Next Article in Journal
Characteristics of Periodic Ultrasonic Assisted TIG Welding for 2219 Aluminum Alloys
Next Article in Special Issue
Phase Formation Behavior and Thermoelectric Transport Properties of P-Type YbxFe3CoSb12 Prepared by Melt Spinning and Spark Plasma Sintering
Previous Article in Journal
Prediction of Tensile Strength in Friction Welding Joins Made of SA213 Tube to SA387 Tube Plate through Optimization Techniques
Previous Article in Special Issue
Insight on the Interplay between Synthesis Conditions and Thermoelectric Properties of α-MgAgSb
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Influence of Pd Doping on Electrical and Thermal Properties of n-Type Cu0.008Bi2Te2.7Se0.3 Alloys

1
Samsung Electronics, Suwon 16678, Korea
2
Department of Materials Science and Engineering, Hongik University, Seoul 04066, Korea
3
Department of Materials Science and Engineering, Yonsei University, Seoul 03722, Korea
4
Department of Materials Science and Engineering, University of Seoul, Seoul 02504, Korea
5
Department of Physics, Kunsan National University, Gunsan 54150, Korea
6
Department of Materials Science and Engineering, Gachon University, Seongnam 13120, Korea
7
School of Energy, Materials and Chemical Engineering, Korea University of Technology and Education, Cheonan 31253, Korea
8
Department of Materials Science and Engineering, Dankook University, Cheonan 31116, Korea
9
Department of Electronic Materials Engineering, Kwangwoon University, Seoul 01897, Korea
*
Authors to whom correspondence should be addressed.
These authors contributed to this work equally.
Materials 2019, 12(24), 4080; https://doi.org/10.3390/ma12244080
Submission received: 18 November 2019 / Revised: 30 November 2019 / Accepted: 4 December 2019 / Published: 6 December 2019
(This article belongs to the Special Issue Novel Thermoelectric Materials and Their Applications)

Abstract

:
Doping is known as an effective way to modify both electrical and thermal transport properties of thermoelectric alloys to enhance their energy conversion efficiency. In this project, we report the effect of Pd doping on the electrical and thermal properties of n-type Cu0.008Bi2Te2.7Se0.3 alloys. Pd doping was found to increase the electrical conductivity along with the electron carrier concentration. As a result, the effective mass and power factors also increased upon the Pd doping. While the bipolar thermal conductivity was reduced with the Pd doping due to the increased carrier concentration, the contribution of Pd to point defect phonon scattering on the lattice thermal conductivity was found to be very small. Consequently, Pd doping resulted in an enhanced thermoelectric figure of merit, zT, at a high temperature, due to the enhanced power factor and the reduced bipolar thermal conductivity.

1. Introduction

Thermoelectric alloys have attracted attention in recent decades because these materials can convert a temperature gradient directly into electrical energy. Bismuth telluride (Bi2Te3)-based alloys are currently the most used bulk thermoelectric alloys near room temperature [1,2]. However, the broader use of Bi2Te3-based alloys is still limited by the rather low thermoelectric conversion performance, evaluated as the thermoelectric figure of merit zT = σ·S2·T/κtot, where σ, S, T, and κtot are the electrical conductivity, Seebeck coefficient, temperature, and total thermal conductivity, respectively. In fact, the zT of n-type Bi2(Te,Se)3 alloys remains below 1, while values significantly higher than 1 have often been reported for p-type (Bi,Sb)2Te3 alloys.
Doping is an effective approach to improving the zT of Bi2Te3 alloys by adjusting the electrical transport properties or reducing the κ through the introduction of additional point defects [3,4,5,6,7,8,9,10,11]. The zT of p-type (Bi,Sb)2Te3 alloys can easily be enhanced using substitutional dopants [3,4,5]. Meanwhile, the influence of doping on n-type Bi2(Te,Se)3 alloys has not been investigated as much as that of doping on p-type (Bi,Sb)2Te3 alloys. It has been found that Cu intercalation in n-type Bi2(Te,Se)3 alloys is a very effective approach to reducing the lattice thermal conductivity (κlatt) by introducing additional point defect scattering centers [12]. However, the accompanying modification of carrier transport properties with the κlatt reduction may reduce the power factor, resulting in zT reduction.
Co-doping of two different substituents was also suggested to further decrease the κlatt while enhancing the power factors in some other thermoelectric materials, such as SnTe and PbTe [13,14,15,16]. In p-type (Bi,Sb)2Te3 alloys, it was found that the co-doping of Ag and Ga reduced the κlatt further compared to that of single doped materials [17], while the power factor can be maintained.
Herein, we investigated the effect of additional Pd substitutional doping on the electrical transport properties and thermal conductivities of Cu-doped n-type Bi2(Te,Se)3, Cu0.008Bi2Te2.7Se0.3. The Pd was anticipated to scatter phonons effectively due to the large mass and ionic radius differences between Pd and Bi (MPd = 106.42 u, MBi = 208.98 u, rPd = 90 pm, rBi = 117 pm). Pd doping increased the electron concentration, electrical conductivity, and power factors. However, the contribution of Pd to additional point defect scattering centers on the lattice thermal conductivity was found to be rather small. As a result, zT enhancement due to Pd doping was observed at high temperatures. To investigate the reason for the limited effectiveness of substitutional Pd doping in reducing the κlatt value in n-type Bi2(Te,Se)3 alloys, the electronic transport properties were analyzed using a single parabolic band model [18], and the reduction in κlatt was quantitatively predicted using the Debye–Callaway model [19].

