Paper The following article is Open access

Three-terminal semiconductor junction thermoelectric devices: improving performance*

, and

Published 23 July 2013 © IOP Publishing and Deutsche Physikalische Gesellschaft
, , Focus on Thermoelectric Effects in Nanostructures Citation Jian-Hua Jiang et al 2013 New J. Phys. 15 075021 DOI 10.1088/1367-2630/15/7/075021

1367-2630/15/7/075021

Abstract

A three-terminal thermoelectric device based on a p–i–n semiconductor junction is proposed, where the intrinsic region is mounted onto what is typically a bosonic thermal terminal. Remarkably, the figure of merit of the deviceis governed also by the energy distribution of the bosons participating in the transport processes, in addition to the electronic. An enhanced figure of merit can be obtained when the relevant distribution is narrow and the electron–boson coupling is strong (such as for optical phonons). We study the conditions for which the figure of merit of the three-terminal junction can be greater than those of the usual thermoelectrical devices made of the same material. A possible setup with a high figure of merit, based on Bi2Te3/Si superlattices, is proposed.

Export citation and abstract BibTeX RIS

Content from this work may be used under the terms of the Creative Commons Attribution 3.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

Thermoelectric energy conversion [1, 2] has for decades stimulated considerable research on fundamentals and applications. For a long time, people strove to find good thermoelectric materials with high thermal to electrical energy conversion efficiency. It was found that the optimal efficiency of a thermoelectric device in the linear-response regime is [1, 2]

Equation (1)

with ηC being the Carnot efficiency. The optimal efficiency ηopt is an increasing function of the figure of merit ZT. However, $ZT=T\sigma S^2/\left (\kappa _{\mathrm { e}}+\kappa _{\mathrm { p}}\right )$ is limited by several competing transport coefficients, the conductivity σ, Seebeck coefficient S and electronic (phononic) thermal conductivity κe (κp), making high values of ZT hard to achieve [3, 4]. Mahan and Sofo (henceforth 'M–S') [5] proposed analyzing and achieving high values of ZT by separating it into two factors: (A) TσS2/κe and (B) κe/(κe + κp). By recognizing that the electronic transport quantities, S and κe/σ, are related to the mean and the variance of E − μ (i.e. the heat transferred by an electron at energy E with μ being the equilibrium value of the chemical potential [6]), over the transport distribution function, namely the energy-dependent conductivity σ(E), they were able to obtain [5]

Equation (2)

For a quantity ${\cal O}$ (which is a function of E) $\langle {\cal O}\rangle = \int \,{\mathrm { d}}E \sigma (E)(-\partial f_0/\partial E) {\cal O}(E)/ \int \,{\mathrm { d}}E (-\partial f_0/\partial E) \sigma (E)$ , with f0 being the equilibrium Fermi distribution function. According to M–S [5], high ZT values can be achieved (i) by increasing the factor (A) through decreasing the variance of E − μ via a sharp structure in σ(E), away from the chemical potential μ, and (ii) by reducing the ratio κp/κe. Following this, there were many proposals to achieve effectively narrow electronic bands, especially in nanostructures with transmission resonances and where the enhanced scattering of phonons at interfaces also reduces the phononic heat conductivity κp [710]. However, narrow electronic bands do not necessarily lead to high ZT values. Specifically, when κe ≪ κp, ZT does not increase via reducing the variance of E − μ when it is already limited by κp, or if σ is concurrently decreased, it has been argued [11] to even harm ZT and reduce the power factor σS2. The best figure of merit can only be obtained by considering the competition of all those factors [11].

One should be aware that the energy-dependent conductivity σ(E) is well defined only for elastic processes. In the direct generalization of the M–S results to include inelastic processes, E in equation (2) effectively becomes the average, $\overline {{E}}=(E_{\mathrm { i}}+E_{\mathrm { f}})/2$ , of the initial and final energies (Ei and Ef, respectively) of the transferred electron [12]. Non-trivial aspects of the inelastic processes are revealed in the 'three-terminal thermoelectric devices' proposed very recently [1317].6 By 'three terminal' we mean a setup with an additional thermal terminal supplying bosons (e.g. phonons, electron–hole excitations) involved in the inelastic processes, besides the two electronic terminals. In such devices, in addition to the normal thermoelectric effect in the two electronic terminals, there can be a thermoelectric effect due to the energy transfer between the thermal terminal and the electronic ones. Physically, this is because the energy exchange between the electronic and bosonic systems induces an electric current, or vice versa. The optimal efficiency of such a three-terminal thermoelectrical device in the linear-response regime was found in [16] to be the same as given by equation (1) but with ZT replaced by the three-terminal figure of merit $\tilde {Z}T$ . By optimizing the efficiency of, say, a refrigerator working at equal temperatures of the two electronic reservoirs, it is found that the three-terminal figure of merit is

