Open access peer-reviewed chapter

Effects of Dimensionality Reduction for High-Efficiency Mg-Based Thermoelectrics

Written By

Övgü Ceyda Yelgel

Submitted: 03 January 2023 Reviewed: 27 January 2023 Published: 28 February 2023

DOI: 10.5772/intechopen.110239

From the Edited Volume

Magnesium Alloys - Processing, Potential and Applications

Edited by Tomasz Arkadiusz Tański, Katarzyna Cesarz-Andraczke and Ewa Jonda

Chapter metrics overview

60 Chapter Downloads

View Full Metrics

Abstract

Over the past decade, there has been significant interest in the field of thermoelectric materials (TEs) owing to their use in clean and sustainable energy sources for cooling and/or power generation applications. Especially, Mg2XIV (XIV = Si, Ge, Sn) based TEs are promising candidates for middle-temperature range energy conversion due to their high thermoelectric performance, environmentally harmless, abundant raw materials, non-toxicity, and relatively inexpensive cost of modules. In this book chapter, we present an overview of the theoretical background of the thermoelectric transport properties (Seebeck coefficient, electrical conductivity, thermal conductivity, and thermoelectric figure of merit ZT) of magnesium-based bulk and low dimensional systems (i.e., quantum wells and quantum wires). A detailed description of the temperature-dependent Fermi level both in extrinsic and intrinsic regimes will be provided whereby it is the primary step in deriving the thermoelectric transport parameters of materials. Following the linearized Boltzmann transport equations temperature-dependent electronic transport properties (Seebeck coefficient, electrical conductivity, and electronic thermal conductivity) of materials under the energy-dependent relaxation time approximation will be defined. By employing Debye’s isotropic continuum model within the single mode relaxation time approximation including various phonon relaxation rates contributed by different scattering mechanisms the lattice contribution to the thermal conductivity will be included.

Keywords

  • thermoelectric figure of merit
  • Mg-based thermoelectrics
  • thermoelectric transport properties
  • Boltzmann transport equation
  • thermal conductivity
  • phonon scattering mechanisms

1. Introduction

Due to the increasing world’s demand for high-efficient clean energy researchers have been looking for environmentally friendly energy technologies. Non-renewable energy sources (nuclear power, coal, oil, and gas) have many disadvantages that can be listed as limited reserves, hazardous environmental risks, acid precipitation, and global climate change. Therefore, renewable energy sources determined as sourced either directly or indirectly from the Sun and solar energy have attracted much attention and scientific studies have advanced rapidly in renewable energy technologies. As known heat and electricity are two different types of energy and thermoelectricity bridges between them. Thermoelectric materials can directly convert thermal energy (such as waste heat generated by automotive exhaust, domestic heating, and industrial processes) into electrical energy (and vice versa); this property makes them noteworthy for the development of sustainable energy-efficient technologies. Compared with other energy conversion technologies, thermoelectric devices have promising advantages such as having no moving parts, noiseless operation, long service life, stability, and less maintenance. Besides, it needs to be noted that the only major disadvantage of thermoelectric devices is their low efficiency for everyday common use compared to other energy conversion technologies. As summarized in Figure 1, thermoelectric devices can be extensively used in many fields such as power generators, cooling devices, sensors, space exploration, solar-based driven thermoelectric technologies, waste heat recovery from automobile engines, and energy harvesting and temperature control [1, 2, 3, 4, 5, 6].

Figure 1.

Overview of application fields of thermoelectric materials.

Different sectors contributing to huge amounts of waste heat energy (∼66%), the biggest energy problem in the world right now, could be transformed to electrical energy via thermoelectric materials with the help of their ability to convert large amounts of wasted thermal energy into useful energy. Figure 2a shows the utilization of waste heat energy from different types of waste energy sources by thermoelectric generator.

Figure 2.

(a) Utilization of waste heat energy from different types of waste energy sources by thermoelectric generator, and (b) scheme and operation of thermoelectric module working as a power generator.

Thermoelectric generators are built of thermoelectric modules and a single module consists of a thermocouple. In Figure 2b, an operation of thermoelectric module working as a power generator mode is given. A single thermocouple generates a small thermoelectric force, and therefore to the formation of one module is used even so many thermocouples in this manner when the thermal energy is supplied the heat is directed to the thermoelectric modules in which a conversion of thermal energy into electricity. As shown in Figure 3a, a thermoelectric couple composed of an n-type (electrons are carriers) and a p-type (holes are carriers) semiconductor material connected through metallic electrical contact pads. Both power generation (Seebeck effect) and refrigeration (Peltier effect) can be accomplished using the same module.

Figure 3.

(a) Operation principle of a power generator (Seebeck effect) and a refrigerator (Peltier effect), and (b) schematic presentation of carrier concentration dependence of electrical transport properties (Seebeck coefficient, electrical conductivity, carrier thermal conductivity) and thermal transport sourced from phonons.

The thermoelectric performance of the materials at a certain temperature is ranked by the dimensionless figure of merit ZT. Its definition is ZT=S2σκtotalT where T is the absolute temperature, S is the Seebeck coefficient, σ is the electrical conductivity, andκtotal is the total thermal conductivity (κtotal=κc+κbp+κph) which has contributions from carriers (i.e., from electrons for n-type materials or holes for p-type materials,κc), electron-hole pairs (bipolar,κbp), and phonons (κph), respectively. The material with a high ZT value has better thermoelectric energy conversion efficiency. Therefore, efficient thermoelectric materials need to have a large Seebeck coefficient to maximize the conversion of heat to electrical power; high electrical conductivity to minimize Joule heating; and low total thermal conductivity to retain heat at the junctions and maintain a large temperature gradient. However, as presented in Figure 3b, these thermoelectric transport properties are strongly coupled with each other and dependent on the material’s band and crystal structure parameters. Owing to the complex interdependence among these properties, it is quite challenging to optimize ZT value and for this reason, the vast majority of research in this field has been focused on this issue. A desirable goal for everyday practical usage of the thermoelectric device is expected to have their efficiency in the range ofZT>1.5 and ZT>2for power generation and thermoelectric cooling, respectively [7, 8, 9]. In the literature, until the 1990s, the highest value of ZT was found to be around 1 which means only 10% conversion efficiency [10, 11, 12]. In the last few years, on the other hand, the highest ZT values have been recorded with the help of using novel materials, following different processing techniques, and having new mechanisms [2, 13, 14, 15, 16]. There are various strategies to improve thermoelectric efficiency; maximizing thermoelectric power factor (PF=S2σ) by following different methods such as convergence of electronic band valleys, carrier energy filtering, and generating resonant levels in electronic bands [13, 15, 17, 18, 19, 20, 21, 22]; reducing total thermal conductivity with introducing new point defects by making alloys, embedding nano inclusion, anharmonicity, nano/mesoscale grain boundaries [2, 14, 23, 24, 25, 26, 27, 28]. Among all these methods we must admit that in the thermoelectric literature Hicks and Dresselhaus made a revolutionary invention by proposing the use of low-dimensionalizing in thermoelectric devices and opened a new pathway in high-efficiency thermoelectric technologies in the early 1990s [29]. In the low dimensionalization method, reduced sizes of materials to the order of a nanometer in one or two directions are used (such as superlattices, nanowires, or quantum dots). Following this strategy allows enhancing ZT to a very large extent by optimizing electronic and thermal properties of materials in ways that are not possible in bulk materials. Also, the thermoelectric efficiency of traditional three-dimensional bulk materials can be improved with the help of the low-dimensionalization method. Low dimensional thermoelectric materials with unusual high ZT values (which can easily break the limits of unity) have been reported by several groups [2, 13, 15, 17, 30, 31, 32, 33]. Moreover, there have been significant improvements in portable electronics and intelligent society based on the internet of things (IoTs), and the internet of nano things (IoNTs) however rechargeable lithium-ion batteries are inconvenient for applications. At this point, by generating electricity from the temperature difference between the human body and the environment, high-efficiency nano-based thermoelectric materials can be an ideal solution to replace lithium-ion batteries in IoTs and IoNTs [34, 35, 36, 37, 38, 39].