2. Materials and Methods

The Cu0.008Bi2Te2.7Se0.3 reference sample and a series of Pd-doped Cu0.008PdxBi2-xTe2.7Se0.3 (x = 0.002, 0.004, 0.01, and 0.02) samples were synthesized by a conventional solid state reaction for 10 h at 1423 K, using high-purity (99.999%) raw materials. The synthesized ingots were ball-milled using a 8000M Mixer/Mill high-energy ball mill (SPEX SamplePrep, Metuchen, NJ, USA) for 10 min, and sieved powders under 45 µm were consolidated by spark plasma sintering at 723 K and 50 MPa for 2 min. Then, the temperature-dependent S and σ parameters were measured over the temperature range between room temperature and 480 K in a direction perpendicular to the pressing direction (ZEM-3, Advanced-RIKO, Yokohama, Japan). The carrier concentrations were determined by Hall measurements in van der Pauw configuration, in a magnetic field of 0.5 T (AHT-55T5, Ecopia, Anyang, South Korea) in the same direction. The κ values of the samples were calculated from their theoretical density (ρs), heat capacity (Cp), and thermal diffusivity (λ) values (κ = ρsCpλ), measured along the same direction.

3. Results and Discussion

Figure 1a shows the X-ray Diffraction (XRD) patterns of the investigated series of Cu0.008PdxBi2-xTe2.7Se0.3. All samples showed single phases without impurities. The lattice parameters a and c are shown in Figure 1b, which reveals that the c parameter generally increased with the Pd doping, while a remained largely unchanged upon doping. The systematic change in the c parameter implies that substitutional doping was successfully achieved.
The measured σ and S values of the Pd-doped Cu0.008PdxBi2-xTe2.7Se0.3 (x = 0, 0.002, 0.004, 0.01, and 0.02) are shown in Figure 2a,b. The σ value of the undoped sample was about 740 S/cm at 300 K, and substantially increased to 1320 S/cm for x = 0.02. On the other hand, the magnitude of the S values at 300 K decreased from −192 to −144 µV/K. As a result, the power factor (S2·σ) at 300 K remained unchanged (around 2.73 mW/m·K2) regardless of the Pd doping level (Figure 2c). However, an enhancement in the power factor was observed at high temperatures upon Pd doping. For example, at 480 K the power factor was enhanced by 19%, from 1.68 to 2.00 mW/m·K2.
Figure 3a shows the electron carrier concentration (nH) and mobility (µH) measured for the Cu0.008PdxBi2-xTe2.7Se0.3 samples at 300 K. The nH gradually increased with the Pd doping, with nH values of 2.4, 2.8, 3.2, 3.4, and 4.2 × 1019 cm−3 for x = 0, 0.002, 0.004, 0.01, and 0.02, respectively. On the other hand, the µH values did not change significantly. Therefore, the increase in σ upon Pd doping is mainly due to the increased nH values. Figure 3b shows the Pisarenko plot of the samples, displaying the S of samples as a function of nH at 300 K. The solid lines were obtained for different effective masses (m* = 0.8, 0.9, and 1.0 m0, where m0 is the electron mass) using Equation (1):
S = 8 π 2 k B 2 3 e h 2 ( π 3 n ) 2 / 3 m * T
where e, h, and kB are the elementary charge, Planck’s constant, and Boltzmann constant, respectively. The m* values of all samples, deduced using Equation (1), are plotted in Figure 3b. As shown in the figure, Pd doping resulted in slightly increased m* values, indicating that the electronic structure of the conduction band of Cu0.008Bi2Te2.7Se0.3 was slightly modified favorably for S.
Figure 4a shows the measured κtot of the Pd-doped Cu0.008Bi2Te2.7Se0.3 samples, revealing that the κtot values gradually increased with the doping level. In order to understand these changes, we analyzed the contributions to κtot, given by the following equation:
κ t o t = κ e l e c + κ b p + κ l a t t
where κelec and κbp are the electronic and bipolar thermal conductivities, respectively. First, κelec was calculated using the equation for the Lorenz number (L, expressed as a simple function with S in Equation (3)) [20], and the results are shown in Figure 4b.
L = 1.5 + exp ( | S | 116 )
Equation (3) describes the relationship between the L and S in a simple function, based on a single parabolic band model [20]. The κelec values increased as the electrical conductivity increased with Pd doping, straightforwardly with the increased carrier concentration (Figure 3a). At 300 K, κelec showed a significant increase from 0.4 to 0.7 W/m·K.
The κbp parameter, related to the bipolar electronic transport properties, can be estimated based on a single parabolic band model and the Boltzmann transport equation (Equation (4)):
κ b p = ( S p 2 σ p + S n 2 σ n S 2 σ )   T
where σp and σn are the electrical conductivities of the valence (p) and conduction (n) bands (VB and CB, respectively), while Sp and Sn are the Seebeck coefficients for the VB and CB, respectively.
