2018 Volume 59 Issue 7 Pages 1013-1021
Although there is a growing demand for first-principles predictions of the thermoelectric properties of materials, the contribution of various errors in Boltzmann transport calculations is not negligible. We conducted a typical first-principles calculation and a Boltzmann transport analysis on a typical semiconductor (Si) at various temperatures T while varying the band gap εg, electron relaxation time τel, and phonon thermal conductivity κph to demonstrate how the calculated thermoelectric properties, which are functions of the carrier doping level, are affected by these parameters. Bipolar conduction drastically decreased zT via a degradation of the Seebeck coefficient S and an increase in the effective Lorenz factor Leff, indicating the importance of a wide enough εg (several multiples of kBT or higher) for high zT. Thus, the underestimation of εg, which frequently happens in first-principles calculations, could induce large errors in calculations for narrow-gap semiconductors. The calculation of the electron thermal conductivity without Peltier thermal conductivity was found to limit the zT of typical semiconductors to below 1. A small value of κph/τel, where κph/τel is the degree to which a material is a phonon-glass electron-crystal, was necessary to achieve a high zT. Fitting the calculations with experimental thermoelectric properties showed that τel can vary by an order of magnitude from 10−15 to 10−14 s, depending on both T and the samples. This indicates that the use of a fixed relaxation time is inappropriate for thermoelectric materials.
Discoveries of new environmentally-friendly thermoelectric materials for solid state power generation and Peltier cooling devices are highly anticipated.1) Although thermoelectric materials have been intensively sought after for over half a century, many potentially thermoelectric compounds remain unstudied or understudied. Experiments tend to underestimate the potential of compounds for use as thermoelectric materials, since the optimum doping level and the optimum method for introducing phonon scattering are rarely achieved during the first set of the experiments.
The best-known and the most popular benchmark for the efficiency of a thermoelectric material is the dimensionless thermoelectric figure of merit
\begin{equation} zT = \frac{S^{2}\sigma T}{\kappa} = \frac{S^{2}\sigma T}{\kappa_{\text{el}} + \kappa_{\text{ph}}}, \end{equation} | (1) |
Equation (1) indicates that a high S, high σ, and low κ are needed to achieve a high zT. Of these parameters, S, σ, and κel cannot be controlled independently because of their strong dependence on the carrier doping level n ≡ ne − nh, where ne and nh are the electron and hole concentrations. In doped semiconductors, a decrease in the carrier density increases S and decreases κel but also decreases σ. As a result, zT is low both in low-n and high-n limits, and there is an “optimum n” (usually between 1019 and 1021 cm−3) where zT is maximized.1) Comparisons between different materials usually rely on the maximum reported zT values. However, the maximum zT of a material can only be achieved after developing effective methods to (1) dope carriers up to the optimum n and (2) to decrease κph while avoiding a decrease in the electron relaxation time τel. Therefore, researchers have to find the correct impurity elements or good defects to control n and scatter phonons without also scattering electrons. Some materials already have intrinsic phonon scattering mechanisms in their pure state and thus exhibit fairly high zT values without additional processes being required to scatter phonons. The search for such intrinsically low-κph materials is currently being conducted with the help of sophisticated phonon calculations and machine learning.2)
Data-driven searches for intrinsically high-zT materials within vast lists of materials are rapidly progressing. Recently, high-throughput first-principles calculations3–7) have been combined with Boltzmann transport calculation codes such as those in the Boltzmann Transport Properties (BoltzTraP8)) program package. In the TE Design Lab database4,5) the electronic properties that are related to the thermoelectric properties of over 2,300 compounds have been calculated theoretically. The database was carefully designed to prevent misleading people into believing that a single representative value for zT can be predicted for each compound. Instead of directly predicting the maximum zT, the database provides the values of the β-factor,5) which is closely related to zT via zT = vβ/(uβ + 1). Properties that are intrinsic to the electronic structure are included in β, and the extrinsic variables related to chemical potential and scattering mechanisms are excluded (they are represented by v and u, respectively). Compounds with a larger β value have higher chances of exhibiting a high zT after optimization of n and suppression of the electron scattering.
In such calculations of thermoelectric materials it is relatively easy to reproduce the experimental zT values, because we can set arbitrary values for n, τel, and κph if necessary. Observing the trends is also possible, because we can fix these unknown variables as constants, such as the common approximation τel = 10−14 s. However, predicting zT is not easy, because all these unknown variables need to be determined accurately.
