Skip to main content
ACS AuthorChoice logoLink to ACS AuthorChoice
. 2023 Jan 9;35(2):755–763. doi: 10.1021/acs.chemmater.2c03542

Fine Tuning of Defects Enables High Carrier Mobility and Enhanced Thermoelectric Performance of n-Type PbTe

Siqi Wang , Cheng Chang ‡,*, Shulin Bai §, Bingchao Qin , Yingcai Zhu , Shaoping Zhan , Junqing Zheng , Shuwei Tang §, Li-Dong Zhao †,*
PMCID: PMC9878722  PMID: 36711054

Abstract

graphic file with name cm2c03542_0010.jpg

High carrier mobility is critical to improving thermoelectric performance over a broad temperature range. However, traditional doping inevitably deteriorates carrier mobility. Herein, we develop a strategy for fine tuning of defects to improve carrier mobility. To begin, n-type PbTe is created by compensating for the intrinsic Pb vacancy in bare PbTe. Excess Pb2+ reduces vacancy scattering, resulting in a high carrier mobility of ∼3400 cm2 V–1 s–1. Then, excess Ag is introduced to compensate for the remaining intrinsic Pb vacancies. We find that excess Ag exhibits a dynamic doping process with increasing temperatures, increasing both the carrier concentration and carrier mobility throughout a wide temperature range; specifically, an ultrahigh carrier mobility ∼7300 cm2 V–1 s–1 is obtained for Pb1.01Te + 0.002Ag at 300 K. Moreover, the dynamic doping-induced high carrier concentration suppresses the bipolar thermal conductivity at high temperatures. The final step is using iodine to optimize the carrier concentration to ∼1019 cm–3. Ultimately, a maximum ZT value of ∼1.5 and a large average ZTave value of ∼1.0 at 300–773 K are obtained for Pb1.01Te0.998I0.002 + 0.002Ag. These findings demonstrate that fine tuning of defects with <0.5% impurities can remarkably enhance carrier mobility and improve thermoelectric performance.

Introduction

Thermoelectric materials can effectively utilize low-grade heat sources to generate electricity, such as industrial waste heat, automobile exhaust, solar energy, geothermal heat, etc.14 To date, the priority of thermoelectric research is to improve the conversion efficiency for more widespread applications. The conversion efficiency depends on the dimensionless figure of merit ZT, ZT = S2σT/κ, where S is the Seebeck coefficient, σ is the electrical conductivity, T is the absolute temperature in kelvin, and κ is the total thermal conductivity (κ = κlat + κele + κbi, where κlat is the lattice thermal conductivity, κele is the electronic thermal conductivity, κbi is the bipolar thermal conductivity). In general, thermoelectric enhancement relies heavily on a large Seebeck coefficient, high electrical conductivity, and low thermal conductivity. Nevertheless, the above three thermoelectric parameters are governed primarily by the carrier concentration and are not independently controllable, making them difficult to be optimized simultaneously.59 Researchers have conducted in-depth and fruitful research in upgrading thermoelectric performance via combined optimization of electrical and thermal transport. For electrical properties, optimizing carrier concentration,1012 increasing carrier mobility,13,14 and modulating band structure1519 are generally applied. For thermal properties, the traditional method is declining the lattice thermal conductivity via hierarchical structures, involving atomic-scale defects,2022 nanoscale defects,2325 and microscale defects.26,27

Among a variety of thermoelectric materials, PbTe is a representative thermoelectric material servicing at medium temperature. Numerous research studies have shown that p-type PbTe materials have achieved extraordinary thermoelectric properties. Many prominent p-type PbTe systems have realized ZT > 2.0, such as Pb0.98Na0.02Te,28 Pb0.98Na0.02Te–8%SrTe,29 Pb0.98Na0.02Te–6%MgTe,30 Pb0.95Na0.05Te–0.5%AgInSe2,31 and Pb0.075K0.025Te0.7S0.3.32 In contrast, their counterpart, n-type PbTe, shows lower ZT values because of the large energy offset (∼0.45 eV) between the conduction bands, so further improvement of n-type PbTe is needed to match p-type PbTe. With time, we discovered several optimization methods that could bring about considerable intensification in the thermoelectric performance of n-type PbTe, such as doping and band engineering.33 However, even though traditional doping undoubtedly plays a significant role in increasing the carrier concentration, dopants inevitably reduce carrier mobility.3436 Band engineering, as another attractive and feasible method, facilitates electrical performance by boosting the effective mass, while it is also detrimental to the carrier mobility.37,38 In general, these methods are capable of enhancing the maximum ZT, whereas a major drawback is their tendency to decrease carrier mobility, which limits the ZTave.39 To ensure good thermoelectric properties over a wide temperature range, comparable n-type PbTe systems with high carrier mobility must be developed.

In this work, fine tuning of defects in PbTe is successively carried out to improve the carrier mobility. First, a small amount of excess Pb (0.004–0.012) is introduced into PbTe. As Pb enters the intrinsic Pb vacancies, n-type PbTe with relatively low carrier concentration is obtained. The dwindling vacancy defects improve carrier mobility, and the maximum carrier mobility reaches ∼3400 cm2 V–1 s–1 at 300 K in Pb1.01Te. Benefiting from the high carrier mobility, a maximum power factor of ∼31.3 μW cm–1 K–2 at 300 K is achieved in Pb1.01Te. Second, excess Ag is subsequently introduced to fill the remaining Pb vacancies. Results reveal that the comparatively low room-temperature carrier concentration of ∼5.5 × 1017 cm–3 leads to a high carrier mobility of ∼7300 cm2 V–1 s–1, resulting in a further increased power factor of ∼39.3 μW cm–1 K–2 at 300 K for Pb1.01Te + 0.002Ag. In particular, Ag atoms in the Pb1.01Te system undergo a dynamic doping process with increasing temperatures, first compensating the Pb vacancies and then entering more interstitial positions due to the temperature-dependent solubility of Ag in the PbTe matrix,40 which improves the carrier concentration from ∼5.5 × 1017 cm–3 at 300 K to ∼8.5 × 1018 cm–3 at 873 K for Pb1.01Te + 0.002Ag. Meanwhile, this unique dynamic doping behavior can effectively decouple the carrier concentration and carrier mobility across the entire temperature range and suppress bipolar diffusion at high temperatures. Last, 0.2% of I element (0.002I) is adopted to optimize the carrier concentration while maintaining superior carrier mobility than other n-type PbTe systems with similar carrier concentrations. Meanwhile, the minimum lattice thermal conductivity can reach as low as ∼0.5 W m–1 K–1 for Pb1.01Te0.998I0.002 + 0.002Ag. As a result, a maximum ZT of ∼1.5 at 773 K and a high average ZTave of ∼1.0 at 300–773 K are obtained for Pb1.01Te0.998I0.002 + 0.002Ag.