The details of the κbp calculations are provided with the two-band model analysis in the Supplementary Materials, while the results of the calculations are shown in Figure 4c. The κbp value was gradually reduced from 0.36 to 0.23 W/m·K at 480 K, which represents a 36% decrease. The decrease in κbp is also mostly related to the increased concentration of electron carriers, which are the majority carriers. Therefore, the influence of the minority carriers is reduced. The inset of Figure 4c highlights a linear relationship between the κbp and σp values at 480 K [21]. The σp values estimated from the two-band model are provided in the Supplementary Materials and Table S1.
Then, the κlatt were deduced by subtracting the κelec and κbp values from the measured κtot, and are shown as symbols in Figure 4d. The κlatt (symbols in Figure 4d) was fitted to the theoretical κlatt (lines in Figure 4d) using the Debye-Callaway equation:
κ l a t t =   k B 2 π 2 ν ( k B T ħ ) 3 0 θ D / T τ t o t ( z ) z 4 e z ( e z 1 ) 2 d z
where τtot, θD, v, and ħ are the total phonon relaxation time, Debye temperature, phonon group velocity, and Planck constant divided by 2π, respectively, while z = ħω/kBT (ω = phonon frequency). Therefore, the determined τtot(z) values describe the theoretical κlatt, whereas τtot(z) can be estimated from the individual phonon relaxation times (τi) for scattering mechanisms, based on Matthiessen’s equation (Equation (6)):
τ total ( z ) 1 = i τ i ( z ) 1 =   τ U ( z ) 1 +   τ B ( z ) 1 +   τ P D ( z ) 1 .
For scattering by point defects, which is the dominant mechanism in the present Pd doping case, the phonon relaxation time can be described using the scattering parameter (Г) within τPD, as shown in Equations (7) and (8):
τ P D 1 = P   f ( 1 f )   ω 4 =   V ω 4 4 π v 3 Г
Г = f ( 1 f ) [ ( Δ M M ) 2 + 2 9 { ( G + 6.4 γ ) 1 + r 1 r } 2 ( Δ a a ) 2 ] .
In Equation (7), P and f are a fitting parameter and substituting fraction, respectively. In Equation (8), ΔM and Δa are the difference in mass and lattice constant between the two constituents of an alloy. The G and γ represent the ratio of the fractional change in the bulk modulus to the local bond length and the Grüneisen parameter, while r is the Poisson ratio. Further details of the calculation were not included here, because we found no differences in the κlatt values.
The theoretical κlatt is shown as solid lines in Figure 4d, along with the experimental κlatt (symbols). The experimental or theoretical κlatt values show rather small changes with the doping level, despite reaching a maximum at x = 0.02, implying that only minor additional scattering originated from the doped Pd. This is a peculiar result, as there is much evidence of additional point defect scattering upon substitutional doping. Due to the effect of the mass and lattice constant differences between two constituents of an alloy, described by Equation (8), we would expect a rather large additional contribution from phonon scattering, due to the large mass and ionic radius differences between Pd and Bi (MPd = 106.42 u, MBi = 208.98 u, rPd = 90 pm, rBi =117 pm). Despite the rather large ΔM and Δa values, we did not observe significant additional scattering due to the Pd doping. A possible explanation would be that intercalated Cu and Te/Se disorder already provide enough point defect scattering, so that Pd substitution would not contribute further in reducing κlatt. Scattering from Cu is known to be rather effective [12]. Consequently, the κtot value at 300 K showed a significant increase due to the increased κelec, whereas that at 480 K increased only slightly, together with the decrease in κbp, seen in Figure 4a.
Figure 5 shows the zT values of all samples. At low temperatures, the zT values were reduced, mainly due to the κtot increase. However, at higher temperatures (over 400 K), enhanced zT values were observed for intermediate Pd doping levels of x = 0.004 and 0.01. This is due to the enhanced power factors, along with the fact that κtot did not increase significantly despite the κelec increase. For instance, the zT at 480 K increased from 0.70 to 0.79 in the x = 0.01 case. However, no clear Pd doping-induced enhancement in zT was observed at doping levels higher than x = 0.01, due to the simultaneous increase in κelec and κtot, resulting from an excessive increase in electron carrier concentration. We found that moderate doping of Pd with levels of x = 0.004 to 0.01 in n-type Cu0.008Bi2Te2.7Se0.3 can be effective in enhancing the power factor. However, the Pd doping in Cu-doped n-type Bi2(Te,Se)3 did not further reduce κlatt despite the rather large ΔM and Δa values.

