Dianta Gintinga,
Chan-Chieh Lina,
Lydia Rathnama,
Byung-Kyu Yub,
Sung-Jin Kimb,
Rabih Al rahal Al Orabic and
Jong-Soo Rhyee*a
aDepartment of Applied Physics and Institute of Natural Sciences, Kyung Hee University, Yong-In 17104, Korea. E-mail: jsrhyee@khu.ac.kr
bDepartment of Chemistry and Nano Science, Ewha Womans Univeristy, Seoul 03760, Korea
cDepartment of Environmental Science and Engineering, Ewha Womans University, Seoul 03760, Korea
First published on 24th June 2016
We investigated thermoelectric properties in K-doped quaternary compounds of (Pb1−xKxTe)0.70(PbSe)0.25(PbS)0.05 (x ≤ 0.03). In terms of two valence bands model, we argue that the L-band approaches to the Σ-band with increasing the K-doping concentration resulting in the increase of carrier concentration and effective mass of carrier due to the increase of band degeneracy. The effective K-doping by x = 0.02 and PbS substitution causes high power factor and low thermal conductivity, resulting in the comparatively high ZT value of 1.72 at 800 K. The low thermal conductivity for (Pb0.98K0.02Te)0.70(PbSe)0.25(PbS)0.05 compound is attributed from the lattice distortion and line dislocation in a phase of nano precipitation. By optimizing K-doping and PbS substitution, we achieved the enhancement of practical thermoelectric performance such as average ZTavg = 1.08, engineering (ZT)eng = 0.81, maximum efficiency ηmax = 11.63%, and output power density Pd = 6.3 W cm−2, with temperature difference ΔT = 500 K, which has practical applicability in waste heat power generation technologies.
Lead tellurides (PbTe) and their alloys have a simple face-centered-cubic (FFC) rock salt structure and have been studied over several decades for thermoelectric applications at the intermediate temperature range (600–800 K).13 Nanostructuring by phase separation and precipitation in PbTe matrix causes the reduction of lattice thermal conductivity, resulting in the enhancement of thermoelectric performance.14,15 In addition, high thermoelectric performance of PbTe can be achieved through the precise control of hole concentration in p-type PbTe (from 1018 to 1020) by substitution of monovalent atoms such as thallium (Tl),2 sodium (Na)16 on the lead sublattice. The Tl- or Na-doping in PbTe increases the density of states (DOS) by controlling the Fermi level close to the heavy hole band17 which attributes to the enhancement of the Seebeck coefficient.18–20
Combining the nano structuring with the Fermi level tuning by Na-doping, high thermoelectric performance was revealed in Na-doped binary composites, for example, PbTe–PbS (ZT = 1.8 at 800 K) and Na-doped PbTe–PbSe (ZT = 1.8 at 850 K) as result of nano structuring with sulfur and band engineering by alloying selenide, respectively.21,22 In order to maximize the performance, complex quaternary composites of Na-doped PbTe–PbSe–PbS were investigated. Quaternary compounds of (PbTe)1−2x(PbSe)x(PbS)x (x = 0.07) exhibited ZT ≈ 2.0 at 800 K due to the valence band modification and phonon scattering from alloy scattering and point defects.23 The Na-doped PbTe0.9−xSe0.1Sx compounds had a high ZT ≈ 1.6 near 750 K.24
Theoretical investigations suggested that power factor of PbTe based compounds can be enhanced by resonant like DOS distortions in p-type Pb1−xAxTe (A = K, Rb, and Cs but not Na).25 However, the other theoretical study showed that K- and Tl-doping do not form the resonance state but can control the energy difference of the maxima of two primary valance sub-bands in PbTe.17,26 The K-doped PbTe compounds showed relatively high ZT values reaching 1.3 at 673 K (ref. 27) which is comparable with those of Na-doped PbTe at the same temperature range.18 Furthermore, a high thermoelectric performance was obtained in binary compounds of K-doped PbTe1−xSex (ZT ≈ 1.6 at 773 K) due to the increase of DOS near the Fermi level, resulting in the enhancement of Seebeck coefficient for band convergence of two valance bands in PbTe1−xSex.27
From the valance band converge of PbTe–PbSe22,27 and the nano precipitation in PbTe–PbS,21 here we examine the effect of K-doping in quaternary system of PbTe–PbSe–PbS composites in order to maximize ZT value by combining of the band converge and nano precipitation. We found that the K-doping enhances the Seebeck coefficient via heavy doping, which increases the DOS near the Fermi level. Combined effect of low lattice thermal conductivity due to nanostructure and dislocation with high power factor, we obtained a high ZT ≈ 1.73 at 750 K for (Pb0.98K0.02Te)0.70(PbSe)0.25(PbS)0.05, which is an 32% improved value comparing with the one of K-doped PbTe and K-doped PbTe1−xSex compounds.27 In addition, we analysed the thermoelectric performance in terms of newly established parameters, such as the engineering ZTeng, device efficiency η, and so on.