Since S, σ, and κel are strongly dependent on the carrier concentration, we need to determine the chemical potential μ that represents the value of n in the sample. We also need to make sure that the calculated value for εg is correct, because determining εg from first-principles calculations often yields incorrect values, even though many thermoelectric materials are narrow-gap semiconductors.
The solutions of the Boltzmann transport equations8,9) yield σ, S, and κ as a function of both μ and T by integrating over energy ε such that
\begin{equation} \sigma (\varepsilon,T) = \frac{e^{2}}{3}D(\varepsilon)v^{2}(\varepsilon)\tau_{\textit{el}}(\varepsilon,T), \end{equation} | (2) |
\begin{equation} \sigma (\mu,T) = \int \sigma (\varepsilon,T)\left(- \frac{\partial f(\varepsilon,T)}{\partial \varepsilon}\right)d\varepsilon, \end{equation} | (3) |
\begin{equation} S(\mu,T) = - \frac{1}{eT\sigma (\mu,T)} \int (\varepsilon - \mu)\sigma (\varepsilon,T)\left(- \frac{\partial f(\varepsilon,T)}{\partial \varepsilon}\right)d\varepsilon, \end{equation} | (4) |
\begin{align} \kappa_{\text{el}}(\mu,T) &= \frac{1}{e^{2}T} \int (\varepsilon - \mu)^{2}\sigma (\varepsilon,T)\left(- \frac{\partial f(\varepsilon,T)}{\partial \varepsilon}\right)d\varepsilon \\ &\quad- S(\mu,T)^{2}\sigma (\mu,T)T. \end{align} | (5) |
Note that there are two terms for κel in eq. (5). The first term κ0 represents the normal diffusion of heat. The second term corresponds to the self-Peltier effect: the electric current induced by the thermoelectric voltage carries the heat backwards from the low-T to the high-T end. In this paper, we refer to this second term as the Peltier thermal conductivity κP ≡ −S2σT. When calculating κel, κP is often neglected because it is negligibly small in ordinary metals.9) However, this term may not be negligible in thermoelectric materials, because they are designed to maximize P.
By using these calculable parameters σ/τel, S, and κel/τel, the full expression for zT becomes
\begin{equation} zT = \frac{S^{2}\sigma T}{\kappa_{\text{el}} + \kappa_{\text{ph}}} = \frac{S^{2}(\sigma/\tau_{\text{el}})T}{(\kappa_{\text{el}}/\tau_{\text{el}}) + (\kappa_{\text{ph}}/\tau_{\text{el}})}. \end{equation} | (6) |
Here, the electron scattering time τel is the interval between electron scattering events. Because of a lack of knowledge regarding the real value of τel, a constant relaxation time approximation with a fixed τel is used for all T to roughly estimate σ and κel. A widely used approximation is τel = 1 × 10−14 s.3–8) Compared with their respective pure compounds, high-zT thermoelectric materials are expected to have a higher concentration of electron scattering centers as they originate from the materials that were introduced to control n and scatter phonons.
Scattering theories predict that τel has a T-dependence that depends on the dominant scattering mechanisms. When acoustic phonon scattering is dominant, τel is proportional to T−1/2 in covalent compounds, T−1 in metals, and T−3/2 in ionic compounds. When ionic impurity scattering is dominant, τel is proportional to T3/2. There are also other scattering centers such as grain boundaries and crystal defects. The overall τel can be calculated from the inverse summation of τel of every scattering mechanism (Στel−1)−1.
There are a few theoretical approaches to calculate the τel values of Si12,13) and PbTe14) from first-principles. Their results suggest that τel lies between 10−15 and 10−13 s. Additionally, there have been attempts to estimate the τel of Mg2Si1−xSnx,15,16) skutterudites,17) SnSe,18) Ga2Ru,19) and AgGaTe220) by fitting the Boltzmann transport equations with experimental σ and carrier concentration values. Further, the electron mobility RHσ = eτel/m*, where RH is the experimental Hall coefficient and m* is the calculated effective mass, can be used to estimate τel in experimental samples, if the correct value for m* is used. Chen et al. report a list of the ratio of both the experimental and calculated mobilities of 31 compounds21) under the assumption that τel = 1 × 10−14 s; the ratio ranged from 0.0005 to 10, corresponding to τel values between 7 × 10−18 s and 9 × 10−14 s. Although this mobility calculation has other possible sources of errors such as effective mass inaccuracy and anisotropy, this implies that τel can vary by orders of magnitude between different samples.
