patterns of [0 1 0] zone axis was performed from the dashed-square area (Figure 7b, inset). The projected period along the c-axes of both P1 and P2 was 9.60 Å, corresponding to (0 0 3) planes, which was close to the reported value of 9.55 Å in Ref. [28].

HRTEM images of a Bi3Se2Te film deposited at 250°C and 40 Pa are shown in Figure 7c and d. Nanocrystallites with sizes of 10–20 nm are clearly observed in Figure 7c, confirming the nanocrystalline type of the Bi3Se2Te films. The interplanar spacing of the Bi3Se2Te (0 0 5) planes in the nanocrystallites is approximately 0.464 nm. Therefore, the c-axis lattice constant is determined to be 23.2 Å, closely agreeing with the value of 23.25 Å for Bi3Se2Te bulk (JCPDS 00-053-1190). In addition, the white lines in Figure 7c indicate the orientations of the (0 0 5) planes. It is seen that the overall orientation of the crystallites is disorganized. Intriguingly, near the interface of the film and substrate, the film has some nanoinclusions with sizes of 12–17 nm, as shown in Figure 7d. The EDS analysis shows that these are Bi semimetal nanoprecipitates (the inset in Figure 7d). The lattice spacing of Bi nanoinclusions is ~0.32 nm, corresponding to the Bi (0 1 2) planes. It has been found that Bi nanoinclusions can lead to the enhanced Seebeck coefficient and reduced lattice thermal conductivity owing to the low-energy electron filtering and phonon scattering at the nanoinclusions, respectively [2931].

Figure 7.

(a) HRTEM images of a high carrier mobility (μ) Bi2Te3 film with nanodisk-like morphology grown at 220°C and PAr of 80 Pa. (b) HRTEM image of an optimized Bi2Se3 film deposited at 300°C and PHe of 40 Pa. The inset shows the FFT patterns of the dashed-square area in the HRTEM image. (c and d) HRTEM images of the Bi3Se2Te film grown at 250°C and PHe of 40 Pa. The white lines in (c) indicate the (0 0 5)-orientation of nanograins. Inset in (d): FFT patterns and EDS spectra performed at film and Bi nanoinclusion positions.

Figure 8a shows the TS-dependent α, σ, and PF (= α2σ) of some nanostructured Bi2Te3 films [15]. The σ value gradually increased from 34.5 ± 0.1 to 814.3 ± 1.5 S/cm when TS was increased from 30 to 300°C, and then sharply decreased to 647.3 ± 0.4 S/cm at 340°C and 414.0 ± 1.2 S/cm at 380°C. The enhanced σ (= 647.3 – 814.3 S/cm) of the films grown at 220–340°C originated from the substantially enhanced μ because the n exhibited a slight decrease [15]. Although the coupled relationship between σ (= neμ) and |α| (~n−2/3) generally constrains the concurrent enhancement of σ and |α|, a reduction in n and a substantial increase in μ in the same optimal range of TS (= 220–340°C) could lead to high values of both σ and |α|. Consequently, the PF of the stoichiometric Bi2Te3 films grown in the range of 220–340°C reached remarkably high values, ranging between 18.2 ± 0.25 and 24.3 ± 0.44 μW cm−1 K−2, whereas the PF was low (≤0.44 μW cm−1 K−2) in the case of nonstoichiometric films deposited at TS ≤ 120 or 380°C (Figure 8a).

Figure 8.

(a) Substrate temperature (TS) dependence of room temperature Seebeck coefficient α (red circles), electrical conductivity σ (blue triangulars), and power factor (PF = α2σ, black squares) of the Bi2Te3 and Bi4Te5 (for “PH” point) films. The morphology abbreviations: CNP, columnar nanoparticle; CNF, columnar nanoflower; ND, nanodisk; CP, compact polycrystalline; LTP, layered triangular platelet; PH, polyhedral. (b) Contour plot of the Bi2Se3 film’s PF as a function of PHe and TS. The morphology abbreviations: SC, smooth and compact; RG, rice grain; TP, triangular polygonal; S-LFs, super-layered flakes; L-HPs, layered hexagonal platelets. (c) Contour plot of the film’s PF as a function of TS from 200 to 350°C and PHe from 0.027 to 86.7 Pa. (d) |S| vs. σ of the films in this study and the relevant novel TE materials in the literature, listed in Table 1. Solid curves denote different PFs from 1 to 50 μW cm−1 K−2.

In order to check the evolution of the PF (= α2σ) as a function of PHe and TS, the contour plot is illustrated Figure 8b. The PF of Bi2Se3 films increased with increasing TS from 200 to 300°C because σ became considerably larger but the Seebeck coefficient diminished only slightly. However, for films deposited at 350°C, PF was lowered primarily because of the reduction in S and not the increase in σ. At intermediate pressures (40–93 Pa), the Bi2Se3 films remained stoichiometric or slightly Se-rich compositions, which in turn led to the reduced carrier concentrations and significantly enhanced the α values [14]. Thus, the PF of Bi2Se3 films grown at intermediate pressure was typically higher than the PF of films grown at a low or high pressure. The optimal value of PF was 5.54 ± 0.34 μW cm−1 K−2 for the layered hexagonal platelet Bi2Se3 films deposited at 300°C and 40 Pa [14].

System Type Method n
(1019
cm−3)
μ
(cm2/Vs) 
σ (S/cm)  α (μV/K)  PF (μW
cm−1 K−2)
References
Bi2Te3 Layered smooth film PLD 10.1 90.6 1464 −186 50.6 [39]
Bi2Te3 Layered compact polycrystalline PLD 5.0 102 814.3 −172.8 24.3 [15]
Bi2Te3 Nanoparticle film PLD 9.7 14.8 230 −91 1.9 [34]
Bi2Te3 Super-assembled film PLD 4.0 12.4 79 −113 1.0 [33]
Bi2Se3 Layered hexagonal platelets PLD 7.4 81.4 963.8 −75.8 5.5 [14]
Bi3Se2Te Nanocrystalline film PLD 35.5 34.4 1747.5 −68.8 8.3 [16]
Bi2Te3 Layered structure Sputtering 95 12.1 1840 −70 8.8 [35]
Bi2Se3 Bulk Melting and
hot-pressing
251.9 −175 7.7 [32]
Bi2Se0.3Te2.7 Bulk Ball milling hot pressing 892 −190 32.2 [36]
Bi2Se1.5Te1.5 Bulk Zone melting 1.2 230 441.6 −193 16.5 [37]
Bi2Se1.8Te1.2 Nanoplatelet bulk Polyol method 199.6 −80.9 1.3 [38]
Bi2Se2Te Bulk Ball milling hot pressing 1613 −60 5.8 [36]

Table 1.