Experimental Section

All raw materials used in the experiment were simple substances with more than 99.99% purity, including Pb, Te, Ag, and I. First, they were placed in quartz tubes according to the stoichiometric ratio. The quartz tubes were sealed below 10–4 Pa and put into a furnace (slowly heating from room temperature to 1323 K for 24 h, holding for 10 h, and finally cooling to room temperature). After the temperature program finished, some silver ingots were obtained, which were then ground into powder for proceeding spark plasma sintering (SPS-211Lx, Dr. Sinter). Finally, the sintered cylindrical, small pieces were cut into rectangles of 10 mm × 3 mm × 3 mm and slices of 8 mm × 8 mm × 1.5 mm to perform electrical and thermal performance tests. The electrical performance was measured by Cryoall CTA (electrical conductivity, σ and Seebeck coefficient, S) and Lake Shore 8400 Series (carrier density, n) instruments. The thermal performance was tested by a Netzsch LFA 457 (thermal diffusivity, D). The thermal conductivity, κ, was estimated by κ = D × Cp × ρ, where Cp is the specific heat capacity based on Debye’s law and ρ is the sample density calculated by the ratio of mass to volume. The optical band gap was measured by a Fourier transform infrared spectrometer, IRAffinity-1S, based on the infrared diffuse reflection method.

Results and Discussion

In this study, the method of fine tuning of defects is used in the PbTe system to achieve superior electrical characteristics while maintaining a relatively low thermal conductivity. Supporting Information to create n-type PbTe, a modest quantity of Pb was first introduced, filling the intrinsic Pb vacancies, which increases carrier mobility and lowers thermal conductivity. Second, excess Ag is adopted for Pb1.01Te to fill the remaining Pb vacancies and also occupy interstitial sites, allowing the Pb1.01Te + xAg system to achieve higher carrier concentration, ultrahigh carrier mobility, and lower thermal conductivity simultaneously across the entire temperature range. Third, the Iodine element is adopted to enhance the carrier concentration, further improving the power factor and restraining the lattice thermal conductivity in the whole temperature range.

Thermoelectric Transport Performance of Pb1+xTe

Figure S1a shows the powder X-ray diffraction (XRD) patterns of Pb1+xTe (x = 0, 0.004, 0.006, 0.008, 0.01, 0.012), and the corresponding lattice parameters are given in Figure S1b. All Pb1+xTe samples are NaCl-type cubic structures without extra diffraction peaks detected, and the lattice parameters increase slightly with excess Pb.

As can be seen in Figure 1a, the electrical conductivities decrease first and then increase with increasing Pb amount. Correspondingly, Pb1+xTe undergoes a p–n transition at low temperatures, as shown in Figure 1b. Specifically, the Seebeck coefficient of bared PbTe is positive at low temperatures; however, when x ≥ 0.008, the Seebeck coefficient is completely negative, resulting in n-type PbTe throughout the measuring temperature range. The combined behavior of electrical conductivity and Seebeck coefficient indicates that the hole concentration in PbTe is suppressed by excess Pb. In Figure 1c, benefitting from the improved electrical conductivity and Seebeck coefficient, the power factor is enhanced significantly, and the maximum value, i.e., room-temperature value, increases notably from ∼18.0 μW cm–1 K–2 for PbTe to ∼31.3 μW cm–1 K–2 for Pb1.01Te. In Figure 1d, the total thermal conductivity of PbTe reduces slightly with excess Pb, and the room-temperature value decreases from ∼2.1 to ∼2.0 Wm–1 K–1 caused by diminished lattice thermal conductivity, as shown in Figure 1e. Pb1+xTe shows higher electronic thermal conductivity, as shown in Figure S2b, because excess Pb fills the intrinsic Pb vacancies, thereby creating weaker carrier scattering. Ultimately, an average ZT value of ∼0.5 for Pb1.01Te is obtained throughout the whole measuring temperature range, as shown in Figure 1f.

Figure 1.

Figure 1

Thermoelectric transport performance of Pb1+xTe (x = 0, 0.004, 0.006, 0.008, 0.01, 0.012): (a) electrical conductivity, (b) Seebeck coefficient, (c) power factor, (d) total thermal conductivity, (e) lattice thermal conductivity, and (f) ZT value.

To understand the electrical transport evolution of Pb1+xTe, we carried out the Hall measurement. In Figure 2a, the carrier concentration of ∼1017 cm–3 in the n-type Pb1+xTe system is relatively low, not as outstanding as the common doping approaches, such as PbTe–Sb41 (∼5.2 × 1019 cm–3), PbTe–I42 (∼1.8 × 1019 cm–3), PbTe–Bi43 (∼2.0 × 1019 cm–3), etc. With the increasing Pb fraction, the carrier concentration is gradually elevated, indicating that more and more Pb enters Pb vacancies. Based on the inverse relationship between carrier concentration and carrier mobility, relatively low carrier concentration brings about high carrier mobility since reduced Pb vacancies alleviate the charge carrier scattering, and the maximum carrier mobility can reach ∼3400 cm2 V–1 s–1 for Pb1.01Te. Figure 2b shows the atomic schematic diagram of the Pb1+xTe system, graphically depicting the phenomenon that excess Pb atoms occupy the intrinsic Pb vacancies.

Figure 2.

Figure 2

(a) Room-temperature carrier concentration and carrier mobility as a function of Pb and (b) atomic schematic diagram of Pb1+xTe.

Electrical Transport Performance of Pb1.01Te + xAg

The powder X-ray diffraction (XRD) measurements of Pb1.01Te + xAg (x = 0–0.005) samples are performed, as shown in Figure S4a. All samples are NaCl cubic structures, and no peaks of impurities are found. Likewise, the small amount of excess Ag results in a slight expansion of the lattice parameters, as shown in Figure S4b.