When we ignore this unknown term κph/τel, the expression becomes zelT: the theoretical upper limit for zT in the limit of κph → 0 is given by
\begin{equation} z_{\text{el}}T(\mu,T) \equiv \frac{S^{2}\sigma}{\kappa_{\text{el}}} = \frac{S^{2}(\sigma/\tau_{\text{el}})T}{\kappa_{\text{el}}/\tau_{\text{el}}} = \frac{S(\mu,T)^{2}}{L_{\text{eff}}(\mu,T)}. \end{equation} | (7) |
\begin{equation} L_{\text{eff}}(\mu,T) \equiv \frac{\kappa_{\text{el}}(\mu,T)}{\sigma (\mu,T)T} = \frac{\kappa_{\text{el}}(\mu,T)/\tau_{\text{el}}}{(\sigma (\mu,T)/\tau_{\text{el}})T} \end{equation} | (8) |
In this study, we attempted to demonstrate how these unknown parameters and calculation errors affect the Boltzmann transport calculation of a material’s thermoelectric properties as this calculation is widely used in high-throughput first-principles calculations. We selected Si (Si–Ge alloy) as the test material for the demonstration. It should be noted here that more precise analysis models exist for the theoretical and experimental transport properties of Si–Ge alloys.11,22–24)
The electronic structure of Si was calculated using WIEN2k code,25) which employs the full-potential linearized augmented plane wave (FLAPW) method. The initial lattice parameters and atomic coordinates were obtained at room temperature. For the exchange-correlation potential, the generalized gradient approximation (GGA) by Perdew, Burke, and Ernzerhof26) was applied. In comparison, the Tran-Blaha modified Becke-Johnson (TB-mBJ)27) potential was applied to the calculation via the GGA to achieve more precise values for the band gaps.28) Spin-polarization and spin-orbit interactions were ignored. The number of k-points was set to 8000 in the first Brillouin zone in the initial calculation, and the mesh was later interpolated by a 5× denser mesh in reciprocal space using code in the BoltzTraP8) program package. The muffin-tin radius RMT was set to 2.10 a.u., while the maximum k was set such that RMTkmax = 7. The core-valence cutoff energy was set to Ecut = −6.0 Ry. The convergence criteria for energy and charge were 0.0001 Ry and 0.0001 e, respectively. The Boltzmann transport analysis was conducted using BoltzTraP8) version 1.5.3; the code was slightly modified29) to include κP during the calculation of the transport tensors.
Figure 1 shows the density of states DOS and energy dispersion curves for Si calculated using the GGA26) and TB-mBJ27) potentials. When GGA was used the εg was underestimated (around 0.571 eV), which is nearly half the experimental εg for Si (1.1701 eV).27) This underestimation is a well-known, common problem of first-principles calculations based on density functional theory (DFT).28) The use of TB-mBJ resulted in a better value of εg of ∼1.156 eV. Although the raw calculation result provides the Fermi energy εF at the top of the valence band, εF at n = 0 (undoped) is shifted to the middle of the band gap during transport calculations by BoltzTraP.29) In reality, εF moves up with electron doping and moves down with hole doping.
(a) Density of states DOS and (b) band structure of Si calculated using the FLAPW method using GGA (black) and TB-mBJ (red) exchange-correlation potentials.
Figure 1(b) shows the band structure of Si along representative reciprocal lattice vectors. The valence band top is composed of one heavy band and one light band, which are degenerate at Γ point, so that the number of hole pockets is two in p-type Si. On the other hand, the conduction band bottom is located along the Γ-X line, so there are six electron pockets in n-type Si.
It can be seen that the band structures obtained using TB-mBJ and GGA are essentially similar, except for the value of εg. The concavity of the bands or the effective masses m* were slightly greater with TB-mBJ than with GGA. We used the calculation results based on the experimental lattice parameter30) a = 5.4310 Å at 300 K since there were negligible differences in the calculated value of S. Changes in the lattice volume from −5% to 10% resulted in differences in S within ±0.5 µV/K at n ≈ 1020 cm−3 in the GGA calculations. An optimization of the lattice parameters from GGA calculations resulted in an increased lattice parameter a = 5.4797 Å (+2.7% in volume); using this value of a, εg increased up to 1.212 eV in the TB-mBJ calculation. Since the agreement with the experimental εg was poorer with the optimized a, we employed the results with the experimental a in the following calculations.