Material, type, method, carrier concentration (n), mobility (μ), electrical conductivity (σ), Seebeck coefficient (S), power factor (PF = S2σ) of the optimal bismuth chalcogenide films in this study as compared to properties of Bi2Te3, Bi2Se3, Bi2SexTe1−x bulk and films reported in the literature. All the selected values were recorded at room temperature.

The TS- and PHe-dependent PF of nanocrystalline Bi3Se2Te films is further shown in Figure 8c. The films grown at 200°C only have PFs of 1.0–2.8 μW cm−1 K−2. The PFs of the films grown at higher TS are significantly enhanced because of their high σ values. Around TS = 250–350°C and PHe = 40 Pa, a window for high PF is clearly observed. An optimal PF of 8.3 μW cm−1 K−2 is achieved for a Bi3Se2Te film deposited at 250°C and 40 Pa.

Table 1 summarizes the transport and room-temperature TE properties of bismuth chalcogenides in the literature [14, 15, 3239]. For PLD growths, the highly (0 0 1)-oriented layered Bi2Te3 films achieved a PF of 50.6 μW cm−1 K−2 [39], and the layered compact polycrystalline film possessed a PF value of 24.3 μW cm−1 K−2 [15]. The Bi2Se3 films generally have lower TE properties than those of Bi2Te3 films. For example, the optimal PF of the Bi2Se3 films grown by PLD was 5.5 μW cm−1 K−2 [14], which was slightly lower than the PF of Bi2Se3 bulk (PF ≈ 7.7 μW cm−1 K−2 ) [32]. The nanocrystalline Bi3Se2Te films had an optimal PF of 8.3 μW cm−1 K−2 [16]. Further, PLD growth allows fabrication of nanostructured TE films with different morphologies of nanoparticle Bi2Te3 film (PF = 1.9 μW cm−1 K−2) [34] and super-assembled Bi2Te3 film (PF = 1.0 μW cm−1 K−2) [33]. The Bi2Te3 film deposited by the sputtering technique had PF of 8.8 μW cm−1 K−2 [35]. There are some reports of TE properties for bulk materials of bismuth chalcogenides, such as Bi2Se1.8Te1.2 nanoplatelet (PF ≈ 1.3 μW cm−1 K−2 ) [38], Bi2Se2Te (PF ≈ 5.8 μW cm−1 K−2 ), Bi2Se1.5Te1.5 (PF ≈ 16.5 μW cm−1 K−2 ) [37], and Bi2Se0.3Te2.7 (PF ≈ 32.2 μW cm−1 K−2 ) [36]. Unfortunately, the thermal conductivity κ of the films is missed in the reports to fully evaluate the TE performance of the films. Nevertheless, the κ of polycrystalline films with small grain sizes should be reduced thanks to the extensive phonon scattering at interfaces and grain boundaries.

Finally, Figure 8d shows the |S| vs. σ plot for the list in Table 1. The solid curves denote different values of PFs (= S2σ). It can be found that TE nanomaterials usually possess low σ values due to the separating or voided structure morphology, but bulk and thin films have superior σ. Note, the significant reduction in thermal conductivity κ is the key factor for employing nanostructured materials in the TE field.

Advertisement

4. Nanomechanical properties of Bi2Te3 and Bi3Se2Te thin films

Effects of helium ambient pressure (in PLD) on the nanomechanical properties of Bi2Te3 and Bi3Se2Te thin films have been investigated [9, 40]. The Bi2Te3 thin films were grown at TS of 250°C on c-plane sapphire substrates using excimer laser PLD with a power density of 5 J/cm2, at a repetition rate of 2 Hz. The helium pressures in PLD growth varied from 2 × 10−5 to 6.5 × 10−3 Torr. Similarly, Bi3Se2Te thin films were deposited on Al2O3 (0 0 0 1) substrates at a fixed substrate temperature of 250°C and PHe ranging from 2 × 10−5 to 6.5 × 10−1 Torr through PLD. The light source of the PLD system was a KrF excimer laser with λ = 248 nm, pulse duration of 20 ns, fluence of 7.0 J/cm2, and repetition rate of 2 Hz. The target-to-substrate distance, the number of laser pulses and deposition time were 40 mm, 3000, and 25 min, respectively. The grown films had the average thickness of 200 nm (the average growth rate was ~0.67 Å/pulse). Nanoindentation experiments were performed on a MTS Nano Indenter® XP system with a three-sided pyramidal Berkovich indenter tip using the continuous stiffness measurement (CSM) technique [41]. The hardness and Young’s modulus of the Bi2Te3 and Bi3Se2Te thin films were determined from the load-displacement results through the analytical method proposed by Oliver and Pharr [42].

As shown in Figure 9a, the hardness monotonically increased from 2.92 ± 0.12 to 4.02 ± 0.14 GPa for Bi2Te3 films, and from 2.5 ± 0.2 to 3.0 ± 0.1 GPa for Bi3Se2Te films when PHe was increased from 2.0 × 10−5 to 2.0 × 10−3 Torr. Similarly, the Young’s modulus of Bi2Te3 thin films was 106.31 ± 0.63, 115.51 ± 1.92, and 127.46 ± 9.21 GPa for PHe at 2.0 × 10−5, 2.0 × 10−4, and 2.0 × 10−3 Torr, respectively. For PHe > 2.0 × 10−3 Torr, the hardness (Young’s modulus) of Bi3Se2Te films continues to increase with increasing PHe, namely 3.2 ± 0.1 GPa (105.2 ± 10.2 GPa) at 2.0 × 10−1 Torr and 5.8 ± 0.2 GPa (188.5 ± 4.3 GPa) at 6.5 × 10−1 Torr (Figure 9b).

Figure 9.

Material and nanomechanical properties of Bi2Te3 and Bi3Se2Te thin films grown on Al2O3 (0 0 0 1) substrates at TS of 250°C and various helium pressures (PHe) between 2.0 × 10−5 and 6.5 × 10−1 Torr [9, 40]: (a and b) the hardness and Young’s modulus, (c) grain size (D), (d) the Hall-Petch behavior observed on the Bi3Se2Te thin films, in which the hardness is observed to increase approximately with D−1/2 (D is grain size).

The crystallite sizes (D) of the films were estimated using the Scherrer equation D = 0.9λ/Bcosθ, where λ, B, and θ are the X-ray wavelength, full width at the half maximum of the Bi2Te3 (0 0 15) peak or Bi3Se2Te (0 0 5) peak, and Bragg diffraction angle, respectively. The PHe-dependent D of the Bi2Te3 and Bi3Se2Te films is shown in Figure 9c. The grain size increases monotonically from 11.0 to 20.0 nm for Bi2Te3 and from 16.1 to 20.5 nm for Bi3Se2Te with increasing PHe from 2.0 × 10−5 to 2.0 × 10−3 Torr. In the nanoscale, grain size can affect significantly the mechanical properties of materials. The dislocation activities can be drastically suppressed in a polycrystalline material when the grain size is decreased, and thus the grain boundary sliding and/or grain rotations become the dominant deformation behavior, which in turn would lead to the manifestations of the inverse Hall-Petch effect [43]. Softening caused by grain boundary sliding is mainly attributed to large amount of defects in grain boundaries, which allow rapid diffusion of atoms and vacancies under stress [44]. Consequently, the plastic deformation of Bi2Te3 films should be dominated by the grain boundary sliding and/or grain rotation rather than the dislocation activity because of D ≤ 20 nm [40], which is consistent with the results in Refs. [4548].