There are so many different thermoelectric materials however among all of them the materials working in the middle-temperature range (namely from 400 K to 800 K) are widely used for thermoelectric applications. Generally, in commercial practice Bi2Te3-, PbTe-, CoSb3-, and GeTe-based thermoelectric compounds are mainly used whereas they have toxic constituent elements and are scarce [40, 41, 42, 43]. The present market search for cheap, environmentally friendly, and non-toxic thermoelectric materials due to safety regulations. Compared to other middle-temperature thermoelectric materials Magnesium-based materials (Mg2X; X = Si, Ge, Sn) have been regarded as promising candidates for high efficiency and large-scale application owing to their high availability, low-cost, reasonable high efficiency, non-toxicity, and low mass density. In Figure 4, the highest ZT values of the state-of-the-art medium-to-high temperature thermoelectric materials are given to compare with Mg-based thermoelectric materials.

Figure 4.

High-efficient hermoelectric materials working in the middle-to-high temperature range. Reproduced from Ref. [40] with permission from the Royal Society of Chemistry.

Mg2X materials are intermetallic alloys that have been investigated for decades as high-efficiency thermoelectric candidates [44, 45]. Mg2X compounds crystallize in the cubic anti-fluorite structure (space group of Fm3¯m #225) phase with X in face-centered cubic position and Mg in tetrahedral sites with three atoms per primitive unit cell in that one X atom is located at a(0 0 0) and 2 Mg atoms are located at a(0.25, 0.25, 0.25) and a(0.75, 0.75, 0.75). The crystal structure model of Mg2X and schematic illustration of the band structures of Mg2X around Γ and X points are shown in Figure 5a and b, respectively.

Figure 5.

(a) Crystal structure model of Mg2X (X = Si, Ge, Sn), and (b) schematic illustration of Mg2X around Γ and X points.

In addition to that, in Table 1, electronic and physical properties are listed for Mg2X-based binary alloys.

Properties/ParametersMg2SiMg2GeMg2Sn
a (Å)6.3386.3856.765
Eg (eV) at 0 K0.770.740.35
ΔE (eV) at 0 K0.40.580.16
Melting point (K)137513881051
γ1.321.381.27
κph (W m K−1)7.96.65.9

Table 1.

List of electronic (lattice parameters, band gaps, and energy offsets), physical (melting points), and phononic properties (Grüneissen parameter, and phonon thermal conductivity) of Mg2X (X = Si, Ge, Sn) binary alloys [46, 47, 48].

The overview of the highest reported ZT values for Mg-based thermoelectric materials (in the form of binary, ternary, and quaternary n- and p-type doped) is illustrated in Figure 6.

Figure 6.

State-of-the-art ZT values for Mg-based thermoelectric materials from different works in the literature [49, 50, 51, 52, 53, 54, 55, 56, 57, 58, 59, 60, 61, 62, 63].

As stated above combining the advantages of using low-dimensionalization and Mg-based thermoelectric materials can lead to highly efficient thermoelectric applications. Despite this, in the literature, there have been very limited theoretical investigations of low-dimensional effects on thermoelectric transport properties of Mg-based materials in detail. Therefore, in this book chapter, unlike the related literature, a comprehensive explanation of the temperature-varied Fermi levels (both at extrinsic and intrinsic regimes) will first be presented for bulk and low-dimensional Mg-based materials. Then, by following the linearized Boltzmann transport equations temperature dependence of electronic transport properties (Seebeck coefficient, electrical conductivity, and carrier thermal conductivity) of bulk and low-dimensional Mg-based materials under the energy-dependent relaxation time approximation will be defined. Also, by employing Debye’s isotropic continuum model within the single mode relaxation time approximation including various phonon relaxation rates contributed by different scattering mechanisms the lattice thermal contribution will be explained meticulously. By presenting a detailed theoretical background of the dimensionality effect on thermoelectric efficiency and highlighting the temperature dependencies of transport properties of low-dimensional systems we shed light on the challenges and perspectives for promoting the development of future Mg-based thermoelectric applications.

Advertisement

2. Thermoelectric transport properties of bulk and low-dimensional systems

The theory of transport properties in thermoelectric materials deals with the temperature dependence of electrical and thermal transport properties. Here, we introduce the theoretical background of thermoelectric transport properties in bulk and low-dimensional systems, respectively.

As stated in the above section, the dimensionless figure of merit includes three interdependent transport properties of S, σ, and κtotal and which impedes the maximization of ZT in bulk systems. To avoid this problem, in 1993, Hicks and Dresselhaus examined the effects of using low-dimensional structures on the ZT with the assumption of parabolic bands and constant relaxation time in a one-band material [29]. To summarize Hicks and Dresselhaus’ approach and describe the effects of dimensionalization as a bird’s eye view, by defining a dimensional factor N (N = 1, 2, and 3 for 1 dimensional (1D), 2 dimensional (2D), and 3 dimensional (3D) systems, respectively) the dimensionality dependence of ZT can be written for the case of conduction along the x direction as [29, 64],

ZNT=N2SNFN/2FN/21ζ2FN/211BN+N+42FN/2+1βNFN/22FN/21E1

where ζ is the reduced chemical potential, SN=146N+N23, βN=349N+2N26, and the Fermi-Dirac function Fi is given by Fiζ=0xidxexpxζ+1. The material property dependent parameter BN is expressed as,

BN=γN2kBT2N/2kB2TμxeκphE2

where kB is the Boltzmann constant and is the reduced Planck’s constant, γ1=2πa2mx1/2 for 1D systems, γ2=12πamxmy1/2 for 2D systems, and γ3=13π2mxmymz1/2 for 3D systems, respectively. Here mx, my, and mz are the effective mass components for the x, y, and z directions, μx is the mobility in the x direction, and a is the width of a 2D quantum well or 1D nanowire. From these equations, we can conclude that ZT depends on ζ and BN parameters. For bulk materials, ZT can be improved by optimizing ζ with suitable doping and BN with reducing κph and/or increasing μx. From the above equations, it is seen that in low-dimensional systems, ZT can be enhanced by adjusting the width/thickness of wells and wires, respectively. Decreasing the width/thickness of the quantum well or nanowire gives rise to boundary scattering mechanisms which lead to a dramatic reduction in phonon thermal conductivity. In Figure 7, the quantum well thickness and quantum wire diameter dependences of ZT are represented for quantum wells and wires of Bi2Te3 fabricated along the x, y, and z directions [64].