Equation (3)

The average in equation (3) is taken with respect to the conductance of each inelastic transport channel. Specifically for a quantity as a function of the initial and final energies ${\cal O}(E_{\mathrm { i}},E_{\mathrm { f}})$

Equation (4)

with gin(Ei,Ef) being the conductance of the inelastic channel with given initial and final energies, and ω = Ef − Ei is the energy of the boson (also equal to the energy change of the carrier) in each inelastic process. This generalizes the results of [16] where only a single inelastic transport channel has been considered. In equation (3) Gel (Gin) is the total elastic (inelastic) conductance, Kpp is the purely boson-mediated thermal conductance between the thermal terminal and the other two terminals, and $K_{\mathrm {pe}}={\mathrm { e}}^{-2}G_{\mathrm { in}}\langle \omega ^2 \rangle $ is the thermal conductance characterizing the heat transfer between the bosons and the electrons. This is the generalization of the theory of M–S to the three-terminal case where the principal quantity is now the energy change ω. A direct consequence is that there is no cancelation of the electron and hole contributions to the three-terminal thermopower. We find that a high three-terminal figure of merit $\tilde {Z}T$ requires (i) the dominance of the inelastic transport Gin ≫ Gel, (ii) a small variance of the energy change $\langle \omega ^2\rangle -\langle \omega \rangle ^2\ll \langle \omega \rangle ^2$ and (iii) a large ratio of Kpe/Kpp that can be realized when Gin and $\langle \omega ^2\rangle $ are large or Kpp is small. Small Kpp values should be achievable by e.g. engineering the interfaces between the central system and the two electronic terminals. Remarkably, the purely electronic heat conductance, Ke, does not appear and does not need to be small!

In addition to the pursuit of a narrow distribution of σ(E) in the M–S proposal, the three-terminal figure of merit may also benefit from a 'selection' of the energy change ω either via the electronic structure or via the bosonic spectrum so that the variance of ω can be small. This can be achieved also by a small bandwidth of the bosons involved in the inelastic transport. The merits of the three-terminal configuration are several. (i) There may be no restriction on the effective electronic bandwidth (or other parameters required for a small variance of σ(E)) as the electronic thermal conductivity κe does not appear in the three-terminal figure of merit. (ii) Smaller effective boson bandwidths usually make the bosonic thermal conductance Kpp smaller, which improves $\tilde {Z}T$ . (iii) In general if, e.g. due to momentum or energy conservation (a 'selection'), only the bosons in a small energy range are involved in the transitions, the effective bandwidth can also be small. As a possible candidate, optical phonons have small bandwidths (see [19]) and their coupling with carriers is relatively strong. For the p–n configurations discussed here, this necessitates an electronic band gap smaller than or of the order of the phonon frequencies. This can happen e.g. in solid solutions of the HgCdTe family [19]. Further examples, such as superlattices, are mentioned below. For acoustic phonons the coupling to the carriers is usually stronger at large wavevectors/frequencies [20] (e.g. around the Debye frequency) where the density of states of phonons is also large. In [15, 17], the Coulomb interaction between the quantum-dot system and the lead(s) plays the same role as bosons to induce the inelastic transport.

The choice of the thermal terminal is very important. It should provide bosonic excitations with an energy matching the required electronic one, Eg. When kBT ≪ Eg, and for up going transitions the product of the effective Eg and the exponentially small boson population has to be optimized (as done for example in [5]). In addition, the Debye energy is often not high enough, and multiple phonon processes are quite weak. This can be remedied by using optical phonons or by using a material with ωD ≳ Eg with ωD being the Debye energy (e.g. diamond, whose ωD = 190 meV) for the thermal terminal. An electronic thermal bath where its electron–hole pair excitations or plasmons interact with the electrons in the intrinsic region is another possibility [15]. The direct tunneling thermal conductance between the thermal terminal and the other two terminals can be made exponentially small by controlling the geometry of the contact. Here we suggest and advocate optical phonons. We will mainly consider the optical phonon bath without implying that it is the only possibility.