We loaded the synthesized precursor powders of PbSe and PbS and the elements Pb (99.999%), Te (99.999%), and K (99.99%) in carbon-coated fused silica tubes and sealed them under vacuum at 10−4 Torr. The elements and powders sealed in the quartz ampoules were heated to 1100 °C, kept at the temperature for 10 hours, and were then quenched in ice cold water. The ice water quenched ingots were annealed at 500 °C for 72 hours. The ingots were hand ground and hot pressed in a graphite mold having a 12.7 mm diameter using a hot press sintering machine under a uniaxial pressure of 40 Mpa for 1 hour at 500 °C under vacuum environment.
The Seebeck coefficient S and electrical conductivity σ were measured by the four-point probe method using a thermoelectric measurement system (ZEM-3 ULVAC, Japan). The isothermal Hall resistivity ρxy was measured by five-point contact method with sweeping magnetic fields from −5 T to 5 T using a physical property measurement system (PPMS Dynacool-14T, Quantum Design, USA). The Hall coefficient RH, Hall carrier density nH, and Hall mobility μH were obtained using the relations of RH = ρxy/H, nH = −1/(RHe), and μH = 1/(ρnHe), respectively.
Thermal conductivity κ was obtained by the relation κ = ρsλCp where ρs, λ, and Cp are sample density, thermal diffusivity, and specific heat, respectively. The ρs, λ, and Cp were measured by Archimedes method, laser flash method (LFA-447, NETZSCH, Germany), and high temperature fitting of the Dulong–Petit law by measuring specific heat using a physical property measurement system (PPMS Dynacool-14T, Quantum Design, USA), respectively. The combined uncertainty for all measurements involved in ZT determination was around 20%.
TEM were prepared by Nova 600 nanolab Focused Ion Beam (FEI) using 5–30 kV. TEM images (High Resolution images/STEM/ED pattern) were collected using a JEOL 2100F at 200 kV. Energy dispersive X-ray spectrometer (EDS) analysis were obtain using Oxford Instruments (INCA platform) detector equipped on JEOM 2100F.
Fig. 1 Powder X-ray diffraction of (Pb1−xKxTe)0.70(PbSe)0.25(PbS)0.05 compounds (x = 0.010, 0.015, 0.020, 0.025, and 0.030). |
x | Pb (at%) | K (at%) | Pb | K |
---|---|---|---|---|
0.01 | 42.17 | 2.69 | 0.94 | 0.06 |
0.015 | 47.48 | 7.58 | 0.86 | 0.14 |
0.02 | 47.81 | 6.83 | 0.87 | 0.13 |
0.025 | 38.2 | 2.39 | 0.94 | 0.06 |
0.03 | 49.5 | 9.31 | 0.84 | 0.16 |
In order to identify the K-doping concentration in the compounds, we measured X-ray photoelectron spectroscopy (XPS) for the whole series compounds of (Pb1−xKxTe)0.70(PbSe)0.25(PbS)0.05 as shown in Fig. 2. Fig. 2(a) represent full energy scan of the compounds. The peaks represent the chemical elements of Pb 4f and 4d5/2, Te 3d3/2, S 3p3/2, Se 3d5/2, K 2p3/2, and other impurities such as O s- and C 1s-orbitals. The oxygen and carbon impurities come from the surface oxidation and contamination which is widely observed in XPS measurement.
Fig. 2 Full scan of the XPS spectra (a) and expanded region near the K 2P3/2 core level (b) for (Pb1−xKxTe)0.70(PbSe)0.25(PbS)0.05 compounds (x = 0.01, 0.015, 0.02, 0.025, and 0.03). |
Fig. 2(b) is the expanded scale near the K 2P3/2 core level for the compounds. From the integration of the peaks for each elements, we quantified the atomic concentration of the elements. Table 1 listed the atomic (at%) and relative concentrations of Pb and K elements for the initial stoichiometric compounds of Pb1−xKxTe0.7Se0.25S0.05 (x = 0.01, 0.015, 0.02, 0.025, and 0.03). It shows the non-systematic concentration of K with respect to initial compositions. It is not surprising because XPS is a surface sensitive measurement tool and the K-doping concentration is relatively small. In spite of that, it seems that the potassium does not be distributed homogeneously.