Figure 7.

Enhanced ZT values as a function of (a) the quantum well thickness, and (b) the quantum wire diameter. Here quantum wells are fabricated in the x-y plane and the x-z plane while quantum wires are fabricated along the x, y, and z directions (data are read from Ref. [64]).

After emphasizing the importance of using low-dimensional materials briefly, we can now move on to examine the detailed transport properties of thermoelectric materials both in bulk and low-dimensional structures to understand the detailed physical background behind these temperature-dependent transport properties and carrier scattering mechanisms. Based on the semiclassical model we follow the Boltzmann transport equation approach to investigate the thermoelectric transport phenomena quantitatively and explain how the temperature variations of carriers and phonons. All the electronic transport properties of thermoelectric materials are dependent on temperature-varied Fermi levels (EF) in extrinsic and intrinsic regimes. Therefore, simpler schemes leading to expressions for the Fermi level in bulk semiconductors have been given in the book by McKelvey [65],

EFextntype=12Ec+Ed+kBT2lnNd2UckBTsinh1Uc8NdexpEi2kBTE3
EFextptype=12Ea+Ev+kBT2lnNa2UvkBTsinh1Uv8NdexpEi2kBTE4

where Ec is the conduction band edge, Ev is the valence band edge, Ed is the donor energy level, and Ea is the acceptor energy level, ΔEi is given as EcEd for the donor ionization and expressed as EaEv for the acceptor ionization energies. Also, Ndand Na are the concentrations of donor impurity and acceptor impurity atoms, respectively. Uc andUv parameters are given as Uc=2mnkBT/2π23/2 and Uv=2mpkBT/2π23/2 where mn is for the electron mass and mp is for the hole mass. At high enough temperatures donors at the Ed or acceptors at the Ea levels become ionized and the material turns out to be an intrinsic semiconductor. The temperature dependence of EF for both types of doped semiconductors is expressed as [65],

EFintnandptypes=12Ev+Ec+34kBTlnmpmn.E5

The temperature dependence of the Fermi level for semiconductors can be summarized as; in n-type doped semiconductors, the EF is in the middle of the Ed and Ec at 0 K, and when the temperature rises the EF first rises slowly and then decreases through the Ed and reaches the center of the gap; similarly, in p-type doped semiconductors the EF is in the middle of the Ea and Ev at 0 K and when the temperature rises the EF first goes down slightly and then increases through the Ea. For both types of doped semiconductors when the donor/acceptor levels are almost completely ionized the EF approaches to EFint. Moreover, the energy dispersion function at the Brillouin zone center is determined as

Ekxkykz=2kx22mx+2ky22my+2kz22mzE6

where kx, ky, and kz are the propagation vector component along thex,y, and z axes, respectively. The effective mass tensor components of the constant energy surfaces are denoted as mx, my, and mz.

In this book chapter, we consider low-dimensional systems as 2D quantum wells and 1D quantum wires and present the theoretical modeling of the temperature dependence of thermoelectric transport properties in these systems. First, in 2D quantum well systems, we assume that the system has B/A/B layers of material A (as a well material with nA atomic layers of thickness a each and total thickness dA) and material B (as a barrier with nB atomic layers of thickness b each and total thickness dB) also band gap differences materials A and B to confine the current carriers to the well material and system is grown in the z-direction. The electronic dispersion in the z-direction is given as,

Ekxky=2kx22mx+2ky22my+Ej,E7

where the quantum confinement term Ej in the jth quantum sublevel is expressed as

Ej=2π22mzdA2j.E8

Assuming the carrier concentration does not change whereas the well thickness varies and up to jth quantum sub-level is occupied by carriers the temperature variation of EF for 2D quantum wells can be defined as

Efext/int2DQWL=Efext/int3DBULKEj.E9

in both extrinsic and intrinsic regimes [66]. Moreover, the energy band gap of this 2D system is given as

Eg2DQWL=Eg3DBULK+2π22dA21mz,e+1mz,hE10

where Eg3DBULK=EcEv and mz,e and mz,h are electrons and holes effective masses in the growth direction, respectively.

In 1D quantum wire systems, we assume that carrier confinement in the y and z directions and free-electron-like motion in the x direction also a is the width of the wire considered to have a square cross-section. The electronic dispersion relation for the 1D system is given as,

Ekx=2kx22mx+Ei,j,E11

where the quantum confinement terms Ei,j in the ith and jth quantum sublevel are expressed as

Ei,j=2π22mydA2i+2π22mzdA2j,E12

with the width of the wire dA assumed to have a square cross-section. Similar to 2D quantum wells, for 1D quantum wire systems the temperature dependence of EF can be written as

Efext/int1DQWR=Efext/int3DBULKEi,jE13

both in extrinsic and intrinsic regimes. Additionally, the band gap of 1D systems is given as

Eg1DQWR=Eg3DBULK+2π22dA21my,e+1my,h+2π22dA21mz,e+1mz,h.E14

As shown in Figure 8, the carrier energy dependence of the density of states (DOS) per unit volume for a single bad with the subband in bulk, quantum well, and quantum wire systems by assuming transport is along the direction of the x-axis. Compared to the 3D system gE enhances at E0 owing to the rise of carriers both with confinement effect in 2D and 1D systems.

Figure 8.

The electronic density of states for 3D, 2D, and 1D systems in relation to carrier energy (E0 corresponds to the energy at the bottom of the conduction band).

The theoretical expressions of electronic transport properties for thermoelectric materials (namely the Seebeck coefficient, electrical conductivity, and electronic thermal conductivity) are explained by following the Boltzmann transport equation. Presenting the theoretical background of each transport coefficient can help to understand the physical origin of these properties in relation to temperature and carrier concentration for bulk, quantum well, and quantum wire systems.