We thus propose a three-terminal device based on the p–i–n junction where the intrinsic region is contacted with e.g. a phonon source or a thermal terminal, as above. In section 2, we present the device structure and show how the proposal works for semiconductors and some of their superlattices where the band gap can be smaller than the phonon energy. We estimate the figure of merit and find that such a device can have better performance than the usual two-terminal device made of the same material. We conclude this study in section 3. The discussions throughout this paper are focused on the linear-response regime. Nevertheless, if the system is not far away from that regime, such treatment may still offer useful information [2].

2. Three-terminal p–i–n junction thermoelectric device

2.1. Device concept and structure

In the following discussions we specifically consider a p–i–n junction made of 'extremely narrow-gap semiconductors'. The structure is depicted in figure 1(a), for the case of converting thermal to electrical energy. It can be viewed as an analogue of the p–i–n photodiode, where photons are replaced by phonons, i.e. the phonon-assisted inter-band transitions lead to current generation in the junction. The device can also be used to e.g. cool the thermal terminal via the electric current between the electronic terminals. We focus on the situation where single-phonon-assisted inter-band transitions are allowed. (The electronic band gap is hence required to be smaller than the phonon energy.) Such processes may play a significant role in the transport across the junction, whereas the transition rates of multi-phonon processes are much smaller. In semiconductors, the optical phonon energy is usually in the range of 20–100 meV. There are several candidates with such small band gaps: (i) Gapless semiconductors resulting from accidental band degeneracy in solid solutions, such as Pb1−xSnxTe and Pb1−xSnxSe [21]. At a certain mole fraction x the band gap closes, around which it can be very small. (ii) Gapless semiconductors originating from band inversion, such as HgTe and HgSe [21]. The band gap can be tuned via the quantum-well effect in the superlattices composed of a gapless semiconductor and a normal semiconductor without band inversion [22]. An example is the HgTe/CdTe superlattices with a tunable band gap [23]. (iii) Multilayers (superlattices) of 'topological' insulators with ordinary insulator layers sandwiched between the topological ones [24]. For example, in Bi2Te3/Si superlattices the energy gap can be tuned by varying the thickness of the topological and ordinary layers [24]. A significant merit of superlattices is that the lattice thermal conductivity along the growth direction can be much smaller than both of the two bulk constituents. For example, the Si/Ge superlattices have a thermal conductivity about two orders of magnitude smaller than the bulk values [25]. We also point out that the same idea can be applied to devices where the role of optical phonons is played by other bosons. If the energy of such bosons is higher, the requirement for a small band gap can be softened.

Figure 1.

Figure 1. (a) Schematic illustration of a possible three-terminal p–i–n junction thermoelectric device. The phonon source acts as the thermal terminal. The two electronic terminals are the p- and n- doped regions, respectively. As an example, we illustrate the situation where the device converts thermal energy from the thermal terminal to electrical energy. The arrows denote the direction of the electric current. (The device can also cool the thermal terminal by consuming the electrical energy, not directly indicated in the figure). (b) Band structure of the p–i–n junction. The dotted line is the chemical potential at equilibrium. The red arrow labels the phonon-assisted inter-band transition that generates electrons and holes. When drifting with the built-in electric field, the generated non-equilibrium carriers lead to current flow across the junction. (c) Schematic illustration of another possible setup of a three-terminal p–i–n junction thermoelectric device. The thermal terminal (labeled as 'T' in the figure) is a thermal finger, which can be either a phonon reservoir or an electronic one. In the latter case the electron–hole pair excitations or plasmons are coupled with the electrons in the intrinsic region via Coulomb interaction.

Standard image High-resolution image

For simplicity, we consider a linear junction where the conduction band edge varies linearly in the intrinsic region with the coordinate along the junction z from z = −L/2 to L/2 as

Equation (5)

with Eg being the band gap [26]. We set 0 < a < 1 so that the p-doped (n-doped) region is at the left (right) side of the junction (see figure 1(b)). These electronic terminals can have temperatures different from that of the thermal terminal. It is favorable to bend the two electronic terminals away from the thermal one so that they will be better thermally isolated from the thermal terminal and from each other (see figure 1(a)). Another possibility is a 'thermal finger' for the boson bath (figure 1(c)), which can be well isolated from the electronic leads.