In contrast, the 200-nm-thick Bi3Se2Te films with D of 16.1–25.1 nm grown at a larger PHe range of 2.0 × 10−5 to 6.5 × 10−1 Torr exhibited the nanomechanical followed Hall-Petch relationship [44, 49]. The hardness and Young’s modulus of the Bi3Se2Te thin films monotonically increased with increasing PHe because of a corresponding decrease in grain sizes (Figure 9ac). Figure 9d shows that hardness (H) increased linearly with D−1/2 (where D is the grain size of the Bi3Se2Te films in the nanoscale regime) which is the typical Hall-Petch relationship [44, 49]. This is because the multiplication and mobility of dislocations are hindered by reducing the grain size [44]. It is reasonable for the observed phenomenon when the present grain sizes ranged between 25.1 and 16.1 nm which is larger than the typical critical Dc of 10 nm [44, 49]. It is demonstrated that the hardness and Young’s modulus of the Bi2Te3 and Bi3Se2Te thin films can be enhanced by proper selection of the ambient pressure in PLD growths.

Advertisement

5. Topological insulator bismuth chalcogenide thin films and their novel properties

5.1. The epitaxial growths of bismuth chalcogenide thin films

Figure 10.

Bi2Te3 films grown on SrTiO3 (1 0 0) substrates [53]: (a) XRD φ-scan patterns of the 200-nm-thick Bi2Te3 (0 1 5) plane and the SrTiO3 (1 1 1) plane; (b) schematics of the in-plane arrangement of h-Bi2Te3 / SrTiO3 (1 0 0). Bi2Se3 films grown on c-plane sapphire substrates [64]; (c) in-plane φ-scan { 011¯5 } planes of the film and { 011¯2 } planes of the substrate; (d) the schematic depicting twin/domain growth: in one case, the basal plane of Bi2Se3 is aligned with that of Al2O3, and in the other, there is a rotation of 60°/180°. Bi3Se2Te films grown on c-plane sapphire substrates [9]; (e) φ-scan patterns of hexagonal h-Bi3Se2Te thin films grown on Al2O3 (0 0 0 1) substrates; and (f) schematics of the in-plane arrangement of hexagonal h-Bi3Se2Te/h-Al2O3 (0 0 0 1).

Topological insulator (TI), a new class of quantum matter, possesses insulating in bulk and robust gapless topological surface states (TSSs) in which the spin of the electron is locked perpendicular to its momentum by strong spin-orbit interaction [6, 50, 51]. TIs have been identified as promising materials for exploiting exciting physics and developing potential applications in optoelectronics, spintronics, and quantum computations [5054]. Dirac fermions in TIs have also been studied by angle-resolved photoemission spectroscopy [5557] or scanning tunneling microscopy [58, 59]. In magnetotransport studies, TSS can be studied by weak antilocalization (WAL) [4, 9, 60, 61] and Shubnikov-de Haas oscillations [3, 62].

Topological insulator bismuth chalcogenide thin films have been grown epitaxially on various substrates using PLD. Onose et al. reported the epitaxial growth of Bi2Se3 thin films on InP (1 1 1) substrates (the lattice mismatch of 0.2%) [7]. A designed Se-rich target with an atomic ratio of Bi:Se = 2:8 was used to compensate for the issue of high doping carriers and to avoid unwanted Se-deficient phases. The pulsed laser power and repetition were 140 mJ and 10–20 Hz, respectively. The Bi2Se3 films obtained a small full-width at half-maximum (FWHM) for the XRD rocking curve of (0 0 0 6) peak. The surfaces of the films are composed of triangular pyramids with step-and-terrace structures, reflecting the hexagonal symmetry of Bi2Se3. The epitaxial relationship is Bi2Se3 (0 0 1) || InP (1 1 1) and Bi2Se3 [112¯0]||InP[001¯]. Le et al. reported the epitaxial growth of Bi2Se3 films on SrTiO3 (STO) (1 1 1) substrates using PLD at TS of 300 and 350°C [63]. The PLD conditions were the pulse fluence of 3.7 J/cm2, helium pressure of 40 Pa (300 mTorr), and repetition rate of 2 Hz. The laser source was KrF excimer laser (λ = 248 nm, duration 20 ns). By comparing the Bi2Se3 {0 1 5} and STO {2 0 0} diffraction peaks, the epitaxial relationship between the film and substrate was determined to be Bi2Se3 (0 0 1) || STO (1 1 1) and Bi2Se3 [1 1 0] || STO [110] [63].