From the room temperature Hall resistivity measurement, we obtained the Hall carrier density nH and Hall mobility μH of p-type (Pb1−xKxTe)0.70(PbSe)0.25(PbS)0.05 system. Fig. 3 shows the Hall carrier density (left axis, black closed square) and Hall mobility (right axis, blue open square) at room temperature of (Pb1−xKxTe)0.70(PbSe)0.25(PbS)0.05 system as function of K-doping concentration x. The Hall carrier concentration is abruptly increased at K-doping range x ≥ 0.02 from 4.33 × 1019 cm−3 (x = 0.02) to 7.08 × 1019 cm−3 (x = 0.025), which indicates that the K-doping tunes the carrier concentration in (Pb1−xKxTe)0.70(PbSe)0.25(PbS)0.05 system, similarly observed in (PbTe)1−x(PbSe)x (x = 0, 0.1, 0.25 and 0.75) compounds.28 The maximum Hall mobility is found μH = 303.2 cm2 V−1 S−1 at x = 0.1 and is decreased with increasing K-doping concentration due to the increase of interfacial and defect scattering of carriers.29
In order to understand the K-doping effect on thermoelectric performance, we measured temperature-dependent Seebeck coefficient S(T) and electrical resistivity ρ(T) from 300 K to 800 K of quaternary (Pb1−xKxTe)0.70(PbSe)0.25(PbS)0.05 compounds and K-doped Pb0.98K0.02Te (black square) for comparison as shown in Fig. 4(a) and (b), respectively. The compounds show a typical behaviour of degenerated semiconductor which S(T) and ρ(T) are increased with increasing temperature. The Seebeck coefficients show a broad maximum at high temperature range (T ≥ 700 K) depending upon the K-doping concentration as shown in Fig. 4(a). The S(T) is not changed significantly with respect to K-doping concentration except the x = 0.02 K-doped one and Pb0.98K0.02Te.
Fig. 4 Temperature dependence of (a) Seebeck coefficient S, (b) electrical resistivity ρ, (c) power factor S2/ρ of (Pb1−xKxTe)0.70(PbSe)0.25(PbS)0.05 and Pb0.98K0.02Te compounds. (d) Room temperature Pisaranko plot of Seebeck coefficient S versus Hall carrier concentration nH for (Pb1−xKxTe)0.70(PbSe)0.25(PbS)0.05 compounds as well as p-type PbTe,18,33 Pb1−xKxTe,27 and Na-doped PbTe18 compounds based on the two band model. |
From the Seebeck coefficient, we can obtain a rough estimation of the effective mass of carrier based on a single band model:30
Stotal = (σLSL + σΣSΣ)/(σL + σΣ) |
ΔEC–Σ = 0.36 + 0.10x |
x | a (Å) | nH (×1019 cm−3) | μH (cm−2 V−1 S−1) | S (μV K−1) | m* (me) |
---|---|---|---|---|---|
0.01 | 6.355 | 4.932 | 303.487 | 46.997 | 0.309 |
0.015 | 6.353 | 4.875 | 268.894 | 33.492 | 0.220 |
0.02 | 6.353 | 4.336 | 224.972 | 51.401 | 0.310 |
0.025 | 6.356 | 7.720 | 166.440 | 41.106 | 0.365 |
0.03 | 6.355 | 7.085 | 123.253 | 40.436 | 0.339 |
The energy band gap ΔEg between the conduction band and L-valence band increases while the extremum valence band difference ΔEv between the L- and Σ-bands decreases with increasing temperature.22,32 Because the L- and Σ-bands come closer to each other with increasing temperature, holes transfer from the L-band to the Σ-band, which is the cause of the broad maximum near 700 K. The electrical resistivity shows the similar trend with Seebeck coefficient as shown in Fig. 4(b), which means the metallic behaviour with temperature. In terms of two valance bands model, the L-band approaches to the Σ-band with increasing K-doping concentration. The two bands overlap increases carrier concentration as well as effective mass of carrier due to the high density of states. Table 2 shows the distinguishably increase of carrier concentration and effective mass at x ≥ 0.02, implying the band degeneracy between two bands.