Assuming an isotropic solid within the nearly free electron/hole model under the consideration of an electric field and a temperature gradient is applied the thermoelectric transport properties can be determined by solving the linearized Boltzmann equation. Here, considering both heat and electricity currents flow in the x-direction, the electric current density ix and the rate of flow of heat per unit cross-sectional area (carried by carriers) wx are given [65, 66, 67, 68, 69, 70].

ix=0evxfEgEdEE15
wx=0vxEEffEgEdEE16

where e is the electronic charge, vx is the velocity of charge carriers in thex direction for an isotropic conductor, fE is the Fermi distribution function, gE is the density of state function, and EEf represents the total energy transported by a carrier. At equilibrium, carrier distribution follows the Fermi Dirac statistics through f0E=expEEfkBT+11where Ef is the Fermi energy. The above equations can be re-written by replacing f with ff0 (since there is no transport ix=wx=0 when f=f0) as

ix=0evx2τgEf0EEfx+EEfTTxdEE17
wX=±Efeix+0Evx2gEτf0EEfx+EEfTTxdE.E18

Here, the assumptions are τx=μxmxe with μx is the carrier mobility in the x-direction and vx2=2E3mx. By taking the condition ix=0 as the Seebeck coefficient is written

S=VT=εxT=1eEf/xT/xE19

where the electric field Ex is given as εx=±1eEfx and S is expressed as

S=±1e0vx2τxgEf0EEEfTdE/0vx2τxgEf0EdEE20

here + sign and − sign represent p- and n-type semiconductors, respectively. The electrical conductivity expression in the extrinsic regime can be defined by setting the temperature gradient as T/x=0 and expressed as

σext=ixεx=0e2vx2τxgEf0EdE.E21

Additionally, in the intrinsic regime, the electrical conductivity of semiconductors is given by

σint=eNiμn+μp=AeEg/2kBTE22

where Ni is the equilibrium carrier density, A is a temperature-independent adjustable parameter, μn and μp is the electron and hole mobility, respectively. Lastly, the total thermal conductivity is the sum of three contributions from carriers (electrons or holes), bipolar (electron-hole pairs), and phonons. The electronic thermal conductivity is related to the electrical conductivity by the Wiedemann-Franz law as κc=LσT where L is the Lorenz number namely a relatively complex transport parameter. The carrier thermal conductivity comes from monopolar carriers (electrons or holes), is determined under the condition of ix=0, and is given as [71].

κc=wxT/x=0vx2EgEτxf0EEf/xT/x+EEfTdE.E23

By determining these integrals we can now write expressions for the temperature dependence of electronic transport properties of 3D, 2D, and 1D systems. As the dimension of a system changes the above integrals need to be varied by the electronic density of states. In Table 2, the expressions for temperature-varied electronic transport properties are listed for bulk, quantum well, and wire systems.

Table 2.

Temperature variation of electronic transport properties (Seebeck coefficient S, electrical conductivity in extrinsic σext and intrinsic σint regimes, and carrier thermal conductivity κc) of 3D, 2D, and 1D systems as a function of Fermi integrals described as Fi=0xidx/exζ+1 where energy (x) and the reduced Fermi energy (ζ) of carriers normalized to the thermal energy are x=E/kBT and ζ=Ef/kBT, respectively.

In Figure 9a–d, the electronic transport properties of bulk and quantum wire structures (with wire widths taken as dA=6,7,and8nm) of Mg3.2−yCoy(Sb0.3Bi0.7)1.99Te0.01 (y = 0.05) are presented in the temperature ranges between 300 and 850 K. Additionally the experimental measurements of the temperature dependence of See beck coefficient and electrical conductivity results for bulk system reported by Ref. [72] are included to compare our theoretical results with experiments. As seen from the Fermi level results in Figure 9a, both bulk and quantum wire systems show extrinsic and intrinsic regimes in the whole temperature range. While the critical extrinsic-to-intrinsic turnover temperature is around 500 K for bulk sample, the temperature shifts to around 600 K for nanowire systems. In Figure 9b, the temperature variations of Seebeck coefficients are illustrated for both 3D and 1D systems. From these results, it is clearly seen that by decreasing the quantum wire width the maximum value of |S| can be improved from 241.1 μV/K at 510 K upto 276.2 μV/K at 593 K by using dA=6nm quantum wire instead of bulk structure. The temperature dependence of electrical conductivity for all systems is seen in Figure 9c. From theoretical calculations, it is seen that a dramatic enhancement can be gained in electrical conductivity by choosing low-dimensional systems. The maximum value of σ at 600 K is found to be 3.4x104 S/m for bulk sample whereas it is found as 14.2x104 S/m for dA=6nm quantum wire sample. Lastly, the temperature variation of the thermoelectric power factor for all samples is shown in Figure 9d and the highest PF values are 18.5 μWcm1K2 at 550 K for the bulk sample and 130.1 μWcm1K2 at 590 K for dA=6nm quantum wire sample.

Figure 9.

The electronic transport properties of bulk Mg3.2yCoy(Sb0.3Bi0.7)1.99Te0.01 (y = 0.05) and its quantum wire structure with the wire widths of dA = 6 nm, 7 nm, and 8 nm. The temperature variation of (a) Fermi level, (b) Seebeck coefficient, (c) electrical conductivity, and (d) power factor. Dashed lines are theoretical calculations and the star symbol is the experimental measurements read from ref. [72].

In Figure 10a and b, the temperature dependence of electronic thermal conductivity of bulk and quantum wire structures (with different quantum wire widths taken as dA=6,7,and8nm) of Mg3.2−yCoy(Sb0.3Bi0.7)1.99Te0.01 (y = 0.05) samples are represented in the temperature range of 300850 K. From our theoretical calculations at room temperature it is found that while the minimum value of κc is 0.051 Wm1 K1 for bulk sample, a dramatic decrease is gained for quantum wire samples with the minimum value of κc is 2.1 × 10−5 Wm1 K1 for dA=6nm sample. As seen from these temperature-varied electronic properties calculations, choosing low dimensional structured thermoelectric materials rather than bulk leads to enhanced thermoelectric power factor values which will help to get a more efficient thermoelectric material with a higher thermoelectric figure of merit.

Figure 10.

Temperature variation of the electronic thermal conductivity of (a) bulk Mg3.2yCoy(Sb0.3Bi0.7)1.99Te0.01 (y = 0.05) and (b) its quantum wire structure with the wire widths of dA = 6, 7, and 8 nm. Dashed lines are theoretical calculations and the star symbol is the experimental measurements done by Ref. [72].

The bipolar diffusion process is determined as at sufficiently high temperatures electron-hole pairs occur by excitation across the forbidden gap hence a continuous stream of charges flows in the direction of temperature gradient. High-efficiency thermoelectric materials are heavily doped narrow-band gap semiconductors andκbp plays a crucial role above room temperature in these small-band-gap thermoelectrics. In the assumption of an idealized semiconductor, has parabolic valence and conduction bands in that carriers only undergo acoustic mode lattice scattering, theκbp is defined by Glassbrenner and Slack

κbp=b1+b2EgkBT+42kBe2σintTE24

here the ratio of electron to hole mobility is determined with the parameter b. This expression is simplified by Yelgel and Srivastava [65].

κbp=FbpTpexpEg/2kBTE25

where Fbp and p are adjustable parameters vary with doping level and dopant type. When a material’s dimension is changed to low dimensional structures in the expression of κbp the band gap property should be updated.

The most dominant contribution on the total thermal conductivity is sourced from lattice vibrations namely the phonon thermal conductivity κph in degenerately doped and small-band-gap semiconductors. In thermal equilibrium condition, the average number of phonons in the qth mode and at temperature T is defined by the Bose-Einstein distribution function given as n¯q=1expωq/kBT1 where ωq is the angular frequency of phonons, q is phonons wae vector, and n¯q is the average number of phonons in the qth mode. At low temperatures (ωkBT), there is an exponentially small probability for a phonon to be present with n¯exp(ω/kBT). Besides, as the temperature rises and reaches the high temperatures (kBTω), the number of phonons increases linearly with temperature as n¯kBT/ω. Following the linearized phonon Boltzmann equation, the phonon contribution can be expressed by employing Debye’s isotropic continuum model within the single-mode relaxation time approximation as [73].