Figure 10 presents the PLD epitaxial growths of Bi2Te3, Bi2Se3, and Bi3Se2Te thin films on large-misfit substrates [9, 53, 64]. The PLD conditions for growing Bi2Te3 films on STO (1 0 0) were as follows: substrate temperature of 300°C; helium ambient pressure of 40 Pa; repetition rate of 2 Hz; pulsed fluence of approximately 3.4 J/cm2. As shown in Figure 10a, a φ-scan was conducted on the (0 1 5) plane of a 200-nm-thick Bi2Te3 film and the (1 1 1) plane of an STO (1 0 0) substrate in skew symmetric geometry by tilting the samples. The in-plane orientation of a hexagonal h-Bi2Te3/STO (1 0 0) film displayed a 12-fold symmetry instead of the expected six-fold symmetry of the (0 1 5) plane in Bi2Te3. Figure 10b shows a schematic drawing of the in-plane atomic arrangement between an h-Bi2Te3 film and an STO (1 0 0) substrate. Since the principal crystallographic orientations of h-Bi2Te3 films grown on STO (1 0 0) substrates can be aligned along either the STO [1 0 0] or STO [0 1 0] directions, the in-plane arrangements result in an observed 12-fold symmetry. The angle differences between STO [0 1 0] and the two orientations of h-Bi2Te3 [1 1 0] were 30° and 0°, respectively, as shown in Figure 10b. In other words, the in-plane relationships were Bi2Te3 [1 1 0] || STO [0 1 0] and Bi2Te3 [1 0 0] || STO [1 0 0]. Lee et al. reported epitaxial growth via domain matching epitaxy of Bi2Se3 thin films on Al2O3 (0 0 0 1) substrates with over 13% lattice misfit and a critical thickness of less than one monolayer [64]. A relatively low repetition rate of 0.2 Hz and low TS of 250°C are key parameters of the PLD growth to achieving high-quality Bi2Se3 epitaxial films. Figure 10c shows φ-scan XRD results performed on { 011¯5 } planes of the film and { 011¯2 } planes of the substrate. Clearly, the presence of six peaks corresponding to the Bi2Se3 { 011¯5 } planes confirmed the epitaxy. A schematic depicting twin/domain growth and direction in the study is illustrated in Figure 10d [64]. In one case, the Bi2Se3 basal plane is aligned with that of Al2O3, and in the other, there is a rotation of 60°/180°. The epitaxial relationships are written as (0 0 0 1) Bi2Se3 || (0 0 0 1) Al2O3 (out-of-plane) and [21¯1¯0] Bi2Se3 || [21¯1¯0] Al2O3 or [21¯1¯0] Bi2Se3 || [112¯0] Al2O3 (in-plane). Figure 10e shows the typical φ-scan patterns of hexagonal h-Bi3Se2Te/h-Al2O3 (0 0 0 1) [9]. With the skew symmetric geometry, the φ-scan measurements were performed on the (1 1 6) plane of Al2O3 substrates and the (1 1 12) plane of Bi3Se2Te films. As shown in Figure 10e, the in-plane orientations of both Al2O3 (1 1 6) and Bi3Se2Te (1 1 12) exhibited six-fold symmetries with a 30° difference. Figure 10f illustrates the in-plane atomic arrangement between h-Bi3Se2Te and h-Al2O3 (0 0 0 1). The epitaxial relationship between the films and substrates is (0 0 0 1) Bi3Se2Te || (0 0 0 1) Al2O3 and [110] Bi3Se2Te || [210] Al2O3 [9]. This in-plane orientation was established to obtain the optimal lattice matching.

5.2. Magnetotransport properties of bismuth chalcogenide thin films

The weak antilocalization (WAL) which is a negative quantum correction to classical MR caused by the wave nature of electrons is used as a signature of TSS. In TIs, WAL is induced by both the helicity of the surface state and the spin-orbit coupling of bulk [4, 61, 65, 66]. In a low B field, the 2D WAL MR of a system with strong spin-orbit interaction can be described using the Hikami-Larkin-Nagaoka model [65], which is [4, 60, 61]

ΔR(B)[R(0)]2=αe22π2[ Ψ(12+BφB)ln(BφB) ]E8

where RW is the sheet resistance, ΔRW=RW(B)RW(0), Ψ(x) is the digamma function, Bϕ=/(4eLϕ2) is a magnetic field varying with the coherence length Lϕ, α is a parameter and reflects the number of conduction channels. In a 3D TI, α = −1/2 for a single coherent transport channel in the 2D surface states, and α = −1 for two independent coherent transport channels with similar Lφ in the 2D surface states [60, 65].

Figure 11.

Magnetoresistance (MR) results of the Bi2Te3, Bi2Se3, and Bi3Se2Te thin films grown by PLD [9, 64, 67]. (a and e) The MR (B = ±1 T) of a 27-nm-thick Bi2Te3 and a 200-nm-thick Bi3Se2Te films grown on Al2O3 (0 0 0 1) substrates. (b and f) Variation in the extracted electron dephasing length Lφ and parameter −α of the films as a function of temperatures. (c and d) MR of the epitaxial Bi2Se3 films grown on Al2O3 (0 0 0 1) substrates at TS = 250°C as a function of temperatures [64]. The solid green lines in (a) and (e) in low B are the theoretical predictions of 2D weak antilocalization (WAL) using Eq. (8).

The typical magnetoresistance (MR) results of some bismuth chalcogenides (i.e., Bi2Te3, Bi2Se3, and Bi3Se2Te) thin films grown by PLD are presented in Figure 11 [9, 64, 67]. The Bi2Te3, Bi2Se3, and Bi3Se2Te thin films were grown on Al2O3 (0 0 0 1) substrates using PLD at TS of 225, 250, and 250°C, respectively [9, 64, 67]. The WAL effect which presents as a sharp increase in resistance when B increases in the low magnetic field B regime is clearly observed on the films. Figure 11a, b, e and f show MR curves at several temperatures (T) and the extracted α(T) and Lϕ(T) values using Eq. (8) for the 27-QL (~27 nm)-thick Bi2Te3 and 200-nm-thick Bi3Se2Te thin films [9, 67]. At 2 K, Lφ are 158.1 and 195.2 nm, and Lφ decreases monotonically with increasing T, obeying the power laws, as Lφ ~ T−0.50 and Lφ ~ T−0.79 for the Bi2Te3 and Bi3Se2Te films, respectively (Figure 11b and f). Theoretically, the result of Lφ ~ T−0.50 observed in the 27-QL-thick Bi2Te3 thin film indicates the predominant electron-electron scattering in 2D weakly disordered systems. The Lφ ~ T−0.79 is closed to the Lφ ~ T−0.75 for 3D systems, if e-e scattering is the dominant dephasing source [9]. The electron screening effect in 3D system is more effective than that in 2D systems. However, the e-e scattering is strongly weakened with high carrier densities (n = 1020 cm−3 for the Bi3Se2Te films [9]). In 3D disordered conductors (namely, in the bulk state), dephasing by electron-phonon (e-ph) scattering would be significant and dominant. The e-ph scattering also causes a faster decay rate on Lφ, that is, Lφ ~ T−1.0 [68]. Additionally, the combination of 2D e-e scattering in the TSS thin layer and e-ph scattering in the bulk would further result in the Lφ ~ T−0.79. Consequently, e-ph scattering could be the force of channel separation as temperatures increase. It is worthy of mentioning that the MR result is a signature (not conclusive evidence) for the presence of TSS on Bi3Se2Te films. Further theoretical calculations and experiments are needed for reaching a final conclusion.

Figure 11b and f also present the –α(T) results of the Bi2Te3 and Bi3Se2Te films. The –α values of the Bi2Te3 film increase with increasing T from 0.43 at 2 K to 0.45 at 10 K, indicating the existence of a single coherent transport channel (i.e., likely a single surface state) [61]. Meanwhile, the –α values of the 200-nm-thick Bi3Se2Te film increase with increasing T from 0.5 at 2 K to 0.85 at 10 K and to 0.71 at 15 K, suggesting increased channel separation with T [69]. In WAL, the independent phase-coherent channels occur when the carriers in one channel lose phase coherence before being scattered into the other channel.