Fig. 4(c) presents the power factor S2/ρ of (Pb1−xKxTe)0.70(PbSe)0.25(PbS)0.05 system and Pb0.98K0.02Te for comparison. The maximum power factor is 27.78 μV cm−1 K−2 at 600 K for x = 0.01. The high power factor for the compound is due to the significant reduction in electrical resistivity. Fig. 4(d) presents room temperature Pisarenko plot (Seebeck coefficient vs. Hall carrier concentration) based on two band model for (Pb1−xKxTe)0.70(PbSe)0.25(PbS)0.05 compounds as well as p-type PbTe,18,33 Pb1−xKxTe,27 and Na-doped PbTe18 compounds. Two band model for p-type PbTe was calculated based on m*h = 1.2 mo and m*1 = 0.36 mo and includes band gap ΔE ≈ 0.12 eV at room temperature. The (Pb1−xKxTe)0.70(PbSe)0.25(PbS)0.05 system follows the Pisarenko plot. The flattening of the Seebeck coefficient with increasing carrier concentration indicates the presence of the second valance band.23,33 The experimental data points are below the dashed line of two band model, which is due to lower effective mass on the compounds than p-type PbTe. Based on the Pisaranko plot in Fig. 4(d), we do not find any indication of the existence of resonance scattering.
Temperature-dependent total thermal conductivity κ(T) of K-doped compounds decreases with increasing temperatures as shown in Fig. 5(a) which is the typical behavior of acoustic phonon scattering. Even though the K-doping effect does not show a systematic behavior, the lowest thermal conductivity is obtained for x = 0.2 K-doped compound (1.9 W m−1 K−1 at 300 K). The significant decrease of thermal conductivity of the compound than those of Pb0.98K0.02Te (3.4 W m−1 K−1 at 300 K) may come from the additional multiple elements doping by Se and S which will be discussed in detail later. The κ(T) for the samples of x ≤ 0.02 weakly increases with increasing temperature at the high temperature region (T ≥ 700 K). The small upturn of total thermal conductivity is also found in Na-doped (PbTe)1−x(PbS)x compound indicating small bipolar contribution of carriers.34
Fig. 5 Total thermal conductivity κ (a) and lattice thermal conductivity κL (b) of (Pb1−xKxTe)0.70(PbSe)0.25(PbS)0.05 and Pb0.98K0.02Te compounds. |
The lattice thermal conductivity κL(T) of the compounds (Pb1−xKxTe)0.70(PbSe)0.25(PbS)0.05 is shown in Fig. 5(b) by subtracting the electronic thermal conductivity κe from the Wiedermann–Franz law: κe = LT/ρ, where ρ, T, and L are the electrical resistivity, absolute temperature, and the Lorenz number. The temperature-dependent Lorenz number is calculated within an assumption of a parabolic band model with acoustic phonon scattering.29 L is expressed by the Fermi integral formalism as following relations:35
As a result, the lattice thermal conductivity of the (Pb1−xKxTe)0.70(PbSe)0.25(PbS)0.05 compounds are shown in Fig. 5(b). The lowest lattice thermal conductivity is found in (Pb0.98K0.02Te)0.70(PbSe)0.25(PbS)0.05 with 0.91 W m−1 K−1 at room temperature and 0.69 W m−1 K−1 at 800 K. The lattice thermal conductivity of the compound is strongly reduced as much as 63.4% and 40.64% at room temperature comparing with the previous report of Pb0.98K0.02Te and (Pb0.98K0.02Te)0.75(PbSe)0.25.27 The compounds showing low lattice thermal conductivity (x = 0.01, 0.02, and 0.025) do not follow the T−1 dependence indicating that the additional contribution of phonon scattering attributes to the thermal conductivity as well as acoustic phonon contribution.
