1. Introduction
Bismuth telluride (Bi
2Te
3) has been the focus of extensive theoretical and experimental studies as a component of materials for thermoelectric (TE) devices, such as solid-state coolers or generators [
1,
2,
3]. The performance of a thermoelectric material in the aforementioned applications is evaluated in terms of a dimensionless figure of merit ZT, which is defined as (S
2σ/κ)T; where S is the Seebeck coefficient (or thermopower), σ is the electrical conductivity, κ is the thermal conductivity, and T is the temperature [
4]. The product (S
2σ) is called the power factor. A larger ZT leads directly to a higher conversion efficiency. The main challenge lies in the decoupling of the interdependent thermoelectric parameters (S, σ, and κ), which are strongly coupled to the carrier concentration. Commercial TE devices comprise series of
p- and
n-type semiconductor pairs.
The ZT values of commercial Bi
2Te
3 compounds are about 1.35 for
p-type and 0.9 for
n-type materials [
5]. The poor performance of
n-type Bi
2Te
3 based materials compared to that of
p-type materials seriously inflicts a limitation on making it a more efficient TE device. Both
p-type and
n-type characteristics of Bi
2Te
3 can be controlled depending on the chemical composition. As is well known,
n-type Bi
2Te
3 have been synthesized by making solid solution with Bi
2Se
3, or addition of excess tellurium as an electron donor [
6,
7]. However, the fabrication of
n-type Bi
2Te
3 thermoelectric materials has a number of technical problems, such as controlling the Se content in Bi
2Te
3-Bi
2Se
3 solid solution is difficult and Te-rich Bi
2Te
3 easily decomposes upon heating. Element doping is a more effective approach to enhance the thermoelectric properties of Bi
2Te
3-based alloys [
8,
9,
10,
11,
12]. Among various dopants, Cu or Cu-halide acts as an excellent additive for improvement of thermoelectric performance of
n-type Bi
2Te
3 [
13,
14,
15,
16]. Cu atoms can be either an acceptor or a donor depending on their location in the compound. Cu is also known to improve the reproducibility of thermoelectric materials, due to the formation of Cu–Te bond in the van der Waals gaps, which suppress the escape of Te atoms [
17]. The Cu-intercalated Bi
2Te
3 bulk shows a significantly enhanced ZT of ~1.12 at 300 K [
13], which is the highest ZT value reported for
n-type Bi
2Te
3 binary material. Cu addition can also prevent the oxidation of the Bi
2Te
3 [
17]. However, the thermoelectric properties of these alloys change with aging time [
18]. Studies of the structure and properties of crystals with a co-dopant with Cu content were carried out [
16,
19,
20]. Cu and I atoms co-doped Bi
2Te
3 was prepared using the Bridgman method for the improvement of its corresponding thermoelectric properties, whereby the power factor was improved through the co-doping effect of Cu and I, while its thermal conductivity was reduced by forming dispersed Cu-rich nanoprecipitates. The maximum ZT of ~1.16 was achieved at a temperature of 368 K for (CuI)
0.01Bi
2Te
3 [
16]. Moreover, various dopants (Au, Mn, Co, Ni, Zn, Ge, Ag, In, Sc, Ti, V, and Sn) in Cu
0.008Bi
2Te
2.7Se
0.3 have been studied [
19,
20]. The addition of dopant atoms at Bi sites in
n-type Cu-intercalated Bi
2Te
3 changes the electronic band structure, such as band position and band degeneracies, resulting in an increase of the Seebeck coefficient. As a consequence, peak
ZT values of 0.88 at 360 K and 0.91 at 320 K were obtained for V-doped and Au-doped Cu
0.008Bi
1.98Te
2.7Se
0.3, respectively [
19,
20]. Therefore, it has great potential to further improve the ZT value of
n-type Bi
2Te
3 based materials via compositional tuning approach by adjusting Cu contents or element doping.
In the present study, CuI-Pb co-doped Bi2Te3 samples were prepared using high temperature solid state reaction method and were consolidated by spark plasma sintering (SPS). The Pb-addition effects on the crystal lattice, the charge transport, and the thermoelectric properties of CuI-doped Bi2Te3 were evaluated.