2.2. 'Ideal' figure of merit

We start by considering only phonon-assisted transport, ignoring phononic thermal conduction and normal diode transport. The phonon-assisted inter-band transitions generate current flow in the junction. In the linear-response regime, the thermoelectric transport equations are written as [16]

Equation (6)

Here I and IeQ are the electronic charge and heat currents flowing between the two electronic terminals, e < 0 is the electronic charge, IpeQ is the heat current from the thermal terminal to the two electronic ones, Gin is the conductance in the inelastic channels, K0e is related to the electronic heat conductance between the electronic terminals and Kpe is that between the thermal terminal and the electronic ones. L3 is the off-diagonal heat conductance. L1 and L2 are related to the currents induced by the temperature differences (thermopower effect) and the current-induced temperature differences (refrigerator and heater effects). δμ = μL − μR (δT = TL − TR) is the chemical potential (temperature) difference between the two electronic terminals, and $\Delta T=T_{\mathrm { p}}-\frac {1}{2}(T_{\mathrm { L}}+T_{\mathrm { R}})$ is the difference between the temperature of the thermal terminal and the average temperature of the two electric ones, with TL, TR and Tp being the temperatures of the left and right electronic terminals and the phonon terminal, respectively. With these definitions of ΔT, δT and IeQ, IpeQ, the Onsager reciprocal relationships are satisfied. In such a setup, as found in [16], the three-terminal Seebeck coefficient and figure of merit are, when Gel and Kpp are neglected,

Equation (7)

respectively. To obtain the transport coefficients and the figure of merit, we need to calculate the phonon-assisted currents through the system.

The Hamiltonian of the inter-band electron–phonon coupling is [27]7

Equation (8)

where V is the volume of the system, λ is the phonon branch index, ν (ρ) runs through the valence (conduction) band indices, M qλνρ is the matrix element of the electron–phonon coupling and c (a) is the electron (phonon) creation operator. Due to momentum (when valid) and energy conservations, phonons involved in such processes will be in a small energy range. For indirect-band semiconductors and with momentum conservation, these phonons can be acoustic as well as optical ones. For simplicity and definiteness, we consider a direct-band semiconductor system and assume that the contribution from the optical phonons is the dominant one. From equation (8), the net electron–hole generation rate per unit volume, gp, is given for single-phonon transitions by the Fermi golden rule as

Equation (9)

where

Equation (10)

Here, E k+qρ, E kν and ω qλ are the electron and phonon energies when the two systems are uncoupled, and f and N are the non-equilibrium distributions of electrons and phonons in the intrinsic region, with ωji ≡ Ej − Ei > 0 due to energy conservation. According to [28], when the length of the intrinsic region L is sufficiently smaller than the carrier diffusion length, the electronic distribution in the conduction (valence) band in the intrinsic region can be well approximated as the distribution in the n-doped (p-doped) electronic terminal [28]. Similarly the phonon distribution is almost the same as that in the thermal terminal when the contact between the intrinsic region and the thermal terminal is good. Finally, a small amount of disorder that always exists in real systems and relaxes the momentum conservation can enhance the phonon-assisted inter-band transitions.

The transport coefficients are determined by studying the currents at a given bias and/or a temperature difference. The key relation is the continuity equation [28, 29]

Equation (11)

where nα (α = e,h) are the electron and the hole densities in the conduction and valence bands, and neqα are the equilibrium values of those densities. Iα are the charge currents, qα are the charges of the electron and the hole, τα are the carrier lifetimes limited by the recombination processes other than the phonon-assisted ones that have already been taken into account in gp, and gp is the net carrier density generation rate given in equation (9). The currents Iα, which consist of diffusion and drift parts, are

Equation (12)

where χα and Dα are the mobilities and the diffusion constants, respectively. They are related by the Einstein relation, Dα = −(kBT/e)χα. ${\cal E}= aE_{\mathrm {g}}/(eL)$ is the built-in electric field in the intrinsic region.