4. Conclusions

We studied the influence of Pd substitution in n-type Cu-doped Bi2Te2.7Se0.3, Cu0.008Bi2Te2.7Se0.3, by analyzing the electrical and thermal properties of a series of n-type Cu0.008PdxBi2-xTe2.7Se0.3 alloys (x = 0, 0.002, 0.004, 0.01, and 0.02) based on a single parabolic band and Debye-Callaway models. As the Pd doping increased, the electron carrier concentration and electrical conductivity increased simultaneously. The power factor was also enhanced, especially at higher temperatures. The bipolar conduction in the Pd-doped Cu0.008Bi2Te2.7Se0.3 samples was reduced; in particular, the bipolar thermal conductivity showed a significant decrease from 0.36 W/m·K in the undoped sample to 0.24 W/m·K in the x = 0.02 doped sample at 480 K. However, the analysis of the lattice thermal conductivity showed that substitutional Pd is not very effective in enhancing phonon scattering when interstitial Cu and Se/Te disorder are already present. Consequently, enhanced zT values at temperatures higher than 400 K were observed for the x = 0.004 and 0.01 doped samples.

Supplementary Materials

The following are available online at https://www.mdpi.com/1996-1944/12/24/4080/s1, Table S1 Band parameters of Pd-doped Cu0.008PdxBi2-xTe2.7Se0.3 samples (x = 0, 0.002, 0.004, 0.01, and 0.02) calculated using the two-band model.

Author Contributions

Conceptualization, W.H.S.; Formal analysis, Y.O. and Y.Y.; Investigation, H.-j.C., S.-s.C., and S.-w.H.; Methodology, K.L. and J.-H.L., S.-M.C.; Project administration, S.-i.K.; Supervision, W.H.S., K.H.L.; Writing—original draft, S.Y.K., H.-S.K. and S.-i.K.; Writing—review and editing, H.J.P. and K.H.L.

Funding

This research was funded by the 2018 Research Fund of the University of Seoul (20180511069).

Conflicts of Interest

The authors declare no conflict of interest” or declare any conflicts of interest.