3.2 Carrier doping levelIn this study, we plotted the thermoelectric properties against the carrier doping level n in units of inverse cubic centimeters by dividing the amount of additional charge by the unit cell volume.32) This corresponds to the carrier doping level or net carrier concentration, i.e., the difference between the electron and hole concentrations. This n is expected to be constant within a sample, because n is related to the initial dopant concentration in the sample and is not affected by the thermal excitation of electron–hole pairs across the band gap. We plotted the materials’ thermoelectric properties against n on a logarithmic scale. This representation is more informative than using a linear scale, which shrinks the information around the optimum n into a “spiky” region around n = 0.
3.3 Seebeck coefficientFigure 2(a) shows the values of S of Si at different T plotted against log n. Both p-type and n-type Si showed very similar curves, even though the band structures of the valence band and the conduction band are different. At equal n, the n-type Si showed slightly greater |S| than p-type Si, especially in bipolar states. This is possibly due to the difference in number of carrier pockets: more carriers exist near the band edge in n-type Si, so that |S| of electrons becomes greater than |S| of holes.
The dependence of the absolute values of calculated Seebeck coefficient |S| on the carrier doping level n in p-type and n-type Si (a) at various T between 300 K and 1300 K with intervals of 100 K, (b) at 900 K using scissors-cut band energies with εg values between 0 and 1.17 eV in GGA calculation, with TB-mBJ calculation (εg ∼ 1.16 eV) as a reference. (c) The comparison between calculated and experimental values of |S| for Si0.8Ge0.2 alloys (reported by Vining et al.22)) plotted against Hall carrier densities for p-type and n-type Si at 300 K. The S–n curves of Si were calculated by using TB-mBJ, and GGA with unshifted εg ≈ 0.571 eV. The dark solid circles correspond to the samples, which were used in the calculation of τel in the following sections.
At 300 K, the absolute value of |S| decreased monotonically with increasing n from 1016 to 1022 cm−3. Extremely high values of |S| were predicted at low n. When we compare |S| at fixed n, the |S|–n curves shifted upwards with increasing T. Above 700 K, peaks started to appear in the |S|–n curves. The peak position moved to higher n with increasing T.
These behaviors in S are in agreement with bipolar conduction occurring; this conduction is caused by the thermal excitation of electron–hole pairs across the band gap. The drastic decrease in |S| is induced by the increase in the total carrier concentration and by the cancellation of S by carriers of the opposite sign. This effect is significant at low n, because thermal excitation adds equal concentrations of electrons and holes, which dominate over the carrier concentration induced by dopants.
The degradation of |S| by bipolar conduction was more significant when εg was smaller. Figure 2(b) shows the calculated S at 900 K for artificially set εg values (from the experimental value 1.17 eV down to 0 eV) using the ‘scissors-cut’ operation implemented in BoltzTraP. Once εg was set to 1.17 eV in GGA, it roughly matched the value of S calculated using TB-mBJ. However, when εg was set to smaller values, |S| was reduced on the left side of the plot, indicating that the decrease in |S| at low n is influenced by bipolar conduction. When εg was decreased down to 0.75 eV (10 kBT), it started to influence S at n ≈ 1019 cm−3, which is the optimum range of n for thermoelectric materials. When εg was decreased down to 0.25 eV (3.3 kBT), the influence of bipolar conduction on S was as high as n ≈ 1020 cm−3. When εg was set to 0, the peak of |S| was around 1021 cm−3; however, the peak value of |S| was very small.
These results tell us two things about εg. First, we should select a parent compound for thermoelectric materials that has a wide enough εg. This concept is known as the 10 kBT-rule.33) At least for Si, bipolar conduction lowers the |S| of the optimum range of n when the εg is less than 10 kBT. Second, when the calculated value of εg is small, the predictions of the thermoelectric properties will contain a large uncertainty. This highlights the difficulty of using first-principles calculations to determine a material’s thermoelectric properties. Although many thermoelectric materials are narrow-gap semiconductors, their band gaps are often underestimated in DFT calculations.