If Boltzmann statistics for the electrons can be assumed everywhere, the net generation rate gp will depend on z very weakly such that its spatial dependence can be ignored. In this situation, the total carrier densities can be divided into two parts, nα = nα,g + nα,n where nα,g = gpτα are the spatially independent carrier densities generated by the phonon-assisted inter-band transitions and nα,n are the 'normal' densities in the junction, determined by the continuity equation with gp = 0. Similarly, the current is divided into two parts, Iα = Iαn + Iαg. The current in the normal diode channel can be obtained from equations (11) and (12) with proper boundary conditions, yielding the celebrated rectification current–voltage relation

Equation (13)

with Iαns = −eDαL−1αnmα being the saturated currents. Here nmα is the density of the minority carrier and Lα is its diffusion length [28].

Inserting nα,g into equation (12), one obtains the currents in the phonon-assisted channel8

Equation (14)

In the linear-response regime,

Equation (15)

where f0 and N0 are the equilibrium distribution functions of the electrons and the phonons, respectively, and $\overline {{E}}_{ij}\equiv (E_{\mathrm { i}}+E_j)/2$ . Consequently

Equation (16)

where g0p is the equilibrium transition rate (defined in equation (15)) and the average is defined in equation (4) with

Equation (17)

The above results are very similar to those obtained in [16]: due to the inelastic nature of the transport, the carrier energies at the p and n terminals are different; the heat transferred between the two terminals is the average one $\overline {{E}}_{ij} - \mu $ , whereas the energy difference ωji is transferred from the thermal terminal to the two electronic ones. This is also manifested in the way the temperature differences are coupled to the heat flows [16] in equation (15), ensuring the Onsager relations. An important feature is that the three-terminal Seebeck coefficient, L2/(TGin), is negative definite because ωji > 0. In contrast, the two-terminal Seebeck coefficient, L1/(TGin), does not possess this property. It can be positive or negative due to the partial cancelation of the contributions from electrons and holes, whereas there is no such cancelation for L2 (see equation (16)). Equation (16) is a generalization of the results in [16] where there was only a single microscopic energy channel. When many inelastic processes coexist, the contribution of each process is weighed by its conductance. From equations (7) and (16), we find that the 'ideal' three-terminal figure of merit is

Equation (18)

For a single microscopic energy channel system where ωji is fixed, this figure of merit goes to infinity [16]. When there are several such energy channels, it becomes finite due to the non-zero variance of ωji. We estimate the figure of merit when 〈ω〉 − Eg ≃ kBT and γ,kBT ≪ 〈ω〉. Here 〈ω〉 is the average phonon energy and γ is the effective bandwidth (i.e. the variance of ωji due to spectral dispersion) of the involved phonons. From equations (16) and (17) one finds that the variance of ωji is limited by γ2 or (kBT)2, whichever is smaller. For example, when the effective phonon bandwidth γ is much smaller than kBT, the variance is rather limited by γ2. In the situations where $\gamma , k_{\mathrm { B}}T\ll \langle \omega _{ji} \rangle $ , the numerator in equation (18) is much larger than the denominator. The 'ideal' figure of merit can be very high thanks to the electronic band gap when Eg ≫ kBT or the narrow bandwidth of the optical phonons 〈ω〉 ≫ γ. However, in realistic situations, as often happens, the parasitic heat conduction is another major obstacle to a high figure of merit. This will be analyzed in the next subsection.

2.3. Realistic figure of merit

Besides the phonon-assisted transport channel, there is the normal diode channel that is dominated by the (elastic) barrier transmission and the diffusion of minority carriers. It contributes to G, L1 and K0e as well. In addition there are the 'parasitic' heat currents carried by phonons flowing between the two electronic terminals and those from the thermal terminal to the electronic ones. Taking into account all these, the thermoelectric transport equations are written as

Equation (19)

Here, IQ = IeQ + IpQ is the total heat current between the two electronic terminals which consists of the electronic IeQ and the phononic IpQ contributions; ITQ = IpeQ + IppQ is the total heat current flowing out of the thermal terminal to the two electronic ones with IppQ being the purely phononic part. Finally, Gel, L1,el and K0e,el are the contributions to the transport coefficients from the normal diode (elastic) channel, and Kp and Kpp are the heat conductances of phonons flowing between the two electronic terminals and those flowing from the thermal terminal to the two electronic ones, respectively. Note that the elastic channel does not contribute to L2 and L3 which are solely related to the inelastic processes. The three-terminal figure of merit can be obtained by optimizing the efficiency of a refrigerator working at δT = 0, but with finite δμ and ΔT. This figure of merit is [16]