κph=2qD56π2kBT2scs401x4τqsn¯qsn¯qs+1dxE26

where τ is the phonon relaxation time, cs is the velocity of phonons for polarization branch s,qD is the Debye radius, and x=q/qD is the reduced wave number. Within the Matthiessen’s rule, the total contribution to the phonon lifetime is given by

τqs1=iτqs,i1E27

with τqs,i1 represents contributions from the ith scattering mechanism. Different scattering rates considered for Mg-based bulk and low-dimensional structured thermoelectric materials can be written as

τqs1total=τqs1bs+τqs1md+τqs1dp+τqs1anh+τqs1ims+τqs1idsE28

where scattering mechanisms are sourced from boundary τqs1bs, mass-defectτqs1md, deformation potential τqs1dp, anharmonic τqs1anh, interface-mass-mixing τqs1ims, and interface-dislocation τqs1ids. Theoretical expressions for these phonon scattering mechanisms in bulk, quantum well, and quantum wire structured materials are listed in Table 3.

Scattering mechanismExpression
Boundary scatteringτqs1bs=cs/L; (L: phonon mean free path)
Mass defect scatteringτqs1md=ΓmdΩ4πc¯3ω4qs; Γmd=ifiΔMiM¯2 (c¯: average phonon speed; fi: the percentage of ith isotope present in the crystal; ΔMi=MiM¯; M¯=fiMi)
Deformation potential scatteringτqs1dp=38πEdf2ρcl242mckBT3/2 (Edf: deformation potential; ρ: mass density; cl: longitudinal acoustic phonon velocity; mc: effective mass of a charge carrier)
Anharmonic scatteringτqs1anhBULK=qD5γ24πρc¯2ssεcscsdxx2x+1ϵ+ϵCx+Dxn¯qsn¯++1n¯qs+1+12dxx2x1ϵ+ϵCxDxn¯qsn¯n¯qs(γ: Grüneisen constant; x=q/qD; x±=Cx±Dx; n¯±=n¯x±, C=cs/cs;D=cs/cs)
τqs1anhQWL=τqs1anhBULKnA+nB2/3
(nA: number of well layers; nB: number of barrier layers)
τqs1anhQWR=1/ATω2
AT=kB2π2cκphBULK0ωDℏω/kBT2eℏω/kBTeℏω/kBT12
Interface dislocation scatteringτqs1idsQWL=ΓidsΩ8πc¯3ω04
Γids=2β0nA+nBiexplil0dΔMiM¯21+eA2eB22+1+eB2eA22
(β0: the fraction of broken bonds in the interface region.)
Interface mass-mixing scatteringτqs1imsQWL=ΓimsΩ4πc¯3ω4qs
Γims=2β0nA+nBiexplil0dΔMiM¯21eA2eB22+1eB2eA22
(d: interlayer distance; li: layer distance from interface; l0:z coordinate of the interface; β0: mass-mixing fraction at the interface; eB/eA the ratio of the amplitudes of eigenvectors in materials B and A.)
eBeA=1M0Δ1Mcoslzqz1M02cos2lzqz+Δ1M2sin2lzqz1/2Δ1M
1M0=121MA+1MB; Δ1M=121MA1MB; lz=dA+dB

Table 3.

Expressions for different phonon scattering mechanisms [65, 73, 74, 75, 76].

The temperature dependence of thermoelectric figure of merit for bulk Mg3.2−yCoy(Sb0.3Bi0.7)1.99Te0.01 (y = 0.05) and its quantum wire structure with the wire widths of dA = 6, 7, and 8 nm is represented in Figure 11 with the experimental measurements reported by Ref. [72] for bulk Mg3.2−yCoy(Sb0.3Bi0.7)1.99Te0.01 (y = 0.05) sample. As seen from our theoretical results, our thermoelectric transport modeling can successfully explain the experimental measurement of ZT; while for the bulk sample the maximum value of ZT is theoretically found to be 1.07 at 550 K, its experimental measured value reported as 0.96 at 575 K by Ref. [72]. Additionally, using low dimensional quantum wires structure of Mg3.2−yCoy(Sb0.3Bi0.7)1.99Te0.01 (y = 0.05) with different wire widths, the highest value of ZT at 600 K can be reached to 5.5, 3.7, and 2.7 for dA = 6 nm, dA = 7 nm, and dA = 8 nm, respectively. Therefore low structured thermoelectric materials warrant dramatically enhanced thermoelectric efficiency compared to bulk materials.

Figure 11.

Temperature dependence of thermoelectric figure of merit ZT for bulk Mg3.2yCoy(Sb0.3Bi0.7)1.99Te0.01 (y = 0.05) and its quantum wire structure with the wire widths of dA = 6, 7, and 8 nm. Dashed lines are our theoretical calculations and the symbols are the experimental measurements done by Refs. [72, 77, 78, 79, 80].

In Figure 12, for the comparison the temperature-varied dimensionless figure of merits of n- and p-type Zn4Sb3 and Bi2Te3 nanowire-based thermoelectric materials, the state-of-the-art bulk and layered structured Sn(Se,S) and Mg-based thermoelectric materials and our theoretical Mg-based quantum wire systems are presented together.

Figure 12.

Comparison of the dimensionless figure of merit for Mg-based nanowire structured thermoelectric materials reported in this work with n- and p-type Zn4Sb3 and Bi2Te3 nanowire-based thermoelectric materials [81, 82, 83, 84, 85], and the state of the art bulk and layered structured Sn(Se,S) and Mg-based thermoelectric materials [25, 47, 86, 87, 88, 89, 90] in the literature.

2.1 Summary and future perspectives

Thermoelectric materials are very useful in various types of applications owing to their advantages as already discussed above. One of the current goals in thermoelectric research area is to find new and innovative thermoelectric material systems. During the last decades, a significant advancement in both theoretical and experimental studies has been achieved in the progress of high-efficient thermoelectric materials. Especially, recent advances in nanotechnology have brought with promising methods that cause fundamental mechanisms to get enhanced thermoelectric efficiencies. In particular, low dimensional quantum well and quantum wire systems are niche applications in electronics as micro- and nano-generators due to a lower energy requirements. Although a tremendous progress has been achieved in the past few years, a number of fundamental questions about theoretical background on thermoelectric transport properties in low-dimensional systems still need to be investigated. Therefore, in this book chapter, we aim to explore the temperature variation of the thermoelectric transport properties of bulk and low dimensional materials systematically. We first start our theoretical investigation with the temperature-dependent Fermi level both in extrinsic and intrinsic regimes. Following the linearized Boltzmann transport equations the temperature variant electronic transport properties under the energy-dependent relaxation time approximation are discussed for bulk and low dimensional systems. The thermal conductivity contributions from bipolar and phonon mechanisms are also discussed in detail. Phonon scattering rates sourced from boundaries, mass defects, anharmonicity, deformation potential, interface mass-mixing, and dislocation are given rigorously.