2. Results and Discussion
Powder X-ray diffraction (PXRD) patterns of
x% CuI-Pb co-doped Bi
2Te
3 (
x = 0.01, 0.03, 0.05, 0.07, and 0.10) samples are shown in
Figure 1a. As a comparison, undoped Bi
2Te
3 and
x% CuI-doped Bi
2Te
3 were prepared under the same synthetic conditions. All of the diffraction peaks are indexed to rhombohedral Bi
2Te
3 structure with the space group of R
3m (JCPDS, No. 15-0863) [
21], with no indication for the existence of a second phase for samples with up to 7% of dopant concentration. Trace amounts of possible impurities including Cu
2−xTe, and CuI were detected in the 10% CuI-Pb co-doped Bi
2Te
3 samples. This result implies that the solubility limit of CuI and Pb dopants in Bi
2Te
3 was x < 0.1. While in a previous report, the impurity phase was observed in less than 5% in CuI-doped Bi
2Te
3 sample [
16], when Pb atoms are co-doped with CuI in Bi
2Te
3, the impurity phase was observed only in a 10% CuI-Pb co-doped Bi
2Te
3 sample. This result indicates that the solubility of CuI in Bi
2Te
3 is increased by addition of Pb atoms.
Figure 1b shows the lattice parameters of CuI-Pb co-doped and CuI-doped Bi
2Te
3 samples as a function of the dopant fraction. In all of the samples, the in-plane parameter
a remains constant, while the unit cell parameter
c along the stacking direction expands with an increasing Pb content in the CuI-Bi
2Te
3 system. The result is presumably a consequence of Cu atoms entering into the interstitial site, which increases the distance between the van der Waals layers [
13]. A comparison of the covalent radius of Pb (r
Pb = 0.147 nm) with that of Bi (r
Bi = 0.146 nm) shows that the size of Pb is very close to that of Bi, and thus the ability of Pb atoms for the substitution of Bi atoms in Bi
2Te
3 should not be neglected. Halogen atoms such as I (r
I = 0.220 nm) are believed to occupy Te (r
Te = 0.221 nm) sites in the lattice [
22]. The incorporation of iodine atoms to Te sites and Pb atoms to Bi sites drive the changes in bonding parameters. The substituted atoms can bridge two neighboring quintuple layers, thus weakening the interface scattering. Such an analysis exceeds the scope of this paper and would demand quantum chemical calculations of bonding parameters, which will be the aim of our next work.
In our previous work, we demonstrated that doping of Bi
2Te
3 samples with 1% CuI enhanced ZT [
16]. Thus, we selected 1% CuI-doped Bi
2Te
3 sample as a reference material to demonstrate the effect of CuI-Pb co-doping on the charge transport properties. The charge transport properties of 1% CuI-Pb co-doping Bi
2Te
3 at room temperature are investigated by Hall effect analysis and are compared to those of 1% CuI-doped Bi
2Te
3 and undoped Bi
2Te
3. Assuming one carrier type and parabolic bands in our analysis, the carrier concentration (
n) was calculated from the room temperature (i.e., well within a single-carrier dominated transport) Hall constants using the relationship
RH = 1/
ne, where
RH is the Hall coefficient,
n is the carrier concentration, and
e is the electronic charge. The Hall coefficients of specimens are negative, indicating
n-type conductions. By incorporating Pb in CuI- Bi
2Te
3 system, the
ne value of the bulk samples decreases from ~7.8 × 10
19/cm
3 (CuI-doped Bi
2Te
3) to ~3.6 × 10
19/cm
3 (CuI-Pb co-doped Bi
2Te
3), and the corresponding mobility value increases from ~164.6 cm
2/V∙s to ~216.9 cm
2/V∙s at 300 K. In comparison, the undoped Bi
2Te
3 sample shows the
n value of ~1.2 × 10
19/cm
3and the mobility of 354.9 cm
2/V∙s at 300 K. This result verifies that the addition of a small amount of Pb significantly decreases the carrier concentration, which should be attributed to the holes generated by the Pb atoms. This demonstrates that facile control of electron concentration can be easily realized by adding Pb atoms to CuI-doped Bi
2Te
3 system, yielding an optimal electron concentration of 3−4.5 × 10
19/cm
3.