Equation (20)

which is equivalent to equation (3). The diode conductance is $G_{\mathrm { el}} = -(e/k_{\mathrm { B}}T)\sum _{\alpha } I_{ns}^{\alpha }$ . A high figure of merit requires Gin ≫ Gel, which is not difficult to achieve according to the analysis in the next subsection (2.4). In such a situation and when γ ≪ 〈ω〉 or kBT ≪ 〈ω〉, i.e. when the energy width due to kBT or γ gives a much weaker limitation to $\tilde {Z}T$ than Kpp, one finds from equation (20)9

Equation (21)

Estimations carried out in section 2.4 indicate that there are parameter regimes where the figure of merit equation (20) can be greater than the usual two-terminal ones in the same material. We repeat that, unlike the two-terminal case, the thermal conductance Kp between the electronic terminals does not affect the three-terminal figure of merit.

Often Kp and Kpp are of the same order of magnitude. We note that Kp and Kpp are small in several gapless semiconductors such as PbSnTe, PbSnSe, BiSb, HgTe and HgCdTe. In addition, superlattice structures (and other planar composite structures) usually have much lower Kp and Kpp along the growth direction than the bulk materials [30]. The geometry where the electric current flows along that direction is promising for high figures of merit.

2.4. Estimation of the figure of merit

We defined τα (α = e,h) as the carrier lifetimes due to recombination processes other than the phonon-assisted ones that have already been taken into account in gp (see equation (9)). The carrier lifetimes due to the phonon-assisted processes, τe,p for the electrons and τh,p for the holes, satisfy the detailed balance relations g0pτe,p = g0pτh,p = ni in the intrinsic region. Here ni is the electron (hole) density in that region and g0p is the equilibrium transition rate per unit volume defined in equation (15). The transition between minibands in semiconductor superlattices is usually dominated by the phonon-assisted processes in the dark limit when the miniband gap is smaller than the phonon energy [31]. We introduce the parameter ζ to write g0pτe ≃ g0pτh = ζni. This parameter is governed by the electron–phonon interaction strength, being of order unity when such a coupling is strong as is the case for some III–V (and other) semiconductors, or smaller (it will be taken as 1/4 below) [31]. Inserting this into equation (16), one finds

Equation (22)

where L stands for the length of the intrinsic region, which will be taken to be much smaller than the diffusion length Lα [28]. In equation (22) we have chosen a (defined in equation (5)) to be close to 1, which amounts to high doping density $N_{\mathrm { d}}\lesssim N_0\equiv n_{\mathrm { i}}\exp [E_{\mathrm {g}}/(2k_{\mathrm { B}}T)]$ in the p-and n-doped regions. The elastic conductance is the slope of the IV rectification characteristics, equation (13), at V =0,

Equation (23)

with Lα denoting the carrier diffusion length. The majority carrier densities (doping densities) in both the n and p regions are taken to be Nd. It follows then that

Equation (24)

A small ratio can be easily achieved because L < Lα,kBT < Eg and ni ≪ Nd. Therefore, the contribution of Gel is not the main obstacle for a high figure of merit in this device.

For Gel ≪ Gin with γ (the effective width of the relevant energy band) and kBT being much smaller than 〈ω〉 (the typical phonon energy), the figure of merit $\tilde {Z}T$ is given by equation (21). The phononic heat conductance Kpp is, as usual, the least known and an extremely important obstacle for increasing $\tilde {Z}T$ . Kpp can be reduced, in principle, by engineering the interfaces. Even without such improvement, according to the geometry Kpp ≃ 4Kp where Kp is the bulk phonon thermal conductance across the junction. In the relevant regime, Kpe ≃ e−2Ginω2 ≳ e−2GinE2g. From equation (22) we then obtain

Equation (25)

We shall introduce yet another parameter to characterize the ratio Ke/Kp with Ke being the electron thermal conductance in a doped sample of the same geometry and size as the junction, with doping density Nd. The temperature dependence of the ratio Ke/Kp varies with the doping, structure and temperature. The temperature dependence is assumed to be of the form Ke/KpTβ with β being a constant. A temperature-independent pre-factor parameter ξ = KeEβg/(KpkβBTβ) is then introduced. We further assume that the transport properties in the n and p regions are similar (up to signs). The Wiedemann–Franz law implies that Ke = η(kBT)2e−2GD, where GD denotes the electrical conductance in the doped sample and η ≃ 2 for Boltzmann statistics. According to [28], $G_{\mathrm { D}}\simeq 0.25G_{\mathrm { el}}\exp [E_{\mathrm {g}}/(k_{\mathrm { B}}T)]$ when Nd ≃ N0 and L/Lα is small. Using equation (23) one can write $K_{\mathrm {pp}}\simeq 4K_{\mathrm { p}}\simeq 4\xi ^{-1}K_{\mathrm { e}} \simeq E_{\mathrm {g}}^\beta (k_{\mathrm { B}}T)^{1-\beta }\xi ^{-1}\eta N_0\sum _\alpha D_\alpha L _\alpha ^{-1}$ . The figure of merit is then estimated as