In Figure 11c, the MR (B,T) of a 50-nm-thick Bi2Se3 film was measured with B perpendicular to the film plane and ranging from −9 T to +9 T. At low B and T, the distinctive dips of WAL are clearly observed. The WAL effect results from strong spin-orbit coupling, showing the absence of backscattering giving rise to the destructive interference between the two time reversal symmetry loops when there is no magnetic field [64]. Resistance increases sharply with increasing B because the quantum interference is destroyed and backscattering increases (Figure 11c). Figure 11d is the enlarged MR at low B (–2 T to +2 T), and it reveals the existence of WAL as a function of T in a discernible way [64]. WAL gradually weakens as temperature increases, eventually disappearing entirely at T = 48 K; thus, the dependence on B is quadratic-like at low field. The MR cusp feature at low B is broadened and finally disappears with increasing temperature owing to the decrease in the phase coherence length. In addition to the WAL effect, Figure 11c shows a 2D, non-saturating linear MR at high B, which usually occurs with several TI materials of Bi2Se3 [70], Bi2Te3 [71], Bi2Te2Se [72], and Bi3Se2Te [9]. Theoretical models propose that the linear MR can appear in the gapless linear-dispersive energy spectrum when only the first landau level is filled [73, 74], or in the presence of both the gapless linear spectrum and Landau level overlaps [75]. Noticeably, the WAL and linear MR simultaneously reflect the 3D contribution of spin-orbit coupling in bulk and the Dirac nature of the 2D surface states. Because the magneto transport is a bulk sensitive measurement, it remains a major challenge to directly probe the topological nature [64].

Figure 12.

(a) Temperature-dependent normalized ab resistivities (ρ/ρ300 K) between 1.8 and 300 K of 46- and 200-nm-thick Bi2Te3 films. Upper inset: an EDS mapping image of a typical Bi-rich cluster. Lower inset: zoomed-in view of the ρ/ρ300 K in the low temperature range. (b) ρ(T) in 1.75–6.0 K of the 200 nm film at various H||c from 0 to 1 T. Inset: the onset Tc of the two superconducting transitions as a function of magnetic field. (c) Auger electron spectroscopy (AES) elemental depth profiling of a non-superconducting (46-nm-thick) and a superconducting (200-nm-thick) Bi2Te3 films. (d and e) The size distribution of Bi-rich clusters and Bi nanoclusters inside the clusters. (f) Schematics of the surface characteristics and a suggested superconducting mechanism in the Bi2Te3 films [53].

5.3. Proximity-induced superconductivities in Bi inclusions/bismuth chalcogenide thin films

Recent studies have shown a two-dimensional interface state between TIs and superconductors resulting from the superconducting proximity effect that supports Majorana fermions [76, 77]. Majorana fermions, novel particles which are their own antiparticles, can potentially be applied to topological quantum computing, which has motivated intense interest in TIs [53]. Koren et al. observed the local superconductivity in Bi2Te2Se and Bi2Se3 films below 2–3 K, which was naturally induced by small amounts of superconducting Bi inclusions or precipitations on the surface [78]. Moreover, Le et al. reported superconductivity at an onset critical temperature of approximately 3.1 K in a topological insulator 200-nm-thick Bi2Te3 thin film grown by pulsed laser deposition [53]. Indeed, Figure 12a shows the normalized resistivity ρ/ρ300 K of a 46-nm-thick Bi2Te3 film (S1) and a 200-nm-thick Bi2Te3 film (S2) as functions of temperatures (T) from 1.8 to 300 K. Both films show a decrease in resistivity (ρ) with decreasing T in the range of 20–300 K, implying that the films exhibit weak metallic properties commonly seen in narrow band-gap semiconductors with high carrier concentrations [53]. Below 20 K (Figure 12a, the lower inset), the ρ/ρ300 K of S1 shows a gentle upturn because of the weak localization of electrons [7], whereas the ρ/ρ300 K of S2 reaches a plateau before dropping slightly at Tc1 ≈ 5.8 K and then sharply at Tc2 ≈ 3.1 K. Figure 12b further shows the H||c-dependent ρ(T) of S2 in the low T regime, where H||c is the applied magnetic field along the c-axis of the film. At H||c = 0, ρ drops abruptly by 8% below Tc2onset but does not go down to zero, even at T = 1.8 K. This nonzero ρ at low T indicates that the superconducting volume ratio is not 100%. The inset in Figure 12b shows that the Tc2onset(Tc1onset) decreases from 3.1 (5.8) to 1.8 (5.4) K with increasing H||c from 0 to 0.2 T, strongly indicating that both transitions are superconducting in nature.

The detailed investigations of S2 strongly suggest the existence of superconducting Bi nanoclusters on the surface that induce the Tc1 ~ 5.8 K. EDS lateral elemental mapping revealed that the distributions of Te and Bi were not uniform, and many Bi-rich (47–54 at.%) clusters were visible (green color), as shown in the upper inset in Figure 12a, differing substantially from the uniform distribution and cluster-free surface observed in film S1. The size distribution and the most probable size of Bi-rich clusters are in the range of 400–2400 nm and 560–772 nm, respectively. Intriguingly, a closer inspection reveals that Bi-rich clusters are composed by some Bi nanoclusters (or nanograins) with a size of 20–62 nm.

The Bi-rich environment on the film surface is confirmed by AES analysis (Figure 12c) [53]. This is because the vapor pressure of Te (at TS = 300°C) is approximately 105 times higher than that of Bi, and therefore, more Te atoms are re-evaporated from the 300°C substrates [79]. The loss of Te is more severe in film S2 than in film S1 (14.3% in S2 and 4.5% in S1 at depth Z = 0, Figure 12c) because of the six times longer deposition time of S2 (60 min) compared to S1 (10 min). In addition, the nonstoichiometric effect is strongly depth-dependent (Figure 12c). The Te/Bi ratio gradually increases toward the stoichiometric ratio of 3/2 in 200-nm-thick films or slightly exceeds it (Te-rich) in 46-nm-thick films when the depth (Z) of the films increases. Under such a sufficiently high surface concentration of Bi atoms, the Bi clusters precipitate and segregate readily on the S2 surface to minimize overall free energy, as long as the substrate temperature of 300°C is higher than the 271°C melting point of Bi, as demonstrated in Figure 12f. Notably, the Bi clusters can only be observed in highly Bi-rich (14.3% at Z = 0) films (S2) and not in low Bi-rich (4.5% at Z = 0) films (S1), suggesting a critical Bi-rich concentration for Bi precipitation (separating a Bi phase) in a Bi2Te3 film [53]. The Tc1 at ~5.8 K found in our samples should be induced by the superconducting transition of the Bi nanoclusters, which is closely consistent with the Tc of 6.3 K for the surface Bi islands observed in Bi2Te2Se films [78]. The tiny resistivity drop at Tc1 = 5.8 K (by approximately 0.5%, Figure 12b) indicates that the amount of superconducting Bi nanoclusters in S2 is likely small and, therefore, the Josephson coupling between these islands is extremely weak. Since the superconductivity of Bi nanoclusters survived until H||c = 1.0 T (Figure 12b), the critical field of Bi nanoclusters is in between 0.3 and 1.0 T. This section demonstrates that natural defects generated during PLD growths, namely superconducting Bi nanoclusters or Bi inclusions, can substantially induce non-superconducting TI thin films (i.e., Bi2Te3, Bi2Se3, and Bi2Te2Se) into superconducting states at low temperatures.