Figure 3 shows that excess Ag significantly improves the electrical transport performance of the Pb1.01Te + xAg system. Figure 3a shows a clear increase in electrical conductivity throughout the whole temperature range. In particular, the electrical conductivity increases anomaly above 500 K. In Figure 3b, the negative Seebeck coefficients prove that Pb1.01Te + xAg is electron-dominated n-type material. The introduction of Ag reduces the Seebeck coefficient, and the maximum absolute value decreases from ∼385.9 μV K–1 for Pb1.01Te to ∼249.0 μV K–1 for Pb1.01Te + 0.002Ag. The Pisarenko curve in Figure S5 reveals that the effective mass of the Pb1.01Te + xAg system remains approximately 0.22m0, indicating that the depressed Seebeck coefficient of Pb1.01Te + xAg is attributed to the elevated carrier concentration rather than the reduced effective mass. As a result, significantly incremental electrical conductivity and abated Seebeck coefficient further enhance the power factor compared to Pb1.01Te. In particular, a superior room-temperature power factor of ∼39.3 μW cm–1 K–2 is acquired for Pb1.01Te + 0.002Ag, as shown in Figure 3c. Also, the power factor above 500 K is twice higher as that of Pb1.01Te. Figure 3d shows the carrier concentration and carrier mobility of Pb1.01Te + xAg at room temperature. Carrier concentration increases as the Ag fraction rises; however, carrier mobility is not impaired. Particularly, the maximum carrier mobility is increased by more than twice from ∼3400 cm2 V–1 s–1 for Pb1.01Te to ∼7300 cm2 V–1 s–1 for Pb1.01Te + 0.002Ag.

Figure 3.

Figure 3

Electrical transport performance of Pb1.01Te + xAg (x = 0–0.005): (a) electrical conductivity, (b) Seebeck coefficient, (c) power factor, and (d) room-temperature carrier concentration and carrier mobility as a function of Ag.

To get an insight into the remarkable electrical transport properties of Pb1.01Te + xAg, we measured the temperature-dependent carrier concentration, as shown in Figure 4a,b. The carrier concentrations of bare PbTe and Pb1.01Te increase with increasing temperatures, as shown in Figure S6, which is derived from the low carrier concentration and thus pronounced bipolar effect at high temperatures.44 However, Pb1.01Te + xAg systems still exhibit increasing carrier concentration throughout the temperature range even though their bipolar effect is suppressed (the suppressed bipolar effect is evidenced by the low bipolar thermal conductivity, which is discussed in detail in the next section). We attribute the abnormal temperature-dependent behavior of carrier concentration to Ag-induced dynamic doping.40 The dynamic doping process in Pb1.01Te + xAg is initiated by the multiple occupancies of Ag atoms with increasing temperatures, and Figure 4c displays the phenomena in detail. At 300–500 K, Ag atoms compensate for intrinsic Pb vacancies and interstitials, enhancing carrier mobility, as shown in Figure 4b. Above 500 K, more Ag atoms occupy interstitial positions to provide excess electrons for the system, resulting in a higher carrier concentration than the Pb1.01Te sample. It is noteworthy that the carrier mobilities of Pb1.01Te + Ag at high temperatures are also higher than those of Pb1.01Te, which can be explained by the suppressed bipolar effect. As a result, the carrier mobility of Pb1.01Te + Ag is greater than that of Pb1.01Te in the entire temperature range.

Figure 4.

Figure 4

Dynamic doping process in Pb1.01Te + xAg: (a) carrier concentration and (b) carrier mobility as a function of temperature, (c) atomic schematic diagram, and (d) calculated defect formation energies of PbTe.

To further reveal the role of Ag in PbTe, we utilized first principles to calculate various defects and their formation energy in the system, such as vacancies, interstitials, antisites, and Ag-filled intrinsic vacancies. Figure 4d depicts the calculated formation energy of related defects. Based on the existence of Pb vacancies, the formation energy of Pb vacancies in the Pb-rich system is greater than that in the Te-rich system, implying that Pb vacancies are more difficult to form in the Pb-rich condition. For Pb1.01Te (Pb-rich system), the formation energy of AgPb (Ag-occupied Pb vacancies) is lower than that of VPb (Pb vacancies), indicating that the introduction of Ag reduces the Pb vacancies, thereby lowering the content of point defects and improving the carrier mobility. The formation energy of Agi (Ag interstitials) is also very low in both Pb-rich and Te-rich systems, demonstrating that Ag not only fills Pb vacancies but also occupies the interstitials in PbTe. The calculated results of defect formation energy well support the experimental results and verify the role of the Ag element in improving the electrical transport performance of the Pb1.01Te system.

Thermal Transport Performance of Pb1.01Te + xAg

Figure 5 displays the temperature-dependent thermal properties of Pb1.01Te + xAg. In Figure 5a, the total thermal conductivity decreases slightly. The electronic thermal conductivity is calculated from equation κele = LσT (L is the Lorenz number, as shown in Figure S7a, determined by the single parabolic band (SPB) model). Because of the greater electrical conductivity, the Pb1.01Te + xAg system displays much higher electronic thermal conductivity than Pb1.01Te, as shown in Figure 5b. Figure 5c presents the lattice thermal conductivity κlat calculated by κlat = κtot – κele as a function of temperature. In the high-temperature region, all of the Pb1.01Te + xAg samples exhibit lower κlat than Pb1.01Te since the elevated electronic thermal conductivity of Pb1.01Te + xAg effectively suppresses the bipolar diffusion. Excess Ag depresses the minimum lattice thermal conductivity from ∼1.2 Wm–1 K–1 for Pb1.01Te to ∼0.7 Wm–1 K–1 for Pb1.01Te + 0.005Ag. The solid black line in Figure 5c is the theoretical lattice thermal conductivity of PbTe without bipolar thermal conductivity. The theoretical lattice thermal conductivity as a function of the temperature follows the relationship30

graphic file with name cm2c03542_m001.jpg

where kB is the Boltzmann constant, is the Planck constant, M is the average atomic mass, V is the average atomic volume, θD is the Debye temperature, and γ is the Gruneisen constant. The κtot – κele values of Pb1.01Te + xAg samples coincide with the calculated values, indicating that the bipolar diffusion is suppressed in Pb1.01Te + xAg. Meanwhile, the enlarged band gap is responsible for restrained bipolar diffusion due to the following formula45,46

graphic file with name cm2c03542_m002.jpg

where A is the constant and Eg is the band gap. Ag enlarges the band gap in Figure 5d, thereby reducing the bipolar thermal conductivity. The (κtot – κele) – 1000 T–1 relationship is plotted in Figure 5e to examine the contribution of bipolar thermal conductivity, revealing that the true lattice thermal conductivity should be a straight line, while the κtot – κele values in the high-temperature region are higher than the predicted values, indicating that bipolar diffusion occurs in this region. According to Figure 5f, bipolar thermal conductivity is the difference between the solid line and the dotted line and decreases as the amount of Ag increases.

Figure 5.

Figure 5

Thermal transport performance of Pb1.01Te + xAg (x = 0–0.005): (a) total thermal conductivity, (b) electronic thermal conductivity, (c) κtot – κele as a function of T, (d) room-temperature band gap, (e) κtot – κele as a function of 1000/T, and (f) bipolar thermal conductivity.

ZT Values of Pb1.01Te + xAg

Comparing room-temperature carrier mobility of the Pb1.01Te + xAg system with those of the other n-type PbTe materials, we found that this system exhibits superior carrier mobility of ∼7300 cm2 V–1 s–1 at low carrier concentrations (∼1017 cm–3), as shown in Figure 6a, which is competitive in high-performance n-type PbTe-based materials. Finally, as a result of ultrahigh carrier mobility and suppressed bipolar thermal conductivity, room-temperature ZT, as shown in Figure 6b, is further enhanced from ∼0.5 for Pb1.01Te to ∼0.6 for Pb1.01Te + 0.002Ag, and high-temperature ZT is significantly increased from ∼0.5 for Pb1.01Te to ∼1.3 for Pb1.01Te + 0.004Ag.

Figure 6.

Figure 6

(a) Comparison of carrier mobility as a function of the carrier concentration of n-type PbTe and (b) ZT values of Pb1.01Te + xAg.

Thermoelectric Transport Performance of Pb1.01Te1–xIx + 0.002Ag

Based on Pb1.01Te + 0.002Ag with the maximum average ZT, I doping is employed to increase the carrier concentration, thereby improving the electrical performance. Figure S8 shows the phase identification of Pb1.01Te1–xIx + 0.002Ag (x = 0–0.003). No new peak is found in Figure S8a, and the samples still denote the NaCl structure. The lattice parameter decreases slightly in Pb1.01Te1–xIx + 0.002Ag, as shown in Figure S8b, because smaller I (∼2.06 Å) substitutes for larger Te2– (∼2.11 Å). Figure 7 depicts the thermoelectric performance of Pb1.01Te1–xIx + 0.002Ag (x = 0–0.003). I doping can effectively improve the electrical conductivity, as shown in Figure 7a. The room-temperature electrical conductivity can be magnified from ∼633.5 S cm–1 for Pb1.01Te + 0.002Ag to ∼3267.8 S cm–1 for Pb1.01Te0.998I0.002 + 0.002Ag. The Seebeck coefficients of all samples in Figure 7b are negative, showing n-type semiconductor characteristics. The absolute value of the Seebeck coefficient of Pb1.01Te1–xIx + 0.002Ag decreases with increasing doping amount of the I element since the more I content, the higher the carrier concentration, as shown in Figure S9. The high electrical conductivity and low Seebeck coefficient of the Pb1.01Te1–xIx + 0.002Ag system indicate that the I element possesses extremely high doping efficiency in PbTe. The carrier concentration can be efficiently increased by doping with as little as 0.002I. Figure 7c shows that the power factor increases across the entire temperature range, with the maximum power factor increasing from ∼39.3 μW cm–1 K–2 for Pb1.01Te + 0.002Ag to ∼48.5 μW cm–1 K–2 for Pb1.01Te0.998I0.002 + 0.002Ag. The average power factor exceeds ∼30.0 μW cm–1 K–2 for Pb1.01Te0.998I0.002 + 0.002Ag, which is higher than those of other n-type PbTe systems (PbTe–S–I,47 PbTe–Sb2Te3–Sb–Cu2Te,23 AgnPb100InTe100+2n,48 PbTe + Cu,49 PbTe–GeTe,50 PbTe–MnTe,51 PbTe–Ga52), as shown in Figure 7d. In Figure 7e, the total thermal conductivity of Pb1.01Te1–xIx + 0.002Ag increases with increasing I content, primarily owing to an enhancement in electronic thermal conductivity. The lattice thermal conductivity decreases with increasing I content, as shown in Figure 7f, due to the enhanced phonon scattering. The minimum lattice thermal conductivity is reduced to ∼0.5 Wm–1 K–1 for Pb1.01Te0.998I0.002 + 0.002Ag.

Figure 7.

Figure 7

Thermoelectric transport performance of Pb1.01Te1–xIx + 0.002Ag (x = 0–0.003): (a) electrical conductivity, (b) Seebeck coefficient, (c) power factor, (d) average power factor, (e) total thermal conductivity, and (f) lattice thermal conductivity.

To comprehensively evaluate the thermoelectric performance of the Pb1.01Te1–xIx + 0.002Ag system, the relationship between carrier concentration and the ratio of carrier mobility to lattice thermal conductivity (μ/κlat) is plotted, as shown in Figure 8a. Compared with other high-performance n-type PbTe systems (PbTe–S–I,47 PbTe–Sb2Te3–Sb–Cu2Te,23 AgnPb100InTe100+2n,48 PbTe + Cu,49 PbTe–GeTe,50 PbTe–MnTe,51 PbTe–Ga52), the Pb1.01Te1–xIx + 0.002Ag system possesses higher μ/κlat values at lower carrier concentrations, indicating that a trace of Pb atoms (Pb-occupied Pb vacancies) and Ag atoms (Ag-occupied Pb vacancies) can effectively regulate the intrinsic defects in PbTe and reduce the scattering for charge carriers, protecting the charge carrier transport while scattering phonon. Accordingly, owing to a better balance between electrons and phonons, the ZT value increases in the whole temperature range, and the maximum ZT value is enhanced from ∼1.2 for Pb1.01Te + 0.002Ag to ∼1.5 for Pb1.01Te0.998I0.002 + 0.002Ag, as shown in Figure 8b. Figure 8c,d shows the comparison of ZT values and average ZTave values with various n-type PbTe materials, respectively. The large ZT values of Pb1.01Te0.998I0.002 + 0.002Ag in the whole temperature zone contribute to the excellent average ZTave value, reaching >1.0 at 300–773 K.

Figure 8.

Figure 8

Comparisons of thermoelectric parameters of Pb1.01Te1–xIx + 0.002Ag with those of other systems: (a) diagram of μ/κlat and (b) ZT values of Pb1.01Te1–xIx + 0.002Ag. Comparisons of (c) ZT values and (d) average ZT values of Pb1.01Te0.998I0.002 + 0.002Ag with those of other PbTe systems.

Conclusions

This work has provided a deeper insight into designing fine-tuned defects to improve the carrier mobility of n-type PbTe. In particular, by introducing a small amount of Pb, the reduced Pb vacancy improves the carrier mobility to ∼3400 cm2 V–1 s–1 at 300 K for Pb1.01Te. Then, the room-temperature carrier mobility can reach as high as ∼7300 cm2 V–1 s–1 for Pb1.01Te + 0.002Ag due to Ag-induced dynamic doping. Finally, Iodine doping dramatically increases the carrier concentration while maintaining superior carrier mobility compared to other PbTe systems with only traditional dopants. The combination of high carrier mobility and the suppressed bipolar effect enables the significant enhancement of thermoelectric performance with a high ZT ∼1.5 at 773 K and average ZTave ∼1.0 at 300–773 K for n-type Pb1.01Te0.998I0.002 + 0.002Ag. These findings provide insights into balancing electron and phonon transport via fine tuning of defects, which could be a promising aspect of co-optimizing thermal and electrical performance of materials with instinct defects.

Acknowledgments

This work was supported by the National Key Research and Development Program of China (2018YFA0702100), the Basic Science Center Project of the National Natural Science Foundation of China (51788104), the National Natural Science Foundation of China (51571007 and 51772012), the Beijing Natural Science Foundation (JQ18004), the 111 Project (B17002), and the National Science Fund for Distinguished Young Scholars (51925101). C.C. acknowledges funding from the FWF “Lise Meitner Fellowship” (grant agreement M2889-N).

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.chemmater.2c03542.

  • Density functional theory (DFT) calculations, formula of the average PFave value, phase identification of Pb1+xTe (x = 0, 0.004, 0.006, 0.008, 0.01, 0.0012) (Figure S1), thermoelectric transport properties as a function of temperature for Pb1+xTe (x = 0, 0.004, 0.006, 0.008, 0.01, 0.0012) (Figure S2), sample density of Pb1+xTe (x = 0, 0.004, 0.006, 0.008, 0.01, 0.0012) (Figure S3); phase identification of Pb1.01Te + xAg (x = 0–0.005) (Figure S4), room-temperature Pisarenko plots of Pb1.01Te + xAg (Figure S5), carrier concentration of PbTe and Pb1.01Te as a function of temperature (Figure S6), thermoelectric transport properties as a function of temperature for Pb1.01Te + xAg (x = 0–0.005) (Figure S7), phase identification of Pb1.01Te1–xIx + 0.002Ag (x = 0–0.003) (Figure S8), room-temperature carrier density and carrier mobility as a function of Ag doping content (Figure S9), thermoelectric transport properties as a function of temperature for Pb1.01Te1–xIx + 0.002Ag (x = 0–0.003) (Figure S10), and sample density of Pb1.01Te1–xIx + 0.002Ag (x = 0–0.003) (Figure S11) (PDF)

Author Contributions

The manuscript was written through the contribution of all authors. All authors have given approval for the final version of the manuscript. Siqi Wang, research design, data collection and analysis, and draft writing and editing; Shulin Bai and Shuwei Tang, density functional theory (DFT) calculations; Bingchao Qin, Yingcai Zhu, Shaoping Zhan, and Junqing Zheng, data analysis; Cheng Chang and Li-Dong Zhao, research design, data analysis, draft writing and editing, and funding acquisition.

The National Key Research and Development Program of China (2018YFA0702100), the Basic Science Center Project of the National Natural Science Foundation of China (51788104), the National Natural Science Foundation of China (51571007 and 51772012), the Beijing Natural Science Foundation (JQ18004), the 111 Project (B17002), the National Science Fund for Distinguished Young Scholars (51925101), and the FWF “Lise Meitner Fellowship” (grant agreement M2889-N). Open Access is funded by the Austrian Science Fund (FWF).

The authors declare no competing financial interest.

Supplementary Material

cm2c03542_si_001.pdf (367.6KB, pdf)

References

  1. Su L. Z.; Wang D. Y.; Wang S. N.; Qin B. C.; Wang Y. P.; Qin Y. X.; Jin Y.; Chang C.; Zhao L. D. High Thermoelectric Performance Realized through Manipulating Layered Phonon-Electron Decoupling. Science 2022, 375, 1385–1389. 10.1126/science.abn8997. [DOI] [PubMed] [Google Scholar]
  2. Hu C. L.; Xia K. Y.; Fu C. G.; Zhao X. B.; Zhu T. J. Carrier Grain Boundary Scattering in Thermoelectric Materials. Energy Environ. Sci. 2022, 15, 1406–1422. 10.1039/D1EE03802H. [DOI] [Google Scholar]
  3. Bu Z. L.; Zhang X. Y.; Hu Y. X.; Chen Z. W.; Lin S. Q.; Li W.; Xiao C.; Pei Y. Z.; Record A. Thermoelectric Efficiency in Tellurium-Free Modules for Low-Grade Waste Heat Recovery. Nat. Commun. 2022, 13, 237 10.1038/s41467-021-27916-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Chang C.; Kanatzidis M. G. High-Entropy Thermoelectric Materials Emerging. Mater. Lab 2022, 1, 220048 10.54227/mlab.20220048. [DOI] [Google Scholar]
  5. Shi X. L.; Zou J.; Chen Z. G. Advanced Thermoelectric Design: From Materials and Structures to Devices. Chem. Rev. 2020, 120, 7399–7515. 10.1021/acs.chemrev.0c00026. [DOI] [PubMed] [Google Scholar]
  6. Zhou C. J.; Lee Y. K.; Yu Y.; Byun S.; Luo Z. Z.; Lee H.; Ge B. Z.; Lee Y. L.; Chen X. Q.; Lee J. Y.; Cojocaru-Miredin O.; Chang H.; Im J.; Cho S. P.; Wuttig M.; Dravid V. P.; Kanatzidis M. G.; Chung I. Polycrystalline SnSe with a Thermoelectric Figure of Merit Greater Than the Single Crystal. Nat. Mater. 2021, 20, 1378–1385. 10.1038/s41563-021-01064-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Zheng Y.; Slade T. J.; Hu L.; Tan X. Y.; Luo Y. B.; Luo Z. Z.; Xu J. W.; Yan Q. Y.; Kanatzidis M. G. Defect Engineering in Thermoelectric Materials: What Have We Learned?. Chem. Soc. Rev. 2021, 50, 9022–9054. 10.1039/D1CS00347J. [DOI] [PubMed] [Google Scholar]
  8. Liu W. S.; Yan X.; Chen G.; Ren Z. F. Recent Advances in Thermoelectric Nanocomposites. Nano Energy 2012, 1, 42–56. 10.1016/j.nanoen.2011.10.001. [DOI] [Google Scholar]
  9. Qin B. C.; Zhao L. D. Carriers: the Less, the Faster. Mater. Lab 2022, 1, 220004 10.54227/mlab.20220004. [DOI] [Google Scholar]
  10. Luo Z. Z.; Cai S. T.; Hao S. Q.; Bailey T. P.; Luo Y. B.; Luo W. J.; Yu Y.; Uher C.; Wolverton C.; Dravid V. P.; Zou Z. G.; Yan Q. Y.; Kanatzidis M. G. Extraordinary Role of Zn in Enhancing Thermoelectric Performance of Ga-doped n-type PbTe. Energy Environ. Sci. 2022, 15, 368–375. 10.1039/D1EE02986J. [DOI] [Google Scholar]
  11. Xing T.; Zhu C. X.; Song Q. F.; Huang H.; Xiao J.; Ren D. D.; Shi M. J.; Qiu P. F.; Shi X.; Xu F. F.; Chen L. D. Ultralow Lattice Thermal Conductivity and Superhigh Thermoelectric Figure-of-Merit in (Mg, Bi) Co-Doped GeTe. Adv. Mater. 2021, 33, 2008773 10.1002/adma.202008773. [DOI] [PubMed] [Google Scholar]
  12. Cui J.; Wang M. M.; Xu X.; Chen Y.; He J. Q. Understanding the Effects of Iodine Doping on the Thermoelectric Performance of n-type PbTe Ingot Materials. J. Appl. Phys. 2019, 126, 025108 10.1063/1.5101034. [DOI] [Google Scholar]
  13. Xiao Y.; Xu L. Q.; Hong T.; Shi H. N.; Wang S. N.; Gao X.; Ding X. D.; Sun J.; Zhao L. D. Ultrahigh Carrier Mobility Contributes to Remarkably Enhanced Thermoelectric Performance in n-type PbSe. Energy Environ. Sci. 2021, 15, 246–355. 10.1039/D1EE03339E. [DOI] [Google Scholar]
  14. 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. 10.1039/C5TC01781E. [DOI] [Google Scholar]
  15. Dutta M.; Biswas R. K.; Pati S. K.; Biswas K. Discordant Gd and Electronic Band Flattening Synergistically Induce High Thermoelectric Performance in n-type PbTe. ACS Energy Lett. 2021, 6, 1625–1632. 10.1021/acsenergylett.1c00450. [DOI] [Google Scholar]
  16. Xie G. Z.; Li Z.; Luo T. T.; Bai H.; Sun J. C.; Xiao Y.; Zhao L. D.; Wu J. S.; Tan G. J.; Tang X. F. Band Inversion Induced Multiple Electronic Valleys for High Thermoelectric Performance of SnTe with Strong Lattice Softening. Nano Energy 2020, 69, 104395 10.1016/j.nanoen.2019.104395. [DOI] [Google Scholar]
  17. Xiao Y.; Wang D. Y.; Zhang Y.; Chen C. R.; Zhang S. X.; Wang K. D.; Wang G. T.; Pennycook S. J.; Snyder G. J.; Wu H. J.; Zhao L. D. Band Sharpening and Band Alignment Enable High Quality Factor to Enhance Thermoelectric Performance in n-type PbS. J. Am. Chem. Soc. 2020, 142, 4051–4060. 10.1021/jacs.0c00306. [DOI] [PubMed] [Google Scholar]
  18. Hong M.; Lyu W. Y.; Wang Y.; Zou J.; Chen Z. G. Establishing the Golden Range of Seebeck Coefficient for Maximizing Thermoelectric Performance. J. Am. Chem. Soc. 2020, 142, 2672–2681. 10.1021/jacs.9b13272. [DOI] [PubMed] [Google Scholar]
  19. Zhou C.; Yu Y.; Lee Y.-L.; Ge B.; Lu W.; Cojocaru-Miredin O.; Im J.; Cho S.-P.; Wuttig M.; Shi Z.; Chung I. Exceptionally High Average Power Factor and Thermoelectric Figure of Merit in n-type PbSe by the Dual Incorporation of Cu and Te. J. Am. Chem. Soc. 2020, 142, 15172–15186. 10.1021/jacs.0c07712. [DOI] [PubMed] [Google Scholar]
  20. Cao Y.; Bai H.; Li Z.; Zhang Z. K.; Tang Y. F.; Su X. L.; Wu J. S.; Tang X. F. Zn-Induced Defect Complexity for the High Thermoelectric Performance of n-type PbTe Compounds. ACS Appl. Mater. Interfaces 2021, 13, 43134–43143. 10.1021/acsami.1c14518. [DOI] [PubMed] [Google Scholar]
  21. Xiao Y. W.; Wu Y. X.; Nan P. F.; Dong H. L.; Chen Z. W.; Chen Z. Q.; Gu H. K.; Ge B. H.; Li W.; Pei Y. Z. Cu Interstitials Enable Carriers and Dislocations for Thermoelectric Enhancements in n-PbTe0.75Se0.25. Chem 2020, 6, 523–537. 10.1016/j.chempr.2020.01.002. [DOI] [Google Scholar]
  22. Wang S. Q.; Yang Z.; Sun Y. J.; Xiao Y.; Zhao L. D. Synergistically Optimizing Charge and Phonon Transport Properties in n-type PbTe via Introducing Ternary Compound AgSb(Se, Te)2. J. Alloys Compd. 2020, 815, 152463 10.1016/j.jallcom.2019.152463. [DOI] [Google Scholar]
  23. Liu S. X.; Yu Y.; Wu D.; Xu X.; Xie L.; Chao X. L.; Bosman M.; Pennycook S. J.; Yang Z. P.; He J. Q. Coherent Sb/CuTe Core/Shell Nanostructure with Large Strain Contrast Boosting the Thermoelectric Performance of n-type PbTe. Adv. Funct. Mater. 2021, 31, 2007340 10.1002/adfm.202007340. [DOI] [Google Scholar]
  24. Abdellaoui L.; Chen Z. W.; Yu Y.; Luo T.; Hanus R.; Schwarz T.; Villoro R. B.; Cojocaru-Miredin O.; Snyder G. J.; Raabe D.; Pei Y. Z.; Scheu C.; Zhang S. Y. Parallel Dislocation Networks and Cottrell Atmospheres Reduce Thermal Conductivity of PbTe Thermoelectrics. Adv. Funct. Mater. 2021, 31, 2101214 10.1002/adfm.202101214. [DOI] [Google Scholar]
  25. Roychowdhury S.; Ghosh T.; Arora R.; Samanta M.; Xie L.; Singh N. K.; Soni A.; He J. Q.; Waghmare U. V.; Biswas K. Enhanced Atomic Ordering Leads to High Thermoelectric Performance in AgSbTe2. Science 2021, 371, 722–727. 10.1126/science.abb3517. [DOI] [PubMed] [Google Scholar]
  26. Xiang B.; Liu J.; Yan J.; Xia M.; Zhang Q.; Chen L.; Li J.; Tan X. Y.; Yan Q.; Wu Y. Local Nanostructures Enhanced the Thermoelectric Performance of n-type PbTe. J. Mater. Chem. A 2019, 7, 18458–18467. 10.1039/C9TA06247E. [DOI] [Google Scholar]
  27. Grossfeld T.; Sheskin A.; Gelbstein Y.; Amouyal Y. Microstructure Evolution of Ag-Alloyed PbTe-Based Compounds and Implications for Thermoelectric Performance. Crystals 2017, 7, 281 10.3390/cryst7090281. [DOI] [Google Scholar]
  28. Wang H. C.; Bahk J.-H.; Kang C.; Hwang J.; Kim K.; Kim J.; Burke P.; Bowers J. E.; Gossard A. C.; Shakouri A.; Kim W. Right Sizes of Nano- and Microstructures for High-Performance and Rigid Bulk Thermoelectrics. Proc. Natl. Acad. Sci. U.S.A. 2014, 111, 10949–10954. 10.1073/pnas.1403601111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Tan G. J.; Shi F. Y.; Hao S. Q.; Zhao L. D.; Chi H.; Zhang X. M.; Uher C.; Wolverton C.; Dravid V. P.; Kanatzidis M. G. Non-Equilibrium Processing Leads to Record High Thermoelectric Figure of Merit in PbTe-SrTe. Nat. Commun. 2016, 7, 12167 10.1038/ncomms12167. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Zhao L. D.; Wu H. J.; Hao S. Q.; Wu C. I.; Zhou X. Y.; Biswas K.; He J. Q.; Hogan T. P.; Uher C.; Wolverton C.; Dravid V. P.; Kanatzidis M. G. All-scale Hierarchical Thermoelectrics: MgTe in PbTe Facilitates Valence Band Convergence and Suppresses Bipolar Thermal Transport for High Performance. Energy Environ. Sci. 2013, 6, 3346–3355. 10.1039/c3ee42187b. [DOI] [Google Scholar]
  31. Zhu Y.; Hu L.; Zhan S.; Ina T.; Gao X.; Hong T.; Zhao L. D. Breaking the Sodium Solubility Limit for Extraordinary Thermoelectric Performance in p-type PbTe. Energy Environ. Sci. 2022, 15, 3958–3967. 10.1039/D2EE01421A. [DOI] [Google Scholar]
  32. Wu H. J.; Zhao L. D.; Zheng F. S.; Wu D.; Pei Y. L.; Tong X.; Kanatzidis M. G.; He J. Q. Broad Temperature Plateau for Thermoelectric Figure of Merit ZT>2 in Phase-Separated PbTe0.7S0.3. Nat. Commun. 2014, 5, 4515 10.1038/ncomms5515. [DOI] [PubMed] [Google Scholar]
  33. Xiao Y.; Zhao L. D. Charge and Phonon Transport in PbTe-Based Thermoelectric Materials. npj Quant. Mater. 2018, 3, 55 10.1038/s41535-018-0127-y. [DOI] [Google Scholar]
  34. Chen T. T.; Wang H. C.; Su W. B.; Mehmood F.; Wang T.; Zhai J. Z.; Wang X.; Wang C. L. Low Thermal Conductivity and High Figure of Merit for Rapidly Synthesized n-type Pb1-xBixTe Alloys. Dalton Trans. 2018, 47, 15957–15966. 10.1039/C8DT03387K. [DOI] [PubMed] [Google Scholar]
  35. Qin C.; Cheng L. X.; Xiao Y. W.; Wen C. L.; Ge B. H.; Li W.; Pei Y. Z. Substitutions and Dislocations Enabled Extraordinary n-type Thermoelectric PbTe. Mater. Today Phys. 2021, 17, 100355 10.1016/j.mtphys.2021.100355. [DOI] [Google Scholar]
  36. Male J.; Agne M. T.; Goyal A.; Anand S.; Witting I. T.; Stevanović V.; Snyder G. J. The Importance of Phase Equilibrium for Doping Efficiency: Iodine Doped PbTe. Mater. Horiz. 2019, 6, 1444–1453. 10.1039/C9MH00294D. [DOI] [PubMed] [Google Scholar]
  37. Chang C.; Wang D. Y.; He D. S.; He W. K.; Zhu F. Y.; Wang G. T.; He J. Q.; Zhao L. D. Realizing High-Ranged Out-of-Plane ZTs in N-Type SnSe Crystals through Promoting Continuous Phase Transition. Adv. Energy Mater. 2019, 9, 1901334 10.1002/aenm.201901334. [DOI] [Google Scholar]
  38. Zhou C. J.; Yu Y.; Lee Y. K.; Cojocaru-Miredin O.; Yoo B.; Cho S. P.; Im J.; Wuttig M.; Hyeon T.; Chung I. High-Performance n-type PbSe-Cu2Se Thermoelectrics through Conduction Band Engineering and Phonon Softening. J. Am. Chem. Soc. 2018, 140, 15535–15545. 10.1021/jacs.8b10448. [DOI] [PubMed] [Google Scholar]
  39. Xiao Y.; Zhao L. D. Seeking New, Highly Effective Thermoelectrics. Science 2020, 367, 1196–1197. 10.1126/science.aaz9426. [DOI] [PubMed] [Google Scholar]
  40. Pei Y. Z.; May A. F.; Snyder G. J. Self-Tuning the Carrier Concentration of PbTe/Ag2Te Composites with Excess Ag for High Thermoelectric Performance. Adv. Energy Mater. 2011, 1, 291–296. 10.1002/aenm.201000072. [DOI] [Google Scholar]
  41. Wang C.; Zhao X. D.; Ning S. T.; Tao Q. R.; Tang Y. F.; Chen Z. Q.; Wu J. S.; Su X. L.; Tang X. F. Phase Boundary Mapping and Suppressing Pb Vacancies for Enhanced Thermoelectric Properties in n-type Sb Doped PbTe Compounds. Mater. Today Energy 2022, 25, 100962 10.1016/j.mtener.2022.100962. [DOI] [Google Scholar]
  42. Chen S. P.; Wang Y. C.; Wang Y. N.; Fan W. H.; Guo J. Y.; Chen J.; Jiang Y.; Al-Yusufi R.; Munir Z. Thermoelectric and Mechanical Characterization of the Utilization of FeTe as an Electrode for Iodine-Doped PbTe. J. Alloys Compd. 2022, 905, 164267 10.1016/j.jallcom.2022.164267. [DOI] [Google Scholar]
  43. Zhu C.; Zhang J.; Ming H. W.; Huang L. L.; Li Y. Y.; Chen T.; Li D.; Zhang B. L.; Xu J. T.; Qin X. Y. Synergistic Optimization of Electrical and Thermal Transport in n-type Bi-Doped PbTe by Introducing Coherent Nanophase Cu1.75Te. J. Materiomics 2021, 7, 146–155. 10.1016/j.jmat.2020.08.001. [DOI] [Google Scholar]
  44. Wu H. J.; Chang C.; Feng D.; Xiao Y.; Zhang X.; Pei Y. L.; Zheng L.; Wu D.; Gong S. K.; Chen Y.; He J. Q.; Kanatzidis M. G.; Zhao L. D. Synergistically Optimized Electrical and Thermal Transport Properties of SnTe via Alloying High-Solubility MnTe. Energy Environ. Sci. 2015, 8, 3298–3312. 10.1039/C5EE02423D. [DOI] [Google Scholar]
  45. Yelgel Ö. C.; Srivastava G. P. Thermoelectric Properties of n-type Bi2(Te0.85Se0.15)3 Single Crystals Doped with CuBr and SbI3. Phys. Rev. B 2012, 85, 125207 10.1103/PhysRevB.85.125207. [DOI] [Google Scholar]
  46. Qian X.; Wu H. J.; Wang D. Y.; Zhang Y.; Wang J. F.; Wang G. T.; Zheng L.; Pennycook S. J.; Zhao L. D. Synergistically Optimizing Interdependent Thermoelectric Parameters of n-type PbSe Through Alloying CdSe. Energy Environ. Sci. 2019, 12, 1969–1978. 10.1039/C8EE03386B. [DOI] [Google Scholar]
  47. Jia B. H.; Huang Y.; Wang Y.; Zhou Y. S. Y.; Zhao X. D.; Ning S. T.; Xu X.; Lin P. J.; Chen Z. Q.; Jiang B. B.; He J. Q. Realizing High Thermoelectric Performance in Non-Nanostructured n-type PbTe. Energy Environ. Sci. 2022, 15, 1920–1929. 10.1039/D1EE03883D. [DOI] [Google Scholar]
  48. Xiao Y.; Wu H. J.; Wang D. Y.; Niu C. L.; Pei Y. L.; Zhang Y.; Spanopoulos I.; Witting I. T.; Li X.; Pennycook S. J.; Snyder G. J.; Kanatzidis M. G.; Zhao L. D. Amphoteric Indium Enables Carrier Engineering to Enhance the Power Factor and Thermoelectric Performance in n-type AgnPb100InnTe100+2n (LIST). Adv. Energy Mater. 2019, 9, 1900414 10.1002/aenm.201900414. [DOI] [Google Scholar]
  49. You L.; Zhang J. Y.; Pan S. S.; Jiang Y.; Wang K.; Yang J.; Pei Y. Z.; Zhu Q.; Agne M. T.; Snyder G. J.; Ren Z. F.; Zhang W. Q.; Luo J. Realization of Higher Thermoelectric Performance by Dynamic Doping of Copper in n-type PbTe. Energy Environ. Sci. 2019, 12, 3089–3098. 10.1039/C9EE01137D. [DOI] [Google Scholar]
  50. Luo Z.-Z.; Zhang X. M.; Hua X.; Tan G. J.; Bailey T. P.; Xu J. W.; Uher C.; Wolverton C.; Dravid V. P.; Yan Q. Y.; Kanatzidis M. G. High Thermoelectric Performance in Supersaturated Solid Solutions and Nanostructured n-type PbTe-GeTe. Adv. Funct. Mater. 2018, 28, 1801617 10.1002/adfm.201801617. [DOI] [Google Scholar]
  51. Xiao Y.; Wu H. J.; Cui J.; Wang D. Y.; Fu L. W.; Zhang Y.; Chen Y.; He J. Q.; Pennycook S. J.; Zhao L. D. Realizing High Performance n-type PbTe by Synergistically Optimizing Effective Mass and Carrier Mobility and Suppressing Bipolar Thermal Conductivity. Energy Environ. Sci. 2018, 11, 2486–2495. 10.1039/C8EE01151F. [DOI] [Google Scholar]
  52. Su X. L.; Hao S. Q.; Bailey T. P.; Wang S.; Hadar I.; Tan G. J.; Song T.-B.; Zhang Q. J.; Uher C.; Wolverton C.; Tang X. F.; Kanatzidis M. G. Weak Electron Phonon Coupling and Deep Level Impurity for High Thermoelectric Performance Pb1–xGaxTe. Adv. Energy Mater. 2018, 8, 1800659 10.1002/aenm.201800659. [DOI] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

cm2c03542_si_001.pdf (367.6KB, pdf)

Articles from Chemistry of Materials are provided here courtesy of American Chemical Society

RESOURCES