In order to clarify the phonon scattering mechanism, we calculate the thermal conductivity based on the PbTe1−xSex alloy model. Theoretical model of lattice thermal conductivity of PbTe1−xSex alloy can be calculated as follows which is given by Klemens:23,36
When we plot the lattice thermal conductivity for various PbTe based compounds including the (Pb1−xKxTe)0.70(PbSe)0.25(PbS)0.05, Pb0.98K0.02Te (in this work) and (Pb0.98K0.02Te)1−x(PbSe)x27 with the theoretical estimation of lattice thermal conductivity in the alloy model as shown in Fig. S2 of ESI,† the experimental lattice thermal conductivity of (Pb0.98K0.02Te)1−x(PbSe)x follows the trend of theoretical of lattice thermal conductivity for (PbTe)1−x(PbSe)x system. Therefore, the strong phonon scattering in (Pb0.98K0.02Te)1−x(PbSe)x comes from the point defect scattering created by Te/Se mixed occupation in rock salt structure. On other hand, the lowest lattice thermal conductivity of (Pb1−xKxTe)0.70(PbSe)0.25(PbS)0.05 compounds lie below the theoretical lattice thermal conductivity of (PbTe)1−x(PbSe)x. It implies that the low lattice thermal conductivity of (Pb1−xKxTe)0.70(PbSe)0.25(PbS)0.05 has another scattering sources as well as the alloy scattering and/or point defect scattering by Te/Se mixed occupation.
One possible phonon scattering mechanism is the nano/micro-grain boundary scattering. We examine the microstructural analysis of (Pb0.98K0.02Te)0.70(PbSe)0.25(PbS)0.05 by high resolution transmission electron microscopy (HR-TEM) as shown in Fig. 6. Fig. 6(a) and (b) show mid-magnified images of (Pb0.98K0.02Te)0.70(PbSe)0.25(PbS)0.05. It displays numerous nano-precipitates with dark contrast for size distribution of 5–15 nm. The electron diffraction pattern, in the inset of Fig. 6(a), of (Pb0.98K0.02Te)0.70(PbSe)0.25(PbS)0.05 sample shows cubic structure. The lattice parameters of PbTe, PbSe, and PbS are 3.22 Å (PbTe), 3.07 Å (PbSe), and 2.965 Å (PbS), respectively, along the (200) plane.34,37 In order to study the nano-precipitates of the compound, we measure high-resolution transmission electron microscope (HR-TEM) images as shown in Fig. 6(c). We can identify the several homogeneous nano-precipitates (dark contrast) in (Pb0.98K0.02Te)0.70(PbSe)0.25(PbS)0.05 with sizes of 2–5 nm in the matrix. From the electron diffraction, we identified the nano-precipitates are composed with (200) and {111} plane as shown in the inset of Fig. 6(c). Fig. 6(d) is the inverse fast Fourier transform (IFFT) image of the Fig. 6(c). The IFFT images along the (200) plane show that there are many line dislocations at the coherent interfaces between the nano-precipitates and PbTe matrix which is marked by red circles.
The nano-precipitates have coherent or semi-coherent interface depending on each of the crystal plane as shown in Fig. 7. Previous reports on the PbTe/PbS system showed that the nano-precipitation was formed by spinodal decomposition or nucleation, depending on composition and temperature conditions, using a miscibility gap in the PbTe/PbS system.38–40 Fig. 7(a)–(c) present high resolution TEM images of [110] plane at different regions. The electron diffraction peaks (inset of Fig. 7(b) and (c)) and HR-TEM images show the identical crystal symmetry. On the other hand, the inverse fast Fourier transform (IFFT) images in Fig. 7(d) and (e) present the line dislocations and small lattice distortion. The line dislocations and lattice distortions may come from the phase separation due to the compositional complexity by multiple elements doping or substitutions.23,41
Although it is difficult to quantitatively determine the composition of precipitates due to overlap with matrix, we investigate the elemental analysis of the matrix and the precipitates by energy dispersive X-ray spectroscopy (EDS) for (Pb0.98K0.02Te)0.70(PbSe)0.25(PbS)0.05. Fig. 8(a) and (b) are the low magnified images of the compound, which clearly show the nano precipitates and spinodal decompositional phase separation in dark regions. By comparing the EDS spectrum 1 (matrix), spectrum 2 (nano-dot), and spectrum 3 (spinodal) regions as shown in the marked regions of Fig. 8(b), we identify the elemental compositions of the compound. We found that the precipitates and spinodal decompositional regions corresponds to the PbS rich phases while the region 1 corresponds to the PbTe matrix as shown in Fig. 8(d).
We analyze the lattice thermal conductivity of (Pb0.98K0.02Te)0.70(PbSe)0.25(PbS)0.05 in terms of various scattering mechanisms based on the Callway model for PbTe and (PbTe)0.70(PbSe)0.25(PbS)0.05. The lattice thermal conductivity can be expressed by the Callaway's model:37,42
Based on TEM studies, for the same frequency, the relaxation time depends mainly on scattering from the nanostructuring precipitates, dislocation boundaries and phonon–phonon interactions. The combined relaxation time is given as:
τ−1c = τ−1U + τ−1N + τ−1B + τ−1S + τ−1D + τ−1P |
The parameters can be obtained from TEM investigations and some from ref. 49.