Advertisement

6. Conclusion

This chapter provides the effects of ambient pressures and substrate temperatures in PLD growths on the structural-morphology, thermoelectric, nanomechanical, and magnetoresistance properties of bismuth chalcogenide thin films. The thermoelectric power factor of the stoichiometric Bi2Te3 films grown in the range of 220–340°C and PAr of 80 Pa reached remarkably high values, ranging between 18.2 ± 0.25 and 24.3 ± 0.44 μW cm−1 K−2. The optimal PF values were 5.54 ± 0.34 μW cm−1 K−2 for the layered hexagonal platelet Bi2Se3 films deposited at 300°C and PHe of 40 Pa and 8.3 μW cm−1 K−2 for the nanocrystalline Bi3Se2Te films deposited at 250°C and PHe of 40 Pa. We also reported the effects of PHe in PLD on nanomechanical properties (i.e., hardness and Young’s modulus) of Bi2Te3 and Bi3Se2Te thin films. It was observed that the hardness and Young’s modulus increased with increasing PHe, depending on the grain sizes following the inverse Hall-Petch effect for Bi2Te3 films grown at PHe ≤ 2.0 × 10−3 Torr and following the Hall-Petch relationship for Bi3Se2Te grown at PHe of 2.0 × 10−5 to 6.5 × 10−1 Torr. PLD has been successfully employed to grow epitaxially bismuth chalcogenide thin films on large-misfit substrates, for example, Bi2Te3/SrTiO3 (1 0 0), Bi2Se3/Al2O3 (0 0 0 1), and Bi3Se2Te/Al2O3 (0 0 0 1). The magnetotransport studies show that the bismuth chalcogenide thin films such as Bi2Te3, Bi2Se3, and Bi3Se2Te films present a two-dimensional weak antilocalization effect in a low magnetic field (B) regime and linear magnetoresistance in a high B regime, which could be attributed to the topological insulator surface states. Furthermore, proximity-induced superconductivities in Bi2Te3 thin films have an onset Tc of approximately 3.1 K, evidently induced by Bi inclusions (nanoclusters with onset Tc at 5.8 K) segregated on the surface of films.

Advertisement

Acknowledgments

Financial support from Vietnam National Foundation for Science and Technology Development (NAFOSTED) under Grant number 103.99–2015.17, the Ministry of Science and Technology, Taiwan under Contract Nos.: 103-2923-M-009-001-MY3, 103-2628-M-009-002-MY3, 103-2119-M-009-004-MY3, and the Ministry of Education (MOE-ATU plan at National Chiao Tung University) are gratefully acknowledged.