Among thermoelectric materials Mg-based thermoelectric materials have attracted intense interests owing to their enhanced ZT values at low-to-medium temperature range, together with the non-toxic, low-cost, earth abundant, low-density, and environmentally friendly. Here some major outlooks are listed:

  1. Mg-based compounds show highly degenerated conduction bands and can be alloyed and/or doped to form a wide range of compositions, providing enhanced ZT values.

  2. Thermoelectric devices, applied for both power generation and refrigeration, have been successfully fabricated using Mg-based materials and can obtain high conversion efficiency.

  3. Low-dimensional Mg-based systems (quantum wells and quantum wires) suggest even more enhanced ZT values than bulk systems with the help of quantum confinement effect.

  4. As stated above, the main reason having higher ZT values in low-dimensional systems is having extremely low phonon thermal conductivity and high thermoelectric power factor at the same time.

  5. Since there are more intensely experimental studies in Mg-based thermoelectric materials in the literature it is essential to imply further theoretical investigations on Mg-based thermoelectric devices. Therefore, this chapter will be a possible guide to focus on thermoelectric transport properties of Mg-based bulk and low-dimensional systems.

Investigations on thermoelectric transport properties both in bulk and low-dimensional materials are becoming more prominent for solving today’s energy challenges. Collabrations from scientists among different areas such as materials science, physics, and chemistry will have clear advantages in this global competition owing to interdisciplinary nature of thermoelectric research. In the light of these outlooks, developing Mg-based thermoelectric devices are well worthy of systematic study which should speed up their practical applications for energy harvesting. Hereby, the present book chapter comprises an overview of fundamental attainments and breakthroughs in experimental and theoretical thermoelectric research based on Mg-based materials. By presenting a detailed theoretical background on Mg-based 3D, 2D, and 1D systems it is clear that using low-dimensional systems certainly helps us to reach higher thermoelectric efficiencies. This also encourages experimentalists to use low dimensional structured Mg-based systems more often in new-generation thermoelectric devices.

Advertisement

Acknowledgments

Ö. C. Y. acknowledges the support from the Scientific and Technical Research Council of Turkey (TÜBİTAK) (project code: 120F331).

Advertisement

Conflict of interest

The authors declare no conflict of interest.