Equation (26)

If the temperature dependence comes mainly from the first two factors, then a high figure of merit can be obtained for Eg/(2kBT) = 3 − β. For large Lα/L, the three-terminal figure of merit can be larger than the two-terminal one. The power factor of the device, P = GinS2p, can be estimated similarly.

An especially appealing setup exploiting the topological insulator–ordinary insulator–superlattices can be built as follows (taking the superlattice Bi2Te3/Si as an example). On the front and back surfaces of each thin Bi2Te3 layer, there are protected surface states with a gapless Dirac-cone like spectrum [32]. Tunneling between the two surfaces opens a gap in the spectrum of each surface band [24]. In superlattices these states form a pair of conduction and valence minibands where the band gap can be controlled via the thickness of the two types of layers (for details see [24] and footnote10). As the energy of the optical phonon in Si is much larger than that in Bi2Te3, the optical phonons in Si layers are well localized in these layers. So are the optical phonons in the Bi2Te3 layers. Similarly, due to the significant mismatch of the mechanical properties of the two types of layers, the acoustic phonons also have difficulty in being transmitted across the interfaces. The phononic thermal conductivity along the growth direction is then considerably reduced. It should be smaller than the values for both materials in the bulk [25]. On the other hand, the phononic heat conductivity within each thin layer of Si is large. At T = 300 K, the former is 1.5 Wm−1 K−1 [33], whereas the latter is about 102 Wm−1 K−1 [34]. When the electric current is along the growth direction, the small phononic heat conductance along this direction greatly reduces the parasitic heat conduction Kpp and benefits the figure of merit. On the other hand, within each Si layer phonons are transferred efficiently from the thermal terminal to the system, which is good for enhancing the output power of the device.

We now estimate the figure of merit of the device using the example of the semiconductor made of the Bi2Te3/Si superlattice. We shall use the transport parameters of bulk Bi2Te3 to do the estimation, although the superlattice should have better thermoelectric performance [35]. First tune the superlattice structure to make Eg ≲ 〈ω〉 with 〈ω〉 being the energy of the optical phonon in Si, which is about 60 meV (see [19]) (equivalent to about 700 K). We will choose Eg = 600 K. In Bi2Te3 with Nd ≃ 1020 cm−3, one finds from  [33] that κe ≃ κp at 300 K. This determines the parameter ξ to be 2β. Using these parameters we calculate and plot the figure of merit as a function of temperature in figure 2 for a practically achievable value Lα/L = 4 and a modest value of ζ = 1/4. In the calculation we ignored the temperature dependence of ζ and Lα. The results are computed for three situations with β = 1,2 and 2.5. It is seen that the figure of merit with the several underestimations made can be larger than 1 around room temperature for all the three situations, indicating potential usefulness. Finally, we note that the same strategy and analysis can also be applied to the superlattice made of an inverted-band gapless semiconductor and a normal semiconductor, where the band gap can be tuned via the quantum-well effect [22] and the normal semiconductor can be chosen to have the proper optical phonon frequency and the high thermal conductivity to enable better thermoelectric performance.

Figure 2.

Figure 2. Figure of merit $\tilde {Z}T$ of Bi2Te3/Si superlattices as a function of temperature for β = 1 (solid curve), 2 (dashed curve) and 2.5 (dotted curve). Eg = 600 K, ζ = 1/4, η = 2, Lα/L = 4 and ξ = 2β as Kp = Ke at T = 300 K.