References

  1. 1. 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
  2. 2. Snyder GJ, Toberer ES. Complex thermoelectric materials. Nat. Mater. 2008; 7: 105–114. doi:10.1038/nmat2090
  3. 3. Qu D-X, Hor YS, Xiong J, Cava RJ, Ong NP. Quantum oscillations and hall anomaly of surface states in the topological insulator Bi2Te3. Science. 2010; 329: 821–824. doi:10.1126/science.1189792
  4. 4. Bao L, He L, Meyer N, Kou X, Zhang P, Chen Z, Fedorov AV, Zou J, Riedemann TM, Lograsso TA, Wang KL, Tuttle G, Xiu F. Weak anti-localization and quantum oscillations of surface states in topological insulator Bi2Se2Te. Sci. Rep. 2012; 2: 726 (1–7). doi:10.1038/srep00726
  5. 5. Hsieh D, Qian D, Wray L, Xia Y, Hor YS, Cava RJ, Hasan MZ. A topological Dirac insulator in a quantum spin Hall phase. Nature. 2008; 452: 970–974. doi:10.1038/nature06843
  6. 6. Moore JE. The birth of topological insulators. Nature. 2010; 464: 194–198. doi:10.1038/nature08916
  7. 7. Onose Y, Yoshimi R, Tsukazaki A, Yuan H, Hidaka T, Iwasa Y, Kawasaki M, Tokura Y. Pulsed laser deposition and ionic liquid gate control of epitaxial Bi2Se3 thin films. Appl. Phys. Express. 2011; 4: 083001. doi:10.1143/APEX.4.083001
  8. 8. Zhang HB, Yu HL, Bao DH, Li SW, Wang CX, Yang GW. Weak localization bulk state in a topological insulator Bi2Te3 film. Phys. Rev. B. 2012; 86: 075102. doi:10.1103/PhysRevB.86.075102
  9. 9. Le PH, Chiu S-P, Jian S-R, Luo CW, Lin J-Y, Lin J-J, Wu KH, Gospodinov M. Nanomechanical, structural, and transport properties of Bi3Se2Te thin films. J. Alloys Compd. 2016; 679: 350–357. doi:10.1016/j.jallcom.2016.03.226
  10. 10. Kaiser N. Review of the fundamentals of thin-film growth. Appl. Opt. 2002; 41: 3053–3060. doi:10.1364/AO.41.003053
  11. 11. Rowe DM. Thermoelectrics handbook: macro to nano. In: Rowe DM (Ed) Boca Raton: CRC Press; 2006. 1014 p. doi:10.1201/9781420038903
  12. 12. Thornton JA. Influence of apparatus geometry and deposition conditions on the structure and topography of thick sputtered coatings. J. Vac. Sci. Technol. 1974; 11: 666–670. doi:10.1116/1.1312732
  13. 13. Karabacak T, Singh JP, Zhao Y-P, Wang G-C, Lu T-M. Scaling during shadowing growth of isolated nanocolumns. Phys. Rev. B. 2003; 68: 125408. doi:10.1103/PhysRevB.68.125408
  14. 14. 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; 285: 657–663. doi:10.1016/j.apsusc.2013.08.107
  15. 15. 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
  16. 16. 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
  17. 17. Ohring M. The materials science of thin films. 2nd ed. San Diego: Academic Press; 2001. 794 p. ISBN: 978-0-12-524975-1
  18. 18. Tritt TM. Thermoelectric phenomena, materials, and applications. Annu. Rev. Mater. Res. 2011; 41: 433–448. doi:10.1146/annurev-matsci-062910-100453
  19. 19. Lind H, Lidin S, Häussermann U. Structure and bonding properties of (Bi2Se3)m(Bi2)n stacks by first-principles density functional theory. Phys. Rev. B. 2005; 72: 184101. doi:10.1103/PhysRevB.72.184101
  20. 20. Bos JWG, Zandbergen HW, Lee M-H, Ong NP, Cava RJ. Structures and thermoelectric properties of the infinitely adaptive series (Bi2)m(Bi2Te3)n. Phys. Rev. B. 2007; 75: 195203. doi:10.1103/PhysRevB.75.195203
  21. 21. Bell LE. Cooling, heating, generating power, and recovering waste heat with thermoelectric systems. Science. 2008; 321: 1457–1461. doi:10.1126/science.1158899
  22. 22. Nolas GS, Sharp J, Goldsmid HJ. Thermoelectrics basic principles and new materials developments. 1st ed. Berlin: Springer; 2001. 293 p. doi:10.1007/978-3-662-04569-5
  23. 23. 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
  24. 24. 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
  25. 25. Pichanusakorn P, Bandaru P. Nanostructured thermoelectrics. Mater. Sci. Eng. R. 2010; 67: 19–63. doi:10.1016/j.mser.2009.10.001
  26. 26. 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
  27. 27. 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
  28. 28. He L, Xiu F, Wang Y, Fedorov AV, Huang G, Kou X, Lang M, Beyermann WP, Zou J, Wang KL. Epitaxial growth of Bi2Se3 topological insulator thin films on Si (1 1 1). J. Appl. Phys. 2011; 109: 103702. doi:10.1063/1.3585673
  29. 29. Faleev SV, Léonard F. Theory of enhancement of thermoelectric properties of materials with nanoinclusions. Phys. Rev. B. 2008; 77: 214304. doi:10.1103/PhysRevB.77.214304
  30. 30. Zebarjadi M, Esfarjani K, Shakouri A, Bian Z, Bahk JH, Zeng G, Bowers J, Lu H, Zide J, Gossard A. Effect of nanoparticles on electron and thermoelectric transport. J. Electron. Mater. 2009; 38: 954–959. doi:10.1007/s11664-008-0656-4
  31. 31. Sumithra S, Takas NJ, Misra DK, Nolting WM, Poudeu PFP, Stokes KL. Enhancement in thermoelectric figure of merit in nanostructured Bi2Te3 with semimetal nanoinclusions. Adv. Energy Mater. 2011; 1: 1141–1147. doi:10.1002/aenm.201100338
  32. 32. 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
  33. 33. 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
  34. 34. 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
  35. 35. 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
  36. 36. 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
  37. 37. 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/c2jm34608g
  38. 38. 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
  39. 39. 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
  40. 40. Tasi C-H, Tseng Y-C, Jian S-R, Liao Y-Y, Lin C-M, Yang P-F, Chen D-L, Chen H-J, Luo C-W, Juang J-Y. Nanomechanical properties of Bi2Te3 thin films by nanoindentation. J. Alloys Compd. 2015; 619: 834–838. doi:10.1016/j.jallcom.2014.09.028
  41. 41. Li X, Bhushan B. A review of nanoindentation continuous stiffness measurement technique and its applications. Mater. Charact. 2002; 48: 11–36. doi:10.1016/S1044-5803(02)00192-4
  42. 42. Oliver WC, Pharr GM. An improved technique for determining hardness and elastic modulus using load and displacement sensing indentation experiments. J. Mater. Res. 1992; 7: 1564–1583. doi:10.1557/JMR.1992.1564
  43. 43. Schiøtz J, Vegge T, Di Tolla FD, Jacobsen KW. Atomic-scale simulations of the mechanical deformation of nanocrystalline metals. Phys. Rev. B. 1999; 60: 11971–11983. doi:10.1103/PhysRevB.60.11971
  44. 44. Zhang S, Sun D, Fu Y, Du H. Recent advances of superhard nanocomposite coatings: a review. Surf. Coatings Technol. 2003; 167: 113–119. doi:10.1016/S0257-8972(02)00903-9
  45. 45. Masumura RA, Hazzledine PM, Pande CS. Yield stress of fine grained materials. Acta Mater. 1998; 46: 4527–4534. doi:10.1016/S1359-6454(98)00150-5
  46. 46. Greer JR, De Hosson JTM. Plasticity in small-sized metallic systems: intrinsic versus extrinsic size effect. Prog. Mater. Sci. 2011; 56: 654–724. doi:10.1016/j.pmatsci.2011.01.005
  47. 47. Chang SY, Chang T-K. Grain size effect on nanomechanical properties and deformation behavior of copper under nanoindentation test. J. Appl. Phys. 2007; 101: 033507. doi:10.1063/1.2432873