References

  1. 1. Rowe DM. Thermoelectrics Handbook. Boca Raton: CRC Press; 2005
  2. 2. Poudel B, Hao Q, Ma Y, Lan Y, Minnich A, Yu B, et al. High-thermoelectric performance of nanostructured bismuth antimony telluride bulk alloys. Science. 2008;320:634
  3. 3. Uher C. Materials Aspect of Thermoelectricity. USA Boca Raton: CRS Press Taylor & Francis Group; 2017
  4. 4. Kraemer D, Poudel B, Feng H-P, Caylor JC, Yu B, Yan X, et al. High performance flat-panel solar thermoelectric generators with high thermal concentration. Nature Materials. 2011;10:532
  5. 5. Mahan G, Sales B, Sharp J. Thermoelectric materials: New approaches to an old problem. Physics Today. 1997;50:42
  6. 6. Zhang Y, Chu Y, Xing W, Zheng L, Cao Y. Mechanical properties of thermoelectric Mg2Si using molecular dynamics simulations. Mechanics of Advanced Materials and Structures. 2017;0:1-6
  7. 7. DiSalvo FJ. Thermoelectric cooling and power generation. Science. 1999;285:703
  8. 8. Bell LE. Cooling, heating, generating power, and recovering waste heat with thermoelectric systems. Science. 2008;321:1457
  9. 9. Szczech JR, Higgins JM, Jin S. Enhancement of the thermoelectric properties in nanoscale and nanostructured materials. Journal of Materials Chemistry. 2011;21:4037
  10. 10. Rowe DM. Applications of nuclear-powered thermoelectric generators in space. Applied Energy. 1991;40:241
  11. 11. Vandersande JW, Fleurial JP. Thermal management of power electronics using thermoelectric coolers. In: Fifteenth International Conference of Thermoelectrics Proceedings—ICT ‘96. Vol. 252. Pasadena, CA. 1996. p. 255
  12. 12. Majumdar A. Materials science: Enhanced: Thermoelectricity in semiconductor nanostructures. Science. 2004;303:777
  13. 13. Heremans JP, Jovovic V, Toberer ES, Saramat A, Kurosaki K, Charoenphakdee A, et al. Enhancement of thermoelectric efficiency in PbTe by distortion of the electronic density of states. Science. 2008;321(5888):554
  14. 14. Biswas K, He J, Blum ID, Wu CI, Hogan TP, Seidman DN, et al. High-performance bulk thermoelectrics with all-scale hierarchical architectures. Nature. 2012;489(7416):414
  15. 15. Pei Y, Shi X, LaLonde A, Wang H, Chen L, Snyder GJ. Convergence of electronic bands for high performance bulk thermoelectrics. Nature. 2011;473(7345):66
  16. 16. Heremans JP, Thrush CM, Morelli DT. Thermopower enhancement in lead telluride nanostructures. Physical Review B. 2004;70(11):115334
  17. 17. Liu W, Tan X, Yin K, Liu H, Tang X, Shi J, et al. Convergence of conduction bands as a means of enhancing thermoelectric performance of n-type Mg2Si1−xSnx solid solutions. Physical Review Letters. 2012;108:166601
  18. 18. Zhao LD, Wu HJ, Hao SQ, Wu CI, Zhou XY, Biswas K, et al. All-scale hierarchical thermoelectrics: MgTe in PbTe facilitates valence band convergence and suppresses bipolar thermal transport for high performance. Energy and Environmental Science. 2013;6:3346
  19. 19. Banik A, Shenoy US, Anand S, Waghmare UV, Biswas K. Mg alloying in SnTe facilitates valence band convergence and optimizes thermoelectric properties. Chemistry of Materials. 2015;27:581
  20. 20. Bahk J-H, Bian Z, Shakouri A. Electron energy filtering by a nonplanar potential to enhance the thermoelectric power factor in bulk materials. Physical Review B. 2013;87:75204
  21. 21. Shakouri A, LaBounty C, Abraham P, Piprek J, Bowers JE. Enhanced thermionic emission cooling in high barrier superlattice heterostructures. MRS Proceedings. 1998;545:449
  22. 22. Zhang Q, Liao B, Lan Y, Lukas K, Liu W, Esfarjani K, et al. High thermoelectric performance by resonant dopant indium in nanostructured SnTe. Proceedings of National Academy of Sciences. 2013;110:13261
  23. 23. Biswas K, He J, Zhang Q, Wang G, Uher C, Dravid VP, et al. Strained endotaxial nanostructures with high thermoelectric figure of merit. Nature Chemistry. 2011;3:160
  24. 24. Zhao L-D, Hao S, Lo S-H, Wu C-I, Zhou X, Lee Y, et al. High thermoelectric performance via hierarchical compositionally alloyed nanostructures. Journal of the American Chemical Society. 2013;135:7364
  25. 25. Zhao L-D, Lo S-H, Zhang Y, Sun H, Tan G, Uher C, et al. Ultralow thermal conductivity and high thermoelectric figure of merit in SnSe crystals. Nature. 2014;508:373
  26. 26. Morelli DT, Jovovic V, Heremans JP. Intrinsically minimal thermal conductivity in cubic I−V−VI2 semiconductors. Physical Review Letters. 2008;101:35901
  27. 27. Guin SN, Chatterjee A, Negi DS, Datta R, Biswas K. High thermoelectric performance in tellurium free p-type AgSbSe2. Energy and Environmental Science. 2013;6:2603
  28. 28. Guin SN, Pan J, Bhowmik A, Sanyal D, Waghmare UV, Biswas K. Temperature dependent reversible p–n–p type conduction switching with colossal change in thermopower of semiconducting AgCuS. Journal of American Chemical Society. 2014;136:12712
  29. 29. Hicks LD, Dresselhaus MS. Effect of quantum-well structures on the thermoelectric figure of merit. Physical Review B. 1993;47:12727
  30. 30. Venkatasubramanian R, Siivola E, Colpitts T, O’Quinn B. Thin-film thermoelectric devices with high room-temperature figures of merit. Nature. 2001;413:597
  31. 31. Harman TC, Taylor PJ, Walsh MP, LaForge BE. Quantum dot superlattice thermoelectric materials and devices. Science. 2002;297:2229
  32. 32. Harman T, Walsh M, Turner G. Nanostructured thermoelectric materials. Journal of Electronic Materials. 2005;34:L19
  33. 33. Heremans JP, Wiendlocha B, Chamoire AM. Resonant levels in bulk thermoelectric semiconductors. Energy & Environmental Science. 2012;5:5510
  34. 34. Miraz MH, Ali M, Excell PS, Picking RA. Review on internet of things (IoT). In: Internet of Everything (IoE) and Internet of Nano Things (IoNT). Internet Technologies and Applications (ITA). New York, USA: IEEE; 2015. p. 219
  35. 35. Hiremath S, Yang G, Mankodiya K. Wearable internet of things: Concept, architectural components and promises for person-centered healthcare. In: 4th International Conference on Wireless Mobile Communication and Healthcare-Transforming Healthcare through Innovations in Mobile and Wireless Technologies (MOBIHEALTH). New York, USA: IEEE; 2014. p. 304
  36. 36. Hung NT, Nugraha AR, Yang T, Saito R. Confinement effect in thermoelectric properties of two-dimensional materials. MRS Advances. 2020;5(10):469
  37. 37. Haras M, Skotnicki T. Thermoelectricity for IoT—A review. Nano Energy. 2018;54:461
  38. 38. Suarez F, Nozariasbmarz A, Vashaee D, Öztürk MC. Designing thermoelectric generators for self-powered wearable electronics. Energy & Environmental Science. 2016;9(6):2099
  39. 39. Hyland M, Hunter H, Liu J, Veety E, Vashaee D. Wearable thermoelectric generators for human body heat harvesting. Applied Energy. 2016;182:518
  40. 40. Santos R, Yamini SA, Dou SX. Recent progress in magnesium-based thermoelectric materials. Journal of Materials Chemistry A. 2018;6:3328
  41. 41. Pei YZ, Wang H, Snyder GJ. Band engineering of thermoelectric materials. Advanced Materials. 2012;24:6125
  42. 42. Zhang Q, Li X, Kang Y, Zhang L, Yu D, He J, et al. High pressure synthesis of Te-doped CoSb3 with enhanced thermoelectric performance. Journal of Materials Science: Materials in Electronics. 2014;26:385
  43. 43. Gelbstein Y, Davidow J, Girard SN, Chung DY, Kanatzidis M. Controlling metallurgical phase separation reactions of the Ge0.87Pb0.13Te alloy for high thermoelectric performance. Advanced Energy Materials. 2013;3:815
  44. 44. Morris RG, Redin RD, Danielson GC. Semiconducting properties of Mg2Si single crystals. Physical Review. 1958;109:1909
  45. 45. Heller MW, Danielson GC. Seebeck effect in Mg2Si single crystals. Journal of Physics and Chemistry of Solids. 1962;23:601
  46. 46. Zaitsev VK, Fedorov MI, Gurieva EA, Eremin IS, Konstantinov PP, Samunin AY, et al. Thermoelectrics of n-type with ZT>1 based on Mg2Si-Mg2Sn solid solutions. In: ICT 2005 24th International Conference on Thermoelectrics. 2005. p. 204