Figure 2a shows SEM images of the fractured surfaces of SPSed undoped Bi
2Te
3, 1% CuI-doped Bi
2Te
3, and 1% CuI-Pb co-doped Bi
2Te
3. All of the samples exhibit lamellar structures at the micron scale and no obvious large-scale preferred orientation. The microstructures are dense (>98% of the theoretical density of
n-type Bi
2Te
3 (7.86 g/cm
3) showing densities of 7.73 g/cm
3, 7.77 g/cm
3, and 7.82 g/cm
3 for undoped Bi
2Te
3, 1% CuI-doped Bi
2Te
3, and 1% CuI-Pb co-doped Bi
2Te
3, respectively. The orientation degree of the (0 0 l) planes, termed as
F, was calculated with the Lotgering method [
23]. In this method,
F is expressed as the following equations:
F =
P −
P0/1 −
P0,
P0 =
I0(0 0 l)/∑
I0(h k l),
P =
I(0 0 l)/∑
I(h k l), where
I0(0 0 l) is the intensity of (0 0
l) peaks and ∑
I0(h k l) is the sum of intensities of all the peaks for the powders with random orientation;
I(0 0 l) is the (0 0
l) peak intensity and ∑
I(h k l) is the sum of the intensities of all peaks for the measured section. We calculated the ratios
I(0015)/
I(015) of the integrated intensity of (0015) to (015), and represented them in
Figure 2b to evaluate the grain orientation anisotropy. All of the samples show anisotropy in the crystal structure; however, the degree of anisotropic orientation is not significant in SPS consolidated polycrystalline samples. The
I(0015)/
I(015) value for 1% CuI-doped Bi
2Te
3 and 1% CuI-Pb co-doped Bi
2Te
3 (17–18%) is slightly higher than those for undoped Bi
2Te
3 (10%). This indicates that the
c-axis of the grains after SPS was preferentially oriented parallel to the pressing direction. This result is consistent with a previous report [
16], showing the strengthening of the two adjacent quintuple layers by substituting Te with I atoms. Effect caused by sample density or sample orientation is negligible since the relative densities and orientation degree determined by the Lotgering method for CuI-doped and CuI-Pb co-doped Bi
2Te
3 samples are nearly same.
The thermoelectric properties depend on the dopants (here, we use CuI only and CuI-Pb), dopant content, and temperature. In order to elucidate the effect of dopants and their contents on the thermoelectric properties, the electrical conductivity (σ), Seebeck coefficient (
S), and power factor of CuI-doped and CuI-Pb co-doped Bi
2Te
3 (
x = 0, 0.01, 0.03, 0.05, 0.07, and 0.10) system as a function of composition was investigated at room temperature, as shown in
Figure 3. In both series, with increasing dopant concentration, the electrical conductivity increase, while the Seebeck coefficient decreases simultaneously for up to 7% of dopant concentration. The room temperature electrical conductivity of the undoped Bi
2Te
3 (~307 S/cm) is increased by CuI-doping (1% of CuI-doped Bi
2Te
3 sample gave ~2673 S/cm). The room temperature electrical conductivity of 1% CuI-Pb co-doped Bi
2Te
3 at 300 K was about ~1462 S/cm. This value is significantly lower than that of 1% CuI-doped Bi
2Te
3. As shown in
Figure 3a, the CuI-Pb co-doped samples show a lower electrical conductivity than that of CuI-doped samples with similar
x values, confirming the role of Pb as an acceptor [
24]. In
Figure 3b, the Seebeck coefficients at room temperature were plotted as a function of dopant contents. The value of Seebeck coefficient at 300 K for CuI-doped and CuI-Pb co-doped Bi
2Te
3 are about −115 μV/K and −157 μV/K, respectively, while that for undoped Bi
2Te
3 is −270 μV/K, which compares well with the previous reported value for
n-type Bi
2Te
3 [
2]. The Seebeck coefficients of the CuI-Pb co-doped bulk samples are observed to be higher than that of the CuI-doped sample due to lower carrier concentrations. Normally, Bi
2Te
3 shows
p-type character, however the undoped Bi
2Te
3 in this study show
n-type character. We assume that these differences may arise from the different experimental conditions used for the preparation of undoped Bi
2Te
3 crystals. The Bi
2Te
3 prepared by the Bridgman method is a
p-type conductor due to the antisite defect of Bi
Te. However, in this work, the SPS pressed Bi
2Te
3 samples show
n-type characteristics, which arises from the Te vacancy at the interface. This decrease in electrical conductivity and the increase in Seebeck coefficient in co-doped samples can be explained by an increased carrier scattering related to the incorporation of Pb atoms in the CuI-doped lattice and by decreased carrier concentrations caused by Pb atoms, which act as electron acceptors [
24]. As shown in
Figure 3c, the CuI-Pb co-doped samples show higher power factors than CuI-doped samples with similar
x values. The power factors decrease with increasing dopant concentrations. The maximum values of the power factors were observed at
x = 0.01 for both CuI and CuI-Pb co-doped samples. The benefit of Pb incorporation into CuI-doped Bi
2Te
3 was not observed in the power factor because of the trade-off relationship between the Seebeck coefficient and the electrical conductivity. (~35 μW/cm∙ K
2 for 1% CuI-doped Bi
2Te
3; ∼36 μW/cm∙K
2 for 1% CuI-Pb co-doped Bi
2Te
3). This corresponds to an >80% enhancement over the typical value of undoped Bi
2Te
3 (22 μW/cm∙K
2).