References

  1. Bell, L.E. Cooling, heating, generating power, and recovering waste heat with thermoelectric systems. Science 2008, 321, 1457–1461. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Scherrer, H.; Scherrer, S. Thermoelectrics Handbook: Macro to Nano, 1st ed.; Rowe, D.M., Ed.; CRC Press: Boca Raton, FL, USA, 2006; pp. 1–27. [Google Scholar]
  3. Cui, J.; Xiu, W.; Xue, H. High thermoelectric properties of p-type pseudobinary (Cu4Te3)x–(Bi0.5Sb1.5Te3)1−x alloys prepared by spark plasma sintering. J. Appl. Phys. 2007, 101, 123713. [Google Scholar] [CrossRef]
  4. Mun, H.; Lee, K.H.; Kim, S.J.; Kim, J.Y.; Lee, J.H.; Lim, J.H.; Park, H.J.; Roh, J.W.; Kim, S.W. Fe-Doping Effect on Thermoelectric Properties of p-Type Bi0.48Sb1.52Te3. Materials 2015, 8, 959–965. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Kim, H.S.; Lee, K.H.; Yoo, J.Y.; Shin, W.H.; Roh, J.W.; Hwang, J.Y.; Kim, S.W.; Kim, S.I. Suppression of bipolar conduction via bandgap engineering for enhanced thermoelectric performance of p-type Bi0.4Sb1.6Te3 alloys. J. Alloy. Compd. 2018, 741, 869–874. [Google Scholar] [CrossRef]
  6. Lee, K.H.; Hwang, S.W.; Ryu, B.K.; Ahn, K.H.; Roh, J.W.; Yang, D.J.; Lee, S.M.; Kim, H.S.; Kim, S.I. Enhancement of the Thermoelectric Performance of Bi0.4Sb1.6Te3 Alloys by In and Ga Doping. J. Electron. Mater. 2013, 42, 1617–1621. [Google Scholar] [CrossRef]
  7. Lee, K.H.; Choi, S.M.; Roh, J.W.; Hwang, S.W.; Kim, S.I.; Shin, W.H.; Park, H.J.; Lee, J.H.; Kim, S.W.; Yang, D.J. Enhanced Thermoelectric Performance of p-Type Bi-Sb-Te Alloys by Codoping with Ga and Ag. J. Electron. Mater. 2015, 44, 1531–1535. [Google Scholar] [CrossRef]
  8. Zhai, R.S.; Wu, Y.H.; Zhu, T.J.; Zhao, X.B. Thermoelectric performance of p-type zone-melted Se-doped Bi0.5Sb1.5Te3 alloys. Rare Met. 2018, 37, 308–315. [Google Scholar] [CrossRef]
  9. Kim, H.; Lee, J.K.; Park, S.D.; Ryu, B.K.; Lee, J.E.; Kim, B.S.; Min, B.K.; Joo, S.J.; Lee, H.W.; Cho, Y.R. Enhanced thermoelectric properties and development of nanotwins in Na-doped Bi0.5Sb1.5Te3 alloy. Electron. Mater. Lett. 2016, 12, 290–295. [Google Scholar] [CrossRef]
  10. Kim, I.H.; Choi, S.M.; Seo, W.S.; Cheong, D.I. Thermoelectric properties of Cu-dispersed Bi0.5Sb1.5Te3. Nanoscale Res. Lett. 2012, 7, 2. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Kim, H.S.; Lee, K.H.; Yoo, J.Y.; Youn, J.H.; Roh, J.W.; Kim, S.I.; Kim, S.W. Effect of substitutional Pb doping on bipolar and lattice thermal conductivity in p-type Bi0.48Sb1.52Te3. Materials 2017, 10, 763. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Cho, H.J.; Shin, W.H.; Choo, S.S.; Kim, J.I.; Yoo, J.Y.; Kim, S.I. Synergistic Influence of Cu Intercalation on Electronic and Thermal Properties of n-Type CuxBi2Te2.7Se0.3 Polycrystalline Alloys. J. Electron. Mater. 2019, 48, 1951–1957. [Google Scholar] [CrossRef]
  13. Tan, X.; Liu, G.; Xu, J.; Tan, X.; Shao, H.; Hu, H.; Jiang, H.; Lu, Y.; Jiang, J. Thermoelectric properties of In-Hg co-doping in SnTe: Energy band engineering. J. Mater. 2018, 4, 62–67. [Google Scholar] [CrossRef]
  14. Roychowdhury, S.; Biswas, K. Effect of In and Cd co-doping on the thermoelectric properties of Sn1−xPbxTe. Mater. Res. Express 2019, 6, 104010. [Google Scholar] [CrossRef] [Green Version]
  15. Cohen, I.; Kaller, M.; Komisarchik, G.; Fuks, D.; Gelbstein, Y. Enhancement of the thermoelectric properties of n-type PbTe by Na and Cl co-doping. J. Mater. Chem. C 2015, 3, 9559–9564. [Google Scholar] [CrossRef]
  16. Kihoi, S.K.; Kim, H.; Jeong, H.; Kim, H.; Ryu, J.; Yi, S.; Lee, H.S. Thermoelectric properties of Mn, Bi, and Sb co-doped SnTe with a low lattice thermal conductivity. Alloys Compd. 2019, 806, 361–369. [Google Scholar] [CrossRef]
  17. Kim, H.-S.; Choo, S.-S.; Cho, H.-J.; Kim, S.-I. Beneficial Influence of Co-Doping on Thermoelectric Efficiency with Respect to Electronic and Thermal Transport Properties. Phys. Status Sol. A 2019, 216, 1900039. [Google Scholar] [CrossRef]
  18. May, A.F.; Snyder, G.J. Materials, Preparation, and Characterization in Thermoelectric; Rowe, D.M., Ed.; CRC Press: Boca Raton, FL, USA, 2012; pp. 1–11. [Google Scholar]
  19. Callaway, J. Model for Lattice Thermal Conductivity at Low Temperatures. Phys. Rev. 1959, 113, 1046. [Google Scholar] [CrossRef]
  20. Kim, H.S.; Gibbs, Z.M.; Tang, Y.; Wang, H.; Snyder, G.J. Characterization of Lorenz number with Seebeck coefficient measurement. APL Mater. 2015, 3, 041406. [Google Scholar] [CrossRef] [Green Version]
  21. Wang, S.; Yang, J.; Toll, T.; Yang, J.; Zhang, W.; Tang, X. Conductivity-limiting bipolar thermal conductivity in semiconductors. Sci. Rep. 2015, 5, 10136. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. (a) X-ray diffraction patterns and (b) calculated lattice parameters a and c of Cu0.008PdxBi2-xTe2.7Se0.3 (x = 0, 0.002, 0.004, 0.01, and 0.02).
Figure 1. (a) X-ray diffraction patterns and (b) calculated lattice parameters a and c of Cu0.008PdxBi2-xTe2.7Se0.3 (x = 0, 0.002, 0.004, 0.01, and 0.02).
Materials 12 04080 g001
Figure 2. (a) σ, (b) S, and (c) power factor of Cu0.008PdxBi2-xTe2.7Se0.3 (x = 0, 0.002, 0.004, 0.01, and 0.02).
Figure 2. (a) σ, (b) S, and (c) power factor of Cu0.008PdxBi2-xTe2.7Se0.3 (x = 0, 0.002, 0.004, 0.01, and 0.02).
Materials 12 04080 g002
Figure 3. (a) Measured carrier concentrations and mobilities and (b) Pisarenko plot.
Figure 3. (a) Measured carrier concentrations and mobilities and (b) Pisarenko plot.
Materials 12 04080 g003
Figure 4. (a) κtot (κtot = κelec + κlatt + κbp), (b) κelec, (c) κbp, and (d) κlatt. The inset of (c) shows the linear relationship between κbp and σp.
Figure 4. (a) κtot (κtot = κelec + κlatt + κbp), (b) κelec, (c) κbp, and (d) κlatt. The inset of (c) shows the linear relationship between κbp and σp.
Materials 12 04080 g004
Figure 5. zT values of Cu0.008PdxBi2-xTe2.7Se0.3 (x = 0, 0.002, 0.004, 0.01, and 0.02).
Figure 5. zT values of Cu0.008PdxBi2-xTe2.7Se0.3 (x = 0, 0.002, 0.004, 0.01, and 0.02).
Materials 12 04080 g005