Standard image High-resolution image

3. Conclusions and discussions

We proposed and studied a thermoelectric device using the phonon bath as an example. The scheme is based on the 'three-terminal' geometry of thermoelectric applications [1317], where inelastic processes play a crucial role. It has been shown that a high thermoelectric figure of merit can be achieved in this geometry in several nanosystems [16], where only one microscopic energy channel, in which the relevant electronic energy is fixed, is available. In this paper, we derived the figure of merit for the multiple energy channel case. We find that when only the inelastic processes are considered, the figure of merit is the ratio of the square of the mean value of energy difference between final and initial states to its variance with the average weighed by the conductance of each microscopic process. A small variance in the energy change is then favorable for a high figure of merit. To achieve such good energy selection, one can use either the electronic band gap, Eg ≫ kBT, for electrons or the narrowness of the phonon band, $\gamma \ll \left \langle \omega \right \rangle $ , for phonons. The realistic figure of merit including other processes, equation (3), is also discussed. It is found that a strong carrier–boson coupling as well as the dominance of inelastic transport and a small purely phononic thermal conductivity between the phonon terminal and the electronic ones are necessary for a high figure of merit. The suppression of elastic transport can be achieved with a semiconductor junction [28], while the coupling is strong when the bosons are e.g. phonons, electron–hole pair excitations, etc. Thanks to these, the proposed three-terminal device can have a figure of merit higher than that of the usual two-terminal device made of the same material.

In comparison with the existing literature, M–S suggested a narrow electronic band for elastic two-terminal transport to achieve high values of ZT. When this scheme is generalized to inelastic processes where the initial Ei and final Ef energies are different, high values of ZT are possible when the distribution of the average energy $\overline {{E}}=(E_{\mathrm { i}}+E_{\mathrm { f}})/2$ (measured from the common chemical potential) is narrow in the two-terminal geometry. In the three-terminal situation, a narrow distribution of ω = Ef − Ei plays a crucial role. The latter can also be achieved by controlling the initial and final electronic states by a barrier enough higher than T, or in a small system with a few initial and final states with fixed ω = Ef − Ei, as in [1517], or for example, via the narrow bandwidth of optical phonons. Finally, we note the analogy of the suggested configuration, which may convert thermal to electrical energy, and a photovoltaic device.

Acknowledgments

OEW acknowledges support from the Albert Einstein Minerva Center for Theoretical Physics, Weizmann Institute of Science. This work was supported by the BMBF within the DIP program, BSF, by ISF and by its Converging Technologies Program.

Footnotes

  • Patent application pending.

  • There are some similar ideas and experimental studies in the literature, e.g. [18].

  • There are other possible inter-band transition processes, such as the defect trap state-mediated inter-band transitions (the Shockley–Read–Hall processes). These are higher-order processes, for which the rate should be smaller. There are also the Auger processes due to inter-band Coulomb interaction. The Auger processes are important for narrow-band semiconductors such as InSb. As far as we know, there is no theoretical work devoted to the comparison of the Auger processes and optical-phonon-induced inter-band transitions when the band gap is smaller than the optical-phonon energy. However, for the superlattice structures studied in this work, the transition should be very much like the inter-subband transitions in semiconductor quantum wells. In the latter, experimental and theoretical investigations (see [27]) revealed that the optical-phonon interaction (especially confined and interface phonons) dominates the inter-subband scattering. This supports our assumption that the inter-miniband transitions in the superlattice-based device are dominated by the optical-phonon processes. Moreover, as the Auger processes conserve the total energy of the electronic system, they will not affect the transfer of energy between the electronic system and the thermal bath. Hence, although we do not include the Auger processes in the theoretical treatment, they will not spoil the functionality and performance of the device.

  • Here we have used certain simplifications. The consideration only holds for the central part of the intrinsic region, whereas the situation near the boundaries is more complicated. However, this is not supposed to change the results considerably [29].

  • Generally one can estimate the figure of merit as $\tilde {Z}T\simeq {\mathrm { min}}\left \{K_{\mathrm { pe}}/K_{\mathrm { pp}}, (E_{\mathrm { g}}/k_{\mathrm { B}}T)^2, (\langle \omega \rangle /\gamma )^2\right \}$ when Gel ≪ Gin. We mostly assume that Kpe/Kpp is the smallest among the three.

  • 10 

    In [24] it is shown that the gap is Eg = 2|ΔS − ΔD| where ΔS and ΔD are the tunneling rates through the topological and the ordinary insulator layer, respectively. They (and the band gap) can be tuned in a large range via the thickness of these layers.

Please wait… references are loading.
10.1088/1367-2630/15/7/075021