Figure 2(c) shows the agreement of the experimental S and calculated S(n) at 300 K. The figure shows the calculated S for T = 300 K in comparison with the experimental S of p-type and n-type Si0.8Ge0.2 alloys.22) Although there was a high agreement between calculated and experimental |S|, there were slight differences in |S|, especially for n-type Si0.8Ge0.2. One of the possible origins for this is that electronic structure of Si0.8Ge0.2 is different from the one for pure Si, especially in the conduction band.
3.4 σ/τel, κel/τel, and the effective Lorenz factorFigure 3(a) and (b) show the n-dependence of the calculated value of σ/τel of Si. Almost no difference in σ/τel was observed between p-type and n-type Si. At 300–600 K, we can see that σ/τel is almost proportional to n. The quantity σ/τel was independent of T, as expected from the Drude model σ/τel = ntotale2/m*, where ntotal = ne + nh $ \simeq $ n under monopolar conduction. Above 700 K, the σ/τel–n curves showed a bend at low n. This is because of the increase in ntotal caused by bipolar conduction. It should be noted that the thermal excitation of carriers does not change n.
Carrier density dependences of calculated values of (a) σ/τel in p-type Si, (b) σ/τel in n-type Si, (c) κel/τel in p-type Si, (d) κel/τel in n-type Si, (e) Leff in p-type Si, and (f) Leff in n-type Si between 300 and 1300 K in intervals of 100 K.
Figure 3(c) and (d) show the n-dependence of κel/τel calculated using two different expressions for κel. At the temperature where the bipolar conduction is negligible (300–600 K), κel/τel increased monotonically with n, just as σ/τel does (see Fig. 3(a) and (b)). This similarity can be explained via the Wiedemann-Franz law (κel = L0σT).
Large discrepancies in the calculated results were observed when we excluded or included the Peltier term κP = −S2σT. When κel/τel was calculated according to κel/τel = κ0 + κP, the value of κel/τel was lower than when κel/τel was calculated according to κel/τel = κ0. This difference was the largest around n ≈ 1020 cm−3, where P = −κP/T tended to reach its maximum. This shows that the approximation κel = κ0 is unsuitable, especially for thermoelectric materials that are designed to achieve a maximum P. In the metallic limit (n > 1021 cm−3) and in the bipolar limit (n < 1018 cm−3 for T = 1300 K), the value of κel/τel calculated with κP was almost equal that calculated without κP. The bipolar effect increases κel not only by increasing ntotal, but also by increasing the energy per carrier particle by εg. Thermally excited carriers possess additional energy equal to εg from the high-T end to the low-T end of the sample. Therefore, zT is suppressed under bipolar conduction both because of the cancellation of S and because of the increase in κel.
By using both the calculated κel/τel and σ/τel values, we calculated the effective Lorenz factor Leff = κel/(σT) = (κel/τel)(σ/τel)−1T−1 and compared the results to those of the theoretical Lorenz factor L0 in the free-electron model
\begin{equation} L_{0} = \frac{\pi^{2}k_{\text{B}}{}^{2}}{3e^{2}} \approx 2.45 \times 10^{-8}\,\text{V$^{2}$\,K$^{-2}$}. \end{equation} | (9) |
The n-dependence of zelT is thought to be material-dependent and a function of both n and T; it is believed that this dependence can be determined purely from its electronic structure. However, as presented in Fig. 4(a) and (b), the n-dependence of zelT calculated using different expressions for κel exhibited very different behaviors. When we used κel = κ0 + κP, the zelT of Si could significantly exceed 1 at low n. The simple expression zelT = S2/L0 also yielded similar zelT curves to the ones obtained with κP, although some discrepancies were observed in the low-n region where bipolar conduction was dominant.
Dependence of zelT on the carrier doping level of (a) p-type and (b) n-type Si calculated using three different expressions for κel between temperatures of 300 and 1300 K in intervals of 100 K.
However, when we used the approximation κel = κ0, a strange upper limit of zelT appeared at zelT = 1, indicating that there was no chance of achieving zT > 1. In our calculations this limit was observed for almost any semiconductors with steep band edges. This is consistent with past theoretical calculations that included κP9,34–37) or used the Wiedemann-Franz law18) to successfully achieve zelT > 1, while studies that only used κ07,38–41) only found values of zelT below 1.