Fig. 9 shows the theoretical lattice thermal conductivity of Pb0.98K0.02Te (black square) and (Pb0.98K0.02Te)0.70(PbSe)0.25(PbS)0.05 (magenta inverse triangle) based on the Callaway's model (open symbols) with experimental data (closed symbols). The experimental data of Pb0.98K0.02Te (black closed square) fit very well with the theoretical lattice thermal conductivity model of Pb0.98K0.02Te (black open square). However, the experimental lattice thermal conductivity (closed magenta inverse triangle) of (Pb0.98K0.02Te)0.70(PbSe)0.25(PbS)0.05 does not fit to the theoretical values (open magenta inverse triangle). The discrepancy of the lattice thermal conductivity of the compound can be considered by several reasons: (1) the Lorenz number calculation is not exact in nano-structured composites,5,50 (2) an anharmonic phonon excitation (3) partial local collapse of nano-structure at high temperature.40,51 This phenomena was also found in (PbTe)0.88(PbS)0.12 by Bi doping51 and (Pb0.95Sn0.05Te)1−x(PbS)x40 alloys. We surmise that the nano precipitation gives rise to significant phonon scattering resulting in the low lattice thermal conductivity. It needs further investigation to clarify the cause of thermal transport on the compounds.
The apparently low thermal conductivity and high power factor of the nanostructured compounds can have a direct impact on thermoelectric performances. Fig. 10(a) displays dimensionless thermoelectric figure-of-merit ZT of (Pb1−xKxTe)0.70(PbSe)0.25(PbS)x compounds. The compound of x = 0.02 shows the highest ZT value of 1.72 at 750 K, which higher than those of Pb0.98K0.02Te (ZT of 1.11 at 800 K) and (Pb0.98K0.02Te)0.75(PbSe)0.25 (ZT of 1.60 at 773 K)27 as much as 35% and 7.0%, respectively (S3 of ESI†). We also calculated the average ZTavg values of (Pb0.98K0.02Te)0.70(PbSe)0.25(PbS)0.05, Pb0.98Na0.02Te, Pb0.98K0.02Te, (Pb0.98Na0.02Te)0.88(PbS)0.12,21 (Pb0.98Na0.02Te)0.75(PbSe)0.25,22 (Pb0.98K0.02Te)0.75(PbSe)0.25 (ref. 27) and (Pb0.98Na0.02Te)0.84(PbSe)0.07(PbS)0.07.23 The average ZTavg is given by:52
For more practical point of view, the engineering ZTeng predicts more reliable and practical conversion efficiency of thermoelectric.52 The engineering ZTeng is expressed as:52
We compare the (ZT)eng of x = 0.02 doped compound with the K- and Na-doped PbTe alloys21–23,27 in S4(a) of ESI.† It clearly shows that (Pb0.98K0.02Te)0.70(PbSe)0.25(PbS)0.05 has the highest (ZT)eng comparing to the other compounds: such as Pb0.98Na0.02Te (ZTeng ≈ 0.43 at ΔT = 500 K), Pb0.98K0.02Te (ZTeng ≈ 0.50 at ΔT = 500 K), (Pb0.98Na0.02Te)0.88(PbS)0.12 (ZTeng ≈ 0.52 at ΔT = 500 K),21 (Pb0.98Na0.02Te)0.75(PbSe)0.25 (ZTeng ≈ 0.57 at ΔT = 500 K),22 (Pb0.98K0.02Te)0.70(PbSe)0.25 (ZTeng ≈ 0.64 at ΔT = 500 K),27 and (Pb0.98Na0.02Te)0.84(PbSe)0.07(PbS)0.07 (ZTeng ≈ 0.74 at ΔT = 500).23
Using the (ZT)eng, we calculated the individual legs' maximum efficiency (ηmax) based on temperature independence. The calculated the maximum efficiency is defined by52
The engineering power factor (PF)eng determines the output power density for given temperature gradient. Due to higher efficiency and engineering power factor, we expect a high output power density of (Pb1−xKxTe)0.70(PbSe)0.25(PbS)0.05 system. The output power density can derived as:52
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra11299d |
This journal is © The Royal Society of Chemistry 2016 |