Figure 3d represents thermal conductivities (closed symbols for κ
tot and open symbols for κ
latt) as a function of dopant content. As the dopant concentration increased, the total conductivity of CuI-doped and CuI-Pb co-doped Bi
2Te
3 increased due to the increase of the electronic contribution. The total thermal conductivity κ
tot of 1% CuI-Pb co-doped samples (κ
tot ~ 1.4 W/m∙K at 300 K) is slightly lower than that of 1% CuI-doped Bi
2Te
3 (κ
tot ~ 1.5 W/m∙K at 300 K) and undoped Bi
2Te
3 (κ
tot ~ 1.6 W/m∙K at 300 K) due to alloy scattering. The lattice part (κ
latt) of the thermal conductivity can be estimated by subtracting the electronic component (κ
elec) from the measured total thermal conductivity, κ
latt = κ
tot − κ
elec. The electronic component is given by the Wiedemann-Franz relation, κ
elec =
LσT, where
L is the Lorenz number. L is taken to be 1.5 × 10
−8 V
2/K
2 for near-degenerate or degenerate semiconductor [
25,
26]. The lattice thermal conductivity of 1% CuI-Pb co-doped Bi
2Te
3 was 0.66 W/m∙K at 300 K. In contrast to the behavior of κ
tot upon increasing the dopant concentration, κ
latt slightly decreased with increasing dopant concentration. This result demonstrates clearly that the lattice κ
latt is reduced by Pb incorporation through the alloy phonon scattering.
Figure 4 shows the electrical transport properties as a function of measured temperature of
x% CuI-Pb co-doped Bi
2Te
3 (
x = 0.01, 0.03, 0.05, 0.07, and 0.10), when compared with 1% CuI-doped Bi
2Te
3 and undoped Bi
2Te
3. For all of the samples, a monotonic decrease in electrical conductivity with increasing temperature is observed (
Figure 4a), which is indicative of heavily degenerated doping. The variation of the Seebeck coefficient is similar to that of the electrical conductivity, as shown in
Figure 4b. The Seebeck coefficient is negative in the whole temperature range, indicating that the majority of charge carriers are electrons (
n-type). The magnitude of the Seebeck coefficient initially increases and reaches a maximum that is strongly depend on the Pb content
x. The onset of intrinsic conduction (the maxima of the curves) in these samples shifts to a higher temperature with an increasing dopant content. While the
x = 0% sample has its maxima at ~300 K, the 1% and 3% sample have their maximum at ~425 K, and the
x > 5% sample at ~525 K. The maximum value of the Seebeck coefficient (~−176 μV/K) was observed at
x = 0.01 CuI-Pb content at 425 K.
Figure 4c shows the power factors (
S2σ) values as a function of temperature. In this system, the power factor values for the 1% CuI-Pb co-doped ranged from 36 μW/cm·K
2 at 300 K to 20 μW/cm·K
2 at 523 K. The CuI-Pb co-doped sample with
x > 0.03 shows a mild temperature dependence.
Figure 4d shows the temperature dependence of the total thermal conductivity κ
tot of the samples. The κ
tot of all the doped samples firstly decreases due to the increasing phonon-phonon scattering, and then increases when upon further increase of the testing temperature due to the increase of ambipolar thermal contributions arising from the diffusion of electron-hole pairs with the onset of intrinsic contribution [
27].
The dimensionless figure of merit ZT of the samples with different dopant concentration (
x) are shown in
Figure 5a as a function of temperature. The magnitude of the ZT initially increases and reaches a maximum that is strongly dependent on the dopant content
x. When the temperature is above ~400 K, the ZT values decrease due to the appearance of intrinsic excitation at a higher temperature. In this experiment, the ZT
max of the 1% CuI-Pb co-doped sample was about 0.96 at 370 K, while the highest value ZT
max was about 0.96 at 422 K for the 1% CuI-doped sample. The incorporation of Pb into CuI-doped Bi
2Te
3 led to a shift of the peak position of ZT
max to lower temperatures. This result shows that the optimization of the operating temperature can be controlled by co-doping. For practical applications of thermoelectric materials, the ZT values at room temperature are also important.
Figure 5b shows the room temperature ZT of the samples as a function of the dopant concentration. The undoped Bi
2Te
3 sample shows a low ZT of ~0.42 at 300 K due to its very poor electrical properties. The highest ZT of 0.79 and 0.70 at 300 K were achieved for the 1% CuI-Pb doped sample and 1% Cu-doped Bi
2Te
3 sample, respectively, which are both significantly improved when compared with those of the undoped sample. All of the evidences about electrical and thermal transport properties suggest that the
n-type ZT of Bi
2Te
3 can be enhanced by the incorporation of Pb with CuI dopant. Further improvement in its TE properties can be expected by choosing suitable combination of dopants.