The origin of this false “wall of zelT = 1” can be explained as follows. In the low-n limit of non-degenerate semiconductors, μ (the center of the Fermi-Dirac distribution) is located nearly at the middle of the band gap, where no hole/electron states exist. In normal semiconductors the DOS at the band edges are very steep. As a result, the hole/electron states are concentrated near the band edge energy ε0, which is either the top of the valence band (εV) for holes or the bottom of the conduction band (εC) for electrons. At low T, only a few carriers can be thermally excited up to these states, and the difference in ε of each carrier from ε0 is negligibly small, compared to the large ε0-μ. Then we can treat ε-μ as a constant (ε0-μ) for all carriers, which can be taken out of the integrals of eqs. (3)–(5).
\begin{align} S(\mu,T) \xrightarrow{| \varepsilon_{0} - \mu | \gg \varDelta \varepsilon} &- \frac{1}{eT\sigma(\mu,T)}(\varepsilon_{0} - \mu)\sigma(\varepsilon_{0},T) \\ &\quad \rightarrow - \frac{1}{eT}(\varepsilon_{0} - \mu) \end{align} | (10) |
\begin{equation} S(\mu,T)^{2}\sigma (\mu,T)T \xrightarrow{| \varepsilon_{0} - \mu | \gg \varDelta \varepsilon}\frac{1}{e^{2}T}(\varepsilon_{0} - \mu)^{2}\sigma(\varepsilon_{0},T) \end{equation} | (11) |
\begin{equation} \kappa_{0}(\mu,T) \xrightarrow{| \varepsilon_{0} - \mu | \gg \varDelta \varepsilon}\frac{1}{e^{2}T}(\varepsilon_{0} - \mu)^{2}\sigma(\varepsilon_{0},T) \end{equation} | (12) |
\begin{equation} \kappa_{\text{P}}(\mu,T) \xrightarrow{| \varepsilon_{0} - \mu | \gg \varDelta \varepsilon} - \frac{1}{e^{2}T}(\varepsilon_{0} - \mu)^{2}\sigma(\varepsilon_{0},T) \end{equation} | (13) |
\begin{equation} \therefore z_{\text{e}}T = \frac{S(\mu,T)^{2}\sigma (\mu,T)T}{\kappa_{0}(\mu,T) + \kappa_{\text{P}}(\mu,T)} \xrightarrow{| \varepsilon_{0} - \mu | \gg \varDelta \varepsilon}\left[ \begin{array}{l} 1\ (\text{$\kappa_{0}$ only})\\ \infty\ (\kappa_{0} + \kappa_{\text{P}}) \end{array} \right. \end{equation} | (14) |
Although knowing zelT is useful for predicting the upper limit for zT, we need to estimate the actual values of zT by including the contribution of the ignored term κph/τel. A theoretical estimation of this parameter is difficult, since both κph and τel are expected to strongly depend on the material and its microstructure since phonons and electrons scatter at impurities, grain boundaries, and crystal defects.
As shown in Fig. 5, the introduction of κph/τel results in a large reduction of zT at low n. The strangely high zelT at low n only occurs when neglecting κph, as it is the dominant term of κ in the low-n region. Therefore, the value of zelT should not be used to estimate zT in the low-n region. However, at high n the value of zelT is more related to zT, because then zT is less influenced by introducing κph/τel as κel is dominant. This unknown term κph/τel can be understood to represent the material’s degree of phonon-glass electron-crystal, a concept that was first proposed by Slack.42)
Dependence of zT on the carrier doping level for different values of κph/τel in [W m−1 K−1 s−1] for (a) p-type Si at 300 K, (b) n-type Si at 300 K, (c) p-type Si at 900 K, and (d) n-type Si at 900 K. The data for κph/τel = 0 are presented as zelT in Fig. 4.
It is difficult to calculate κph and τel from first-principles calculations. Even if the values of κph and τel of pure crystals were calculated, these values will be decreased in the actual samples owing to extrinsic electron and phonon scattering mechanisms. In this study, we therefore analyzed whether κph and τel should instead be obtained by comparing the calculation results with the experimental values.
By combining the calculation results with the reported experimental values for Sexp(T), σexp(T), and κexp(T) we first tried to find the dataset of values of εF that best agrees with the experimental data. This method is useful when the carrier density is not available, although it requires that the calculation reproduces the actual electronic structure. The Boltzmann transport calculation can provide thousands of datasets for different values of (εF, T). The parameter ncalc, which is the doped carrier concentration, is determined by εF. The other parameters Scalc, (σ/τel)calc, and (κel/τel)calc are all obtained as functions of (εF, T). We searched for the εF that best reproduced Sexp and nexp. Instead of using the value of nexp from Hall measurements, the corresponding dataset for T was found by searching for a dataset with Scalc = Sexp, because in general nexp at higher T was not reported. Two or more datasets were found to have Scalc = Sexp when the S–n curve peaked; we selected the dataset with the most reasonable value for n.
Figure 6(a) shows the T-dependence of εV-μ for p-type and εC-μ for n-type semiconductors that was estimated by comparing the value of S calculated using TB-mBJ to the values of Sexp of three samples of Si0.8Ge0.2 alloys that were reported by Vining et al.22) and Dismukes et al.41) Here, the DOS of Si0.8Ge0.2 and Si was assumed to be the same. Vining’s p-type (No. 75) and n-type (No. 93) samples were obtained via ball-milling of the alloys followed by hot-pressing. A Dismukes’ n-type sample (No. 1834) was obtained using the zone-levelling crystal growth technique so that the sample would have a higher crystallinity and larger grains compared with the Vining’s samples. Since εV-μ for the p-type semiconductor yielded almost 0 meV and εC-μ for the n-type yielded about 6 meV, we can see that μ is located almost at the band edge, as expected for semiconductors near the degenerate regions. The n estimated from S in Fig. 6(b) was on the order of 1020 cm−3, which is consistent with the experimental values. The cause of the slight increase with T thus appears to be either calculation errors or the thermal excitation of localized carriers.
Temperature dependence of (a) the chemical potential, where εV is the top of the valence band and εC is the bottom of the conduction band; (b) the carrier doping level; (c) the effective relaxation time; (d) the inverse of the effective relaxation time, with linear fits of the data points; and (e) the phonon thermal conductivity in comparison with κtotal–L0σT (displayed as pale lines and symbols) that was estimated using experimental values of S, σ, and κ of 0.08% boron-doped Si0.8Ge0.2 alloys reported in the literature. The ball-milled, pressure-sintered samples prepared by Vining et al.22) is shown as BM, and the zone-levelled, high-crystallinity sample by Dismukes is shown as ZL.
The effective relaxation time $\tau _{\text{el}}^{\text{eff}}$ was obtained as
\begin{equation} \tau_{\text{el}}^{\text{eff}} = \frac{\sigma_{\text{exp}}}{(\sigma/\tau_{\text{el}})_{\text{calc}}} \end{equation} | (15) |
\begin{equation} \frac{1}{\tau_{\text{el}}^{\text{eff}}} = \frac{1}{\tau_{\text{el}}^{\text{ideal}}} + \frac{1}{\tau_{\text{el}}^{\text{ext}}}, \end{equation} | (16) |
Figure 6(c) shows the T-dependence of $\tau _{\text{el}}^{\text{eff}}$ estimated for the three samples. All samples exhibited a value of $\tau _{\text{el}}^{\text{eff}}$ on the order of 10−15 s, while a monotonic decrease of $\tau _{\text{el}}^{\text{eff}}$ was observed with increasing T. We also estimated the value of τel of the other thermoelectric materials by using the values of Sexp, σexp, and κexp at their best values of zT. Although the calculations were not reliable enough, the values of τel were estimated to be between 10−15 and 10−14 s; single crystals43) tended to exhibit longer τel values than polycrystalline samples.44)
This analysis of $\tau _{\text{el}}^{\text{eff}}$ enabled us to determine the dominant scattering mechanism in the samples. Figure 6(d) shows that in the two n-doped samples $(\tau _{\text{el}}^{\text{eff}})^{ - 1}$ and T had a linear relationship to each other. Therefore, the dominant scattering mechanism at high T in these samples was expected to be scattering caused by an acoustic distortion potential (in metals), where τel is proportional to T−1. The scattering mechanism in the p-type sample appears to be more independent of T.
In Fig. 6(e), the effective phonon thermal conductivity was calculated according to
\begin{equation} \kappa_{\text{ph}}^{\text{eff}} = \kappa_{\text{exp}} - \tau_{\text{el}}^{\text{eff}}(\kappa_{\text{el}}/\tau_{\text{el}})_{\text{calc}}, \end{equation} | (17) |
The sources of errors in first-principles calculation of the thermoelectric properties of materials can be classified into three categories: over-simplification of the crystal structure, inaccuracy in the electronic structure calculation, and inaccuracies in the transport calculation.
An over-simplification of the crystal structure can occur in calculations of almost any thermoelectric material since the highest zT values are usually achieved in extremely complex structures.1) For carrier concentration control and phonon scattering, the crystals are solid solutions of various impurity elements that are randomly distributed throughout the crystal. Defects and disorders can act in a similar way, making it impossible to define unit cells. A number of thermoelectric materials have incommensurate or misfit-layer structures, which are difficult to define by periodic unit cells. The presence of such structures would modify the electronic structure near the band gap so that it is different from that of the parent compound.
The second type of error is related to the electronic structure calculations. Underestimation (and sometimes, overestimation) of εg is an unavoidable error in DFT. This error is critical for narrow-gap semiconductors. However, if the bandgap was at least open, the scissors-cut operation of the band energies in BoltzTraP8) could be used to improve the accuracy.21) The experimental values of the band gaps can be obtained from databases such as citrination.45) The use of other high-cost exchange-correlation potentials that reproduce εg may also be helpful, although the errors in m* need to be small enough. The introduction of various interactions, such as spin-polarizations for magnetic elements and spin-orbit interactions for heavy elements, is also necessary to reduce errors. Magnetic insulators exhibit a finite εg only in spin-polarized calculations, and Mott insulators may require an electron localization potential U to open the band gaps. Spin-orbit interactions should be dominant in many heavy-element thermoelectric materials; these split the bands into two similar bands separated by several hundred millielectronvolts. At the crossing-points of those bands, band reconstruction occurs to generate bands with small m* that are separated by a small gap. In some compounds, the electronic structures are sensitive to the lattice volume and to atomic positions. In such cases a careful selection of the calculation conditions is necessary.
The third type of error is related to the Boltzmann transport calculations. This type of error can be usually be mitigated by increasing the density of the k-point mesh. In BoltzTraP, the use of a coarse k-point mesh results in oscillations and noise in the thermoelectric properties at high n. The introduction of anisotropy would improve the calculations of anisotropic compounds. A more precise treatment of τel, which should be dependent on energy and bands, may also improve accuracy of the Boltzmann transport calculations.
We investigated how the Boltzmann transport calculation of the thermoelectric properties of a typical semiconductor (Si) from a first-principles calculation is affected by unknown variables such as n, εg, and κph/τel. We determined that to achieve a high zT the parent compound should have a wide εg, low κph, and long τel. We demonstrated that calculated thermoelectric properties such as S, σ/τel, κel/τel, zelT, and zT can vary by orders of magnitude when n is varied. The prediction of an extremely high value for S was not found to be special, as it was achieved in a normal semiconductor with a wide enough εg in the low-n limit. These properties were strongly affected by εg, even though the estimation of εg is difficult in DFT. This error is critical for thermoelectric materials, because narrow-gap semiconductors are favored for ease of carrier doping. We also found that κP is required for calculations of κel when we need to show that zT can exceed 1. The calculated Leff was nearly equal to L0 in the metallic limit of n > 1021 cm−3; however, it was ∼20% lower than L0 when n < 1020 cm−3 and increased by orders of magnitude when bipolar conduction became dominant. The values of zT were largely dependent on κph and τel, although both parameters are strongly influenced by typical engineering processes such as impurity doping and microstructural control.
We attempted to estimate the T-dependence of μ, n, τel, and κph by fitting the calculation results with the T-dependences of Sexp, σexp, and κexp provided in experimental papers on the thermoelectric properties of certain materials. We showed that τel was on the order of several 10−15 s and varied with T as expected for the dominant electron scattering mechanism. It was also shown that the values of τel differed between samples fabricated in different ways.
To predict semiconductor’s thermoelectric properties correctly, it is necessary to reduce many errors in the values of parameters related to the crystal structures, electronic structure calculations, and Boltzmann transport calculations. We also found that it is necessary to careful check that the calculated results are in agreement with the experimental data for any calculations of the thermoelectric properties of semiconductors from first-principles.
This work was supported by KAKENHI grants (grant numbers 23760647 and 16K14379) from the Ministry of Education, Culture, Sports, Science and Technology and by the “Materials research by Information Integration” Initiative (MI2I) project of the Support Program for Starting Up Innovation Hub from Japan Science and Technology Agency (JST). We thank Jim Bailey, PhD, from Edanz Group (www.edanzediting.com/ac) for editing a draft of this manuscript.