  47. 47. Zaitsev VK, Fedorov MI, Gurieva EA, Eremin IS, Konstantinov PP, Samunin AY, et al. Highly effective Mg2Si1-xSnx thermoelectrics. Physical Review B. 2006;74:045207
  48. 48. Zhou Z, Han G, Lu X, Wang G, Zhou X. High-performance magnesium-based thermoelectric materials: Progress and challenges. Journal of Magnesium and Alloys. 2022;10:1719
  49. 49. Tani J, Kido H. Thermoelectric properties of Bi-doped Mg2Si semiconductors. Physica B: Condensed Matter. 2005;364:218
  50. 50. Gao HL, Zhu TJ, Zhao XB, Deng Y. Influence of Sb doping on thermoelectric properties of Mg2Ge materials. Intermetallics. 2015;56:33
  51. 51. Mars K, Ihou-Mouko H, Pont G, Tobola J, Scherrer H. Thermoelectric properties and electronic structure of Bi- and Ag-doped Mg2Si1-xGex compounds. Journal of Electronic Materials. 2009;38:1360
  52. 52. Zhang Q, Yin H, Zhao XB, He J, Ji XH, Zhu TJ, et al. Thermoelectric properties of n-type Mg2Si0.6-ySbySn0.4 compounds. Physica Status Solidi A. 2008;205:1657
  53. 53. Liu W, Kim HS, Chen S, Jie Q, Lv B, Yao M, et al. N-type thermoelectric material Mg2Sn0.75Ge0.25 for high power generation. In: Proceedings of the National Academy of Sciences of U.S.A. 2015;112:3269
  54. 54. Yin K, Su X, Yan Y, You Y, Zhang Q, Uher C, et al. Optimization of the electronic band structure and the lattice thermal conductivity of solid solutions according to simple calculations: A canonical example of the Mg2Si1-x-yGexSny ternary solid solution. Chemistry of Materials. 2016;28:5538
  55. 55. An T-H, Park C, Seo W-S, Choi S-M, Kim I-H, Kim S-U. Enhancement of p-type thermoelectric properties in an Mg2Sn system. Journal of Korean Physical Society. 2012;60:1717
  56. 56. Ihou-Mouko H, Mercier C, Tobola J, Pont G, Scherrer H. Thermoelectric properties and electronic structure of p-type Mg2Si and Mg2Si0.6Ge0.4 compounds doped with Ga. Journal of Alloys and Compounds. 2011;509:6503
  57. 57. Zhang Q, Cheng L, Liu W, Zheng Y, Su X, Chi H, et al. Low effective mass and carrier concentration optimization for high performance p-type Mg2(1-x)Li2xSi0.3Sn0.7 solid solutions. Physical Chemistry Chemical Physics. 2014;16:23576
  58. 58. De Boor J, Saparamadu U, Mao J, Dahal K, Müller E, Ren Z. Thermoelectric performance of Li doped, p-type Mg2(Ge,Sn) and comparison with Mg2(Si,Sn). Acta Materialia. 2016;120:273
  59. 59. Zhang Q, He J, Zhu TJ, Zhang SN, Zhao XB, Tritt TM. High figures of merit and natural nanostructures in Mg2Si0.4Sn0.6 based thermoelectric materials. Applied Physics Letters. 2008;93:102109
  60. 60. Liu W, Tang XF, Li H, Yin K, Sharp J, Zhou XY, et al. Enhanced thermoelectric properties of n-type Mg2.16(Si0.4Sn0.6)1-ySby due to nano-sized Sn-rich precipitates and an optimized electron concentration. Journal of Materials Chemistry. 2012;22:13653
  61. 61. Ning H, Mastrorillo GD, Grasso S, Du B, Mori T, Hu C, et al. Enhanced thermoelectric performance of porous magnesium tin silicide prepared using pressure-less spark plasma sintering. Journal of Materials Chemistry A. 2015;3:17426
  62. 62. Ozen M, Yahyaoglu M, Candolfi C, Veremchuk I, Kaiser F, Burkhardt U, et al. Enhanced thermoelectric performance in Mg3+xSb1.5Bi0.49Te0.01 via engineering microstructure through melt-centrifugation. Journal of Materials Chemistry A. 2021;9:1733
  63. 63. Zhang JW, Song LR, Iversen BB. Probing efficient n-type lanthanide dopants for Mg3Sb2 thermoelectrics. Advanced Science. 2020;7:2002867
  64. 64. Mao J, Liu Z, Ren Z. Size effect in thermoelectric materials. NPJ Quantum Materials. 2016;1:16028
  65. 65. Yelgel ÖC, Srivastava GP. Thermoelectric properties of n-type Bi2(Te0.85Se0.15)3 single crystals doped with CuBr and SbI3. Physical Review B. 2012;85:125207
  66. 66. Yelgel ÖC, Srivastava GP. Thermoelectric properties of Bi2Se3/Bi2Te3/Bi2Se3 and Sb2Te3/Bi2Te3/Sb2Te3 quantum well systems. Philosophical Magazine. 2014;94:2072
  67. 67. Ziman JM. Electrons and Phonons: The Theory of Transport Phenomena in Solids. Great Britain: Oxford University Press; 2001
  68. 68. Richard F. Semiconductor Physics and Applications. Great Britain: Oxford University Press; 2000
  69. 69. Tritt TM. Thermal Conductivity: Theory, Properties, and Applications. New York: Kluwer Academic/Plenum Publishers; 2004
  70. 70. Yu PY, Cardona M. Fundamentals of Semiconductors : Physics and Materials Properties. London: Springer; 2010
  71. 71. Wilson AH. The Theory of Metals. Great Britain: Cambridge University Press; 2011
  72. 72. Chen Y, Wang C, Ma Z, Li L, Li S, Wang J. Improved thermoelectric performance of n-type Mg3Sb2-Mg3Bi2 alloy with Co element doping. Current Applied Physics. 2021;21:25
  73. 73. Srivastava GP. The Physics of Phonons. Bristol: Adam Hilger; 1990
  74. 74. Yelgel ÖC. Chapter6: Recent Advances on Mg2XIVbased Thermoelectric Materials: A Theoretical Approach. U.S.A: John Wiley & Sons, Inc; 2019
  75. 75. Landry ES, McGaughey AJH. Effect of interfacial species mixing on phonon transport in semiconductor superlattices. Physical Review B. 2009;79:75316
  76. 76. Hepplestone SP, Srivastava GP. Theory of interface scattering of phonons in superlattices. Physical Review B. 2010;82:144303
  77. 77. Al H, Chen R, Delgado RD, Liang W, Garnett EC, Najarian M, et al. Enhanced thermoelectric performance of rough silicon nanowires. Nature. 2008;451(7175):163
  78. 78. Al B, Bunimovich Y, Tahir-Kheli J, Yu JK, Goddard WA III, Heath JR. Silicon nanowires as efficient thermoelectric materials. Nature. 2008;451(7175):168
  79. 79. Yang L, Huh D, Ning R, Rapp V, Zeng Y, Liu Y, et al. High thermoelectric figure of merit of porous Si nanowires from 300 to 700 K. Nature Communications. 2021;12:3926
  80. 80. Hernandez JA, Ruiz A, Fonseca LF, Pettes MT, Yacaman MJ, Benitez A. Thermoelectric properties of SnSe nanowires with different diameters. Scientific Reports. 2018;8:11966
  81. 81. Liu Y-B, Zhou S-M, Yuan X-Y, Lou S-Y, Gao T, Shi X-J, et al. Synthesis and high-performance thermoelectric properties of beta-Zn4 Sb3 nanowires. Materials Letters. 2012;84:116
  82. 82. Li L, Xu S, Li G. Enhancement of thermoelectric properties in Bi–Sb–Te alloy nanowires by pulsed electrodeposition. Energy Technology. 2015;3:825
  83. 83. Chen C-L, Chen Y-Y, Lin S-J, Ho JC, Lee P-C, Chen C-D, et al. Fabrication and characterization of electrodeposited bismuth telluride films and nanowires. Journal of Physical Chemistry C. 2010;114:3385
  84. 84. Zhang G, Kirk B, Jauregui LA, Yang H, Xu X, Chen YP, et al. Rational synthesis of ultrathin n-type Bi2Te3 nanowires with enhanced thermoelectric properties. Nano Letters. 2011;12:56
  85. 85. Tan M, Deng Y, Wang Y. Ordered structure and high thermoelectric properties of Bi2(Te, Se)3 nanowire array. Nano Energy. 2014;3:144
  86. 86. Peng KL, Zhang B, Wu H, Cao XL, Li A, Yang DF, et al. Ultra-high average figure of merit in synergistic band engineered SnxNa1-xSe0.9S0.1 single crystals. Materials Today. 2018;21:501
  87. 87. Peng KL, Lu X, Zhan H, Hui S, Tang XD, Wang GW, et al. Broad temperature plateau for high ZTs in heavily doped p-type SnSe single crystals. Energy & Environmental Science. 2016;9:450
  88. 88. Tan XJ, Liu W, Liu HJ, Shi J, Tang XF, Uher C. Multiscale calculations of thermoelectric properties of n-type Mg2Si1-xSnx solid solutions. Physical Review B. 2012;85:205212
  89. 89. Zhang JW, Song LR, Sist M, Tolborg K, Iversen BB. Chemical bonding origin of the unexpected isotropic physical properties in thermoelectric Mg3Sb2 and related materials. Nature Communications. 2018;9:4716
  90. 90. Zhao HZ, Sui JH, Tang ZJ, Lan YC, Jie Q, Kraemer D, et al. High thermoelectric performance of MgAgSb based materials. Nano Energy. 2014;7:97

Written By

Övgü Ceyda Yelgel

Submitted: 03 January 2023 Reviewed: 27 January 2023 Published: 28 February 2023