  48. 48. Kumar KS, Suresh S, Chisholm MF, Horton JA, Wang P. Deformation of electrodeposited nanocrystalline nickel. Acta Mater. 2003; 51: 387–405. doi:10.1016/S1359-6454(02)00421-4
  49. 49. Conrad H, Narayan J. On the grain size softening in nanocrystalline materials. Scr. Mater. 2000; 42: 1025–1030. doi:10.1016/S1359-6462(00)00320-1
  50. 50. Hasan MZ, Kane CL. Colloquium: topological insulators. Rev. Mod. Phys. 2010; 82: 3045–3067. doi:10.1103/RevModPhys.82.3045
  51. 51. Hasan MZ, Moore JE. Three-dimensional topological insulators. Annu. Rev. Condens. Matter Phys. 2011; 2: 55–78. doi:10.1146/annurev-conmatphys-062910-140432
  52. 52. Luo CW, Wang HJ, Ku SA, Chen H-J, Yeh TT, Lin J-Y, Wu KH, Juang JY, Young BL, Kobayashi T, Cheng C-M, Chen CH, Tsuei KD, Sankar R, Chou FC, Kokh KA, Tereshchenko OE, Chulkov EV, Andreev YM, Gu GD. Snapshots of Dirac fermions near the Dirac point in topological insulators. Nano Lett. 2013; 13: 5797–802. doi:10.1021/nl4021842
  53. 53. Le PH, Tzeng W-Y, Chen H-J, Luo CW, Lin J-Y, Leu J. Superconductivity in textured Bi clusters/Bi2Te3 films. APL Mater. 2014; 2: 096105. doi:10.1063/1.4894779
  54. 54. Tu C-M, Yeh T-T, Tzeng W-Y, Chen Y-R, Chen H-J, Ku S-A, Luo C-W, Lin J-Y, Wu K-H, Juang J-Y, Kobayashi T, Cheng C-M, Tsuei K-D, Berger H, Sankar R, Chou F-C. Manifestation of a second Dirac surface state and bulk bands in THz radiation from topological insulators. Sci. Rep. 2015; 5: 14128. doi:10.1038/srep14128
  55. 55. Chen YL, Analytis JG, Chu J-H, Liu ZK, Mo S-K, Qi XL, Zhang HJ, Lu DH, Dai X, Fang Z, Zhang SC, Fisher IR, Hussain Z, Shen Z-X. Experimental realization of a three-dimensional topological insulator, Bi2Te3. Science. 2009; 325: 178–181. doi:10.1126/science.1173034
  56. 56. Hsieh D, Xia Y, Qian D, Wray L, Meier F, Dil JH, Osterwalder J, Patthey L, Fedorov AV, Lin H, Bansil A, Grauer D, Hor YS, Cava RJ, Hasan MZ. Observation of time-reversal-protected single-Dirac-cone topological-insulator states in Bi2Te3 and Sb2Te3. Phys. Rev. Lett. 2009; 103: 146401. doi:10.1103/PhysRevLett.103.146401
  57. 57. Wang G, Zhu X-G, Sun Y-Y, Li Y-Y, Zhang T, Wen J, Chen X, He K, Wang L-L, Ma X-C, Jia J-F, Zhang SB, Xue Q-K. Topological insulator thin films of Bi2Te3 with controlled electronic structure. Adv. Mater. 2011; 23: 2929–2932. doi:10.1002/adma.201100678
  58. 58. Zhang T, Cheng P, Chen X, Jia J-F, Ma X, He K, Wang L, Zhang H, Dai X, Fang Z, Xie X, Xue Q-K. Experimental demonstration of topological surface states protected by time-reversal symmetry. Phys. Rev. Lett. 2009; 103: 266803. doi:10.1103/PhysRevLett.103.266803
  59. 59. Alpichshev Z, Analytis JG, Chu J-H, Fisher IR, Chen YL, Shen ZX, Fang A, Kapitulnik A. STM imaging of electronic waves on the surface of Bi2Te3: topologically protected surface states and hexagonal warping effects. Phys. Rev. Lett. 2010; 104: 016401. doi:10.1103/PhysRevLett.104.016401
  60. 60. Chiu SP, Lin JJ. Weak antilocalization in topological insulator Bi2Te3 microflakes. Phys. Rev. B. 2013; 87: 035122. doi:10.1103/PhysRevB.87.035122
  61. 61. Cha JJ, Kong D, Hong S-S, Analytis JG, Lai K, Cui Y. Weak antilocalization in Bi2(SexTe1−x)3 nanoribbons and nanoplates. Nano Lett. 2012; 12: 1107−1111. doi:10.1021/nl300018j
  62. 62. Yu X, He L, Lang M, Jiang W, Xiu F, Liao Z, Wang Y, Kou X, Zhang P, Tang J, Huang G, Zou J, Wang KL. Separation of top and bottom surface conduction in Bi2Te3 thin films. Nanotechnology. 2013; 24: 015705. doi:10.1088/0957-4484/24/1/015705
  63. 63. Le PH, Wu KH, Luo CW, Leu J. Growth and characterization of topological insulator Bi2Se3 thin films on SrTiO3 using pulsed laser deposition. Thin Solid Films. 2013; 534: 659–665. doi:10.1016/j.tsf.2013.01.104
  64. 64. Lee YF, Punugupati S, Wu F, Jin Z, Narayan J, Schwartz J. Evidence for topological surface states in epitaxial Bi2Se3 thin film grown by pulsed laser deposition through magneto-transport measurements. Curr. Opin. Solid State Mater. Sci. 2014; 18: 279–285. doi:10.1016/j.cossms.2014.07.001
  65. 65. Hikami S, Larkin AI, Nagaoka Y. Spin-orbit interaction and magnetoresistance in the two dimensional random system. Prog. Theor. Phys. 1980; 63: 707–710. doi:10.1143/PTP.63.707
  66. 66. Suzuura H, Ando T. Crossover from symplectic to orthogonal class in a two-dimensional honeycomb lattice. Phys. Rev. Lett. 2002; 89: 266603. doi:10.1103/PhysRevLett.89.266603
  67. 67. Le PH, Liu P-T, Luo CW, Lin J-Y, Wu KH. Thickness-dependent magnetotransport properties and terahertz response of topological insulator Bi2Te3 thin films. J. Alloys Compd. doi:10.1016/j.jallcom.2016.09.109
  68. 68. Lin JJ, Bird JP. Recent experimental studies of electron dephasing in metal and semiconductor mesoscopic structures. J. Phys. Condens. Matter. 2002; 14: R501–596. doi:10.1088/0953-8984/14/18/201
  69. 69. Steinberg H, Laloё J-B, Fatemi V, Moodera JS, Jarillo-Herrero P. Electrically tunable surface-to-bulk coherent coupling in topological insulator thin films. Phys. Rev. B. 2011; 84: 233101. doi:10.1103/PhysRevB.84.233101
  70. 70. Tang H, Liang D, Qiu RLJ, Gao XPA. Two-dimensional transport-induced linear magnetoresistance in topological insulator Bi2Se3 nanoribbons. ACS Nano. 2011; 5: 7510–7516. doi:10.1021/nn2024607
  71. 71. Wang X, Du Y, Dou S, Zhang C. Room temperature giant and linear magnetoresistance in topological insulator Bi2Te3 nanosheets. Phys. Rev. Lett. 2012; 108: 266806. doi:10.1103/PhysRevLett.108.266806
  72. 72. Assaf BA, Cardinal T, Wei P, Katmis F, Moodera JS, Heiman D. Linear magnetoresistance in topological insulator thin films: quantum phase coherence effects at high temperatures. Appl. Phys. Lett. 2013; 102: 012102. doi:10.1063/1.4773207
  73. 73. Abrikosov AA. Quantum magnetoresistance. Phys. Rev. B. 1998; 58: 2788–2794. doi:10.1103/PhysRevB.58.2788
  74. 74. Abrikosov AA. Quantum linear magnetoresistance; solution of an old mystery. J. Phys. A Math. Gen. 2003; 36: 9119–9131. doi:10.1088/0305-4470/36/35/301
  75. 75. Wang CM, Lei XL. Linear magnetoresistance on the topological surface. Phys. Rev. B. 2012; 86: 035442. doi:10.1103/PhysRevB.86.035442
  76. 76. Qi X-L, Hughes TL, Raghu S, Zhang S-C. Time-reversal-invariant topological superconductors and superfluids in two and three dimensions. Phys. Rev. Lett. 2009; 102: 187001. doi:10.1103/PhysRevLett.102.187001
  77. 77. Wilczek F. Majorana returns. Nat. Phys. 2009; 5: 614–618. doi:10.1038/nphys1380
  78. 78. Koren G, Kirzhner T, Lahoud E, Chashka KB, Kanigel A. Proximity-induced superconductivity in topological Bi2Te2Se and Bi2Se3 films: robust zero-energy bound state possibly due to Majorana fermions. Phys. Rev. B. 2011; 84: 224521. doi:10.1103/PhysRevB.84.224521
  79. 79. 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

Written By

Phuoc Huu Le and Chih Wei Luo

Submitted: 22 March 2016 Reviewed: 21 September 2016 Published: 21 December 2016