Share and Cite

MDPI and ACS Style

Kim, S.Y.; Kim, H.-S.; Lee, K.H.; Cho, H.-j.; Choo, S.-s.; Hong, S.-w.; Oh, Y.; Yang, Y.; Lee, K.; Lim, J.-H.; et al. Influence of Pd Doping on Electrical and Thermal Properties of n-Type Cu0.008Bi2Te2.7Se0.3 Alloys. Materials 2019, 12, 4080. https://doi.org/10.3390/ma12244080

AMA Style

Kim SY, Kim H-S, Lee KH, Cho H-j, Choo S-s, Hong S-w, Oh Y, Yang Y, Lee K, Lim J-H, et al. Influence of Pd Doping on Electrical and Thermal Properties of n-Type Cu0.008Bi2Te2.7Se0.3 Alloys. Materials. 2019; 12(24):4080. https://doi.org/10.3390/ma12244080

Chicago/Turabian Style

Kim, Se Yun, Hyun-Sik Kim, Kyu Hyoung Lee, Hyun-jun Cho, Sung-sil Choo, Seok-won Hong, Yeseong Oh, Yerim Yang, Kimoon Lee, Jae-Hong Lim, and et al. 2019. "Influence of Pd Doping on Electrical and Thermal Properties of n-Type Cu0.008Bi2Te2.7Se0.3 Alloys" Materials 12, no. 24: 4080. https://doi.org/10.3390/ma12244080

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop