Next Article in Journal
Study on Dynamic Recrystallization under Thermal Cycles in the Process of Direct Energy Deposition for 316 L Austenitic Stainless Steel
Previous Article in Journal
Review of Flexible Robotic Grippers, with a Focus on Grippers Based on Magnetorheological Materials
Previous Article in Special Issue
Phase Transitions and Thermoelectric Properties of Charge-Compensated ZnxCu12−xSb4Se13
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Thermoelectric Characteristics of Permingeatite Compounds Double-Doped with Sn and S

Department of Materials Science and Engineering, College of Engineering, Korea National University of Transportation, Chungju 27469, Republic of Korea
*
Author to whom correspondence should be addressed.
Materials 2024, 17(19), 4859; https://doi.org/10.3390/ma17194859
Submission received: 9 July 2024 / Revised: 27 September 2024 / Accepted: 28 September 2024 / Published: 2 October 2024

Abstract

:
Sn/S double-doped permingeatites, Cu3Sb1−xSnxSe4−ySy (0.02 ≤ x ≤ 0.08 and 0.25 ≤ y ≤ 0.50) were synthesized, and crystallographic parameters and thermoelectric characteristics were examined as a function of doping level. The lattice parameters of permingeatite were significantly modified by the dual doping of Sn and S, with S doping exerting a greater influence on lattice constants and variations in tetragonality compared to Sn doping. With an increase in the level of Sn doping and a decrease in S doping, the carrier concentration increased, leading to enhanced electrical conductivity, indicative of a degenerate semiconducting state. Conversely, an increase in S doping and a decrease in Sn doping led to a rise in the Seebeck coefficient, demonstrating p-type conductivity characteristics with positive temperature dependence. Additionally, the double doping of Sn and S substantially improved the power factor, with Cu3Sb0.98Sn0.02Se3.75S0.25 exhibiting 1.12 mWm−1K−2 at 623 K, approximately 2.3 times higher than that of undoped permingeatite. The lattice thermal conductivity decreased with increasing temperature, while the electronic thermal conductivity exhibited minimal temperature dependence. Ultimately, the dimensionless figure of merit (ZT) was improved through the double doping of Sn and S, with Cu3Sb0.98Sn0.02Se3.50S0.50 recording a ZT of 0.68 at 623 K, approximately 1.7 times higher than that of pure permingeatite.

1. Introduction

The performance evaluation of thermoelectric materials is conducted using a dimensionless figure of merit at the operating temperature (T in Kelvin), ZT = α2σκ−1T, where α2σ represents the power factor (α: Seebeck coefficient and σ: electrical conductivity), and κ (=κE + κL) is thermal conductivity (κE: carrier contribution and κL: phonon contribution) [1,2]. In order to attain a high ZT value, it is imperative to reduce thermal conductivity while simultaneously increasing the power factor. Nevertheless, the augmentation of carrier concentration through doping results in both an increase in electrical conductivity and a reduction in the Seebeck coefficient. Hence, optimizing carrier concentration becomes essential for enhancing the power factor. Additionally, the electrical conductivity correlates with the electronic thermal conductivity as per the Wiedemann-Franz law (κE/σ = LT), where L represents the Lorenz number, contingent upon carrier concentration and temperature [3].
Chalcogenides such as Bi2Te3 and PbTe have long been recognized for their exceptional thermoelectric properties, characterized by high ZT values [4]. However, the scarcity of Te, one of the rarest elements on Earth, along with the toxicity of Pb, presents significant challenges for the widespread application of these materials [5]. In recent years, copper-based chalcogenides have garnered considerable attention as eco-friendly and cost-effective thermoelectric materials. These materials not only avoid the use of toxic and rare elements but also demonstrate high Seebeck coefficients and low lattice thermal conductivities [6]. Among these, Cu3SbSe4, commonly known as permingeatite, has emerged as a promising candidate for thermoelectric applications due to its large carrier effective mass, narrow bandgap, and reduced lattice thermal conductivity [7,8,9].
Despite these favorable properties, pure Cu3SbSe4 suffers from intrinsically low carrier concentration, which limits its electrical conductivity and, consequently, its thermoelectric efficiency [10]. To address this issue, several studies have explored the effects of doping various elements into permingeatite, aiming to enhance its thermoelectric performance. These studies have yielded significant improvements in ZT values, demonstrating the potential of doped Cu3SbSe4 as a viable thermoelectric material. For instance, Cu3Sb0.98Sn0.02Se4 and Cu3Sb0.985Sn0.015Se4 have achieved ZT values of 0.72 at 630 K and 1.05 at 650 K, respectively, while Cu3Sb0.99Sn0.01Se4 has recorded a ZT of 0.127 at 374 K [11,12,13,14,15]. Similarly, sulfur doping has been shown to enhance thermoelectric performance, with Cu3SbSe3.2S0.8 achieving a ZT of 0.32 at 623 K. Notably, Li et al. [16] reported a ZT of 1.1 at 700 K for Cu3Sb0.94Sn0.06Se3.5S0.5, obtained through simultaneous Sn doping at the Sb site and S doping at the Se site. Furthermore, Ahn and Kim [17] achieved a ZT of 0.75 at 623 K for Cu3Sb0.92Sn0.06Bi0.02Se4 by co-doping Sn and Bi at the Sb site, while Liu et al. [18] reported a ZT of 1.26 at 673 K for Cu3Sb0.88Sn0.10Bi0.02Se4.
In this study, we investigate the phase, crystallinity, charge transport, and thermoelectric properties of Sn/S double-doped Cu3SbSe4. The thermoelectric performance of this doped material is subsequently compared with that of undoped Cu3SbSe4, drawing upon results from our previous research [19,20]. This comparative analysis aims to provide a comprehensive evaluation of the potential of Sn/S double-doped Cu3SbSe4 as an efficient thermoelectric material.

2. Experimental Procedure

Cu3Sb1−xSnxSe4−ySy (x = 0.02, 0.04, 0.06, and 0.08; y = 0.25 and 0.50) permingeatites were fabricated via a solid-state process to mitigate the volatilization of constituent elements and ensure uniform synthesis. High-purity (99.9–99.999%) powders of Cu, Sb, Sn, Se, and S were utilized as starting materials for mechanical alloying (MA). MA was carried out at 350 rpm for a duration of 12 h utilizing a planetary ball milling instrument (Pulverisette 5, Fritsch, Pittsboro, NC, USA). Subsequently, hot pressing (HP) was conducted at 573 K for 2 h under 70 MPa. The optimal MA–HP process conditions for permingeatite were determined through our prior investigation [19].
X-ray diffraction (XRD; D8-Advance, Bruker, Billerica, MA, USA) was employed to analyze the phases of the specimens synthesized by MA–HP. Crystallographic data and lattice parameters were derived from the measured diffraction peaks utilizing Rietveld refinement. The sintered body’s microstructure was examined using scanning electron microscopy (SEM; Quanta400, FEI, Lausanne, Switzerland) in backscattered electron (BSE) mode. Additionally, compositional analysis and elemental distribution were verified using energy-dispersive X-ray spectroscopy (EDS; Quantax200, Bruker, Billerica, MA, USA).
Charge transport parameters were analyzed using the Hall-effect measurement system (Keithley 7065, Cleveland, OH, USA). The Seebeck coefficient and electrical conductivity were measured using ZEM-3 (Advance Riko, Yokohama, Japan) equipment, and then the power factor was evaluated. Thermal diffusivity was measured using the laser-flash TC-9000H (Advance Riko, Yokohama, Japan) instrument, and the thermal conductivity was calculated using the theoretical specific heat and density. The ZT values were examined based on the power factor and thermal conductivity measured within the temperature range of 323–623 K.

3. Results and Discussion

Figure 1 shows the XRD patterns of Cu3Sb1−xSnxSe4−ySy fabricated by the MA–HP process. The diffraction peaks indicated a single phase corresponding to tetragonal permingeatite with the I 4 ¯ 2 m space group, matching the standard diffraction data (PDF# 00-066-0482). As shown in Table 1, the relative density of the Sn/S double-doped specimens prepared in this study ranged from 95.2% to 99.7% in comparison to the theoretical density (5.82 gcm−3) of pure Cu3SbSe4 [21]. The lattice constants of pure Cu3SbSe4 were reported as a = 0.5649 nm, c = 1.1243 nm, and c/a = 1.9904 [20]. In this study, with the S content of y = 0.25, the lattice constants decreased to a = 0.5635(2)–0.5636(2) nm and c = 1.1219(7)–1.1223(5) nm, resulting in c/a = 1.9906–1.9913. However, for y = 0.50, the a-axis shrank to 0.5618(2)–0.5619(3) nm while the c-axis expanded significantly to 1.1874(7)–1.1904(6) nm, which resulted in c/a = 2.1132–2.1189, indicating a substantial increase in tetragonality. When the S content remained constant, an increase in Sn content resulted in minimal changes in the a-axis lattice constant but a slight decrease in the c-axis lattice constant. The variation in lattice constants due to Sn and S doping was speculated to be attributed to the interaction between the differences in ionic radii of Sb5+ (60 pm), Sn4+ (69 pm), Se2− (198 pm), and S2− (184 pm) [22]. Bhardwaj et al. [14] reported that as the Sn doping amount increased in Cu3Sb1-xSnxSe4 (x = 0.005–0.175), the c-axis decreased to 1.136–1.130 nm, and the c/a ratio also decreased to 2.013–1.992. Lee and Kim [15] recorded that Cu3SbSe3.2S0.8 among Cu3SbSe4−ySy (y = 0–4) had a = 0.542 nm and c = 1.082 nm, with both axes increasing as the S content increased.
In pure permingeatite Cu3SbSe4, Cu1, Cu2, Sb1, and Se1 occupied their Wyckoff positions (atomic coordinates) of 2b (0, 0, 0.5), 4d (0, 0.5, 0.25), 2a (0, 0, 0), and 8i (0.2410, 0.2410, and 0.36871), respectively. Additionally, in Sn/S double-doped permingeatite Cu3Sb1−xSnxSe4−ySy, Sn substituted the Sb site (2a) and S substituted the Se site (8i). Min et al. [23] also reported very similar results (Wyckoff positions, atomic coordinates, and occupancies) for Sn-doped permingeatite Cu3Sb1−xSnxSe4. The positions of the constituent elements were the same as in our results, with slight differences in atomic coordinates and occupancies, which were interpreted as due to differences in the types and amounts of dopants. In this study, the crystallite size of Sn/S double-doped permingeatite Cu3Sb1−xSnxSe4−ySy ranged from 55.68(6) to 78.77(8) nm, and the lattice microstrain was analyzed to be 0.736(2)–0.819(2)%.
Figure 2 and Figure 3 show BSE–SEM micrographs and EDS analysis results of Cu3Sb1−xSnxSe4−ySy. Elemental line scans and two-dimensional maps confirmed a dense and homogeneous phase of permingeatite, and all elements were uniformly distributed. No peculiarities in microstructure were found according to the contents of Sn and S.
The analysis results of the measured charge-transport parameters are summarized in Table 1. A comparison between our findings and those of other researchers reveals two significant trends. First, regarding the effect of Sn doping on carrier concentration, this study shows that Cu3Sb0.94Sn0.06Se3.75S0.25 recorded the highest carrier concentration of 1.38 × 1019 cm−3, indicating that Sn doping significantly increased the carrier concentration. Similarly, Skoug et al. [11] reported that Sn doping in Cu3Sb1−xSnxSe4−ySy (x = 0.01–0.03) increased carrier concentrations to the order of 1019–1021 cm−3. This suggests that Sn doping is a key mechanism for increasing carrier concentration. In contrast, in this study, S doping slightly decreased the carrier concentration. This aligns with the findings of Lee and Kim [15], who also observed a decrease in both carrier concentration and mobility with increasing S doping in Cu3SbSe3.2S0.8. On the other hand, Li et al. [16] found that in Cu3Sb0.90Sn0.10Se3.5S0.5, both the carrier concentration and mobility increased with increasing Sn content, which suggests that the effect of Sn doping outweighed the impact of S doping. In terms of mobility, this study recorded a notably high mobility of 285 cm2V−1s−1 for Cu3Sb0.94Sn0.06Se3.75S0.25, which is much higher than the 10 cm2V−1s−1 reported by Skoug et al. [11] for Sn-doped samples, likely due to less scattering by ionized impurities. Pi et al. [20] reported a mobility of 50 cm2V−1s−1 for pure Cu3SbSe4, which falls within the range of 47–230 cm2V−1s−1 for Cu3Sb1−xSnxSe4 (x = 0–0.04) reported by Prasad and Rao [13]. Overall, the trends indicate that Sn doping consistently increases carrier concentration, albeit often at the expense of mobility, while S doping generally has a detrimental effect on both carrier concentration and mobility.
Figure 4 represents the electrical conductivity of Cu3Sb1−xSnxSe4−ySy. A slight decrease in electrical conductivity with increasing temperature was observed across all specimens, indicating degenerate semiconducting behavior. This is consistent with the results of Pi et al. [20], who reported similar behavior in undoped permingeatite, with electrical conductivities ranging from (4.2–4.5) × 103 Sm−1 over the temperature range of 323–623 K. Similarly, Skoug et al. [11] reported that Cu3Sb0.97Sn0.03Se4 exhibited conductivity values from (10.0–3.3) × 104 Sm−1 at 320–630 K, with a decrease in conductivity as temperature increased, following a similar trend to the one observed in this study. An increase in Sn doping resulted in higher electrical conductivity, which correlates well with the observed rise in carrier concentration. Specifically, Cu3Sb0.96Sn0.04Se3.75S0.25 showed the highest electrical conductivity of (5.1–3.7) × 104 Sm−1 in the 323–623 K range. This is in line with the results of Skoug et al. [11], where Sn-doped Cu3Sb0.97Sn0.03Se4 showed comparable conductivity values, indicating that Sn doping significantly enhances electrical conductivity by increasing carrier concentration. Ahn and Kim [17] also found that Cu3Sb0.92Sn0.06Bi0.02Se4 exhibited conductivity values of (8.3–5.8) × 104 Sm−1 within the same temperature range, further supporting the role of Sn doping in improving conductivity. In this study, the combination of Sn and S doping in Cu3Sb0.96Sn0.04Se3.75S0.25 maintained relatively high conductivity levels. This result aligns with the findings of Li et al. [8], who reported even higher conductivity values of (1.0–0.4) × 105 Sm−1 for Cu3Sb0.94Sn0.06Se3.5S0.5 at 300–673 K. The double doping of Sn and S in their study also resulted in a significant increase in conductivity, suggesting that S doping, when combined with Sn, can further optimize the material’s electrical performance. In contrast, Lee and Kim [15] observed lower conductivity values of (0.8–7.0) × 103 Sm−1 for Cu3SbSe3.2S0.8 over the 373–623 K temperature range, where S doping alone appeared to reduce the electrical conductivity compared to the undoped material or Sn-doped systems. Therefore, Sn doping effectively enhances electrical conductivity, especially at higher doping levels, which is consistent with other findings in the literature. However, S doping, when used in combination with Sn, maintains high conductivity levels, while S doping alone tends to decrease the electrical conductivity. The combination of Sn and S doping provides an effective strategy to optimize the electrical conductivity in permingeatite compounds, particularly in applications requiring degenerate semiconducting behavior over a wide temperature range.
Figure 5 shows the Seebeck coefficient of Cu3Sb1−xSnxSe4−ySy. An increase in Sn doping, while the S content remained constant, led to a reduction in the Seebeck coefficient. This is attributed to the inverse relationship between the Seebeck coefficient and carrier concentration, where Sn doping increases the carrier concentration and thus lowers the Seebeck coefficient. For example, Cu3Sb0.94Sn0.02Se3.50S0.50 exhibited Seebeck coefficient values of 135–222 μVK−1 over the temperature range of 323–623 K. This result is consistent with the findings of Skoug et al. [11], who reported a Seebeck coefficient of 75–120 μVK−1 for Cu3Sb0.99Sn0.01Se4 at 320–630 K, demonstrating a similar reduction in the Seebeck coefficient with increasing Sn content. Ahn and Kim [17] also observed that increasing Sn doping in Cu3Sb1−x−ySnxBiySe4 led to a reduction in the Seebeck coefficient, with values of 141–225 μVK−1 for Cu3Sb0.94Sn0.02Bi0.04Se4 at 323–623 K, further reinforcing the relationship between higher carrier concentration due to Sn doping and a lower Seebeck coefficient. In contrast, when the Sn content was held constant in this study, increasing the S doping resulted in an increase in the Seebeck coefficient. This is likely due to S doping reducing the carrier concentration, thereby enhancing the Seebeck coefficient. This trend aligns with the results reported by Lee and Kim [15], where Cu3SbSe3.2S0.8 exhibited a maximum Seebeck coefficient of 400 μVK−1 at 523 K, indicating that S doping enhances the Seebeck coefficient. Li et al. [16] also found that Cu3Sb0.98Sn0.02Se3.5S0.5 exhibited Seebeck coefficients in the range of 140–240 μVK−1 at 300–700 K, which supports the observation that S doping can offset the reduction in Seebeck coefficient caused by Sn doping. Pi et al. [20] reported a peak Seebeck coefficient of 348 μVK−1 at 523 K for pure Cu3SbSe4, which is significantly higher than the Seebeck coefficients observed in Sn-doped samples. This further highlights the impact of increased carrier concentration due to Sn doping, which reduces the Seebeck coefficient compared to undoped materials. The values reported in this study, as well as those by Skoug et al. [11] and Ahn and Kim [17], indicate a consistent trend of reduced Seebeck coefficients in Sn-doped samples relative to undoped Cu3SbSe4. In cases of double doping, different dopants can have varying effects on the Seebeck coefficient, and optimal doping strategies may involve balancing dopants that influence carrier concentration in opposite ways. Sn doping increases the carrier concentration, leading to a reduction in the Seebeck coefficient, while S doping counteracts this effect by decreasing the carrier concentration, resulting in an increased Seebeck coefficient. These findings suggest that careful tuning of Sn and S doping levels is crucial to optimizing thermoelectric performance, particularly when targeting a balance between electrical conductivity and the Seebeck coefficient.
Figure 6 shows the power factor of Cu3Sb1−xSnxSe4−ySy. The power factor was significantly enhanced through double doping with Sn and S. The highest power factor values of 0.57–1.12 mWm−1K−2 were observed in Cu3Sb0.98Sn0.02Se3.75S0.25 at 323–623 K, representing a 2.3-fold improvement over undoped Cu3SbSe4, which exhibited a power factor of 0.49 mWm−1K−2 at 623 K [20]. This increase can be attributed to the optimization of carrier concentration through the dual doping strategy, which balances the opposing effects of carrier concentration on electrical conductivity and the Seebeck coefficient. Skoug et al. [11] reported power factor values of 0.1–1.3 mWm−1K−2 for Cu3Sb0.98Sn0.02Se4 and Cu3Sb0.97Sn0.03Se4 at 320–630 K. These values are comparable to those observed in this study, particularly at higher temperatures, indicating that Sn doping is effective in increasing the power factor. The slightly higher maximum value reported by Skoug et al. (1.3 mWm−1K−2) can be attributed to a higher Sn doping level (x = 0.03), which further enhances carrier concentration but also reduces the Seebeck coefficient, limiting the overall power factor improvement at lower doping levels. In contrast, the study by Lee and Kim [15] reported a lower power factor range of 0.29–0.31 mWm−1K−2 for Cu3SbSe3.2S0.8 at 323–623 K. This suggests that S doping alone, without the addition of Sn, results in a moderate increase in the power factor compared to undoped Cu3SbSe4, but does not achieve the same level of improvement as observed with dual Sn and S doping. The lower carrier concentration resulting from S doping likely contributes to this limitation. Li et al. [8] achieved a power factor of 0.96 mWm−1K−2 at 640 K for Cu3Sb0.94Sn0.06Se3.5S0.5, which is consistent with our observation that dual Sn and S doping enhances the power factor. The slightly higher Sn content (x = 0.06) in their study contributed to the increased power factor at elevated temperatures, reinforcing the conclusion that double doping is an effective strategy for optimizing thermoelectric performance. Similarly, Ahn and Kim [17] reported power factor values of 0.64–1.29 mWm−1K−2 for Cu3Sb0.92Sn0.06Bi0.02Se4 at 323–623 K, demonstrating the effectiveness of co-doping (in this case, with Sn and Bi) in improving the power factor by balancing the carrier concentration and optimizing the Seebeck coefficient.
Figure 7 represents the thermal conductivity of Cu3Sb1−xSnxSe4−ySy. The thermal conductivity followed a temperature dependence of T−1. In the measured temperature range, no bipolar conduction occurred, showing thermal conductivities of 1.24–1.80 Wm−1K−1 at 323 K and 0.81–1.32 Wm−1K−1 at 623 K., with the lowest values recorded for Cu3Sb0.94Sn0.04Se3.50S0.50. This indicates that the dual doping of Sn and S successfully suppressed thermal conductivity, with a substantial reduction at higher temperatures. The absence of bipolar conduction in our results suggests that the material maintains a single-carrier transport mechanism over the measured temperature range. Pi et al. [20] reported that undoped Cu3SbSe4 exhibited thermal conductivities of 1.19–0.74 Wm−1K−1 over the temperature range of 323–623 K. These values are comparable to those observed in this study, particularly at higher temperatures, indicating that the Sn and S doping approach achieves a similar degree of thermal conductivity reduction as undoped materials. Skoug et al. [11] reported significantly higher thermal conductivities for Sn-doped Cu3Sb1−xSnxSe4 (x = 0.02–0.03), with values of 3.5–3.6 Wm−1K−1 at 323 K and 1.5–2.0 Wm−1K−1 at 623 K. This suggests that increasing Sn content, while effective at increasing carrier concentration, can also lead to higher thermal conductivities due to reduced phonon scattering. The much higher thermal conductivity observed in Skoug et al.’s study compared to our results indicates that the additional S doping in our samples plays a critical role in reducing thermal conductivity by enhancing phonon scattering. The study by Lee and Kim [15] on Cu3SbSe4−ySy (y = 0–4) showed that increasing S doping results in lower thermal conductivities. For example, Cu3SbSe3.2S0.8 exhibited values of 1.1 Wm−1K−1 at 323 K and 0.7 Wm−1K−1 at 623 K. These values are in line with those observed in this study, reinforcing the conclusion that S doping is highly effective in reducing thermal conductivity. The combination of Sn and S doping provides a synergistic effect in reducing thermal conductivity, particularly at higher temperatures. Li et al. [8] demonstrated that dual Sn and S doping can significantly lower thermal conductivity. Their results for Cu3Sb0.94Sn0.06Se4−ySy (y = 0.5–1.5) indicated thermal conductivities of 1.7–0.9 Wm−1K−1 at 300 K and 673 K, with Cu3Sb0.94Sn0.06Se2.5S1.5 showing the lowest thermal conductivity. These findings are consistent with our results and demonstrate the effectiveness of dual doping in reducing thermal conductivity by enhancing phonon scattering and reducing lattice thermal transport. Ahn and Kim [17] reported that Cu3Sb0.94Sn0.02Bi0.04Se4 exhibited a minimum thermal conductivity of 0.91 Wm−1K−1 at 523 K. This is slightly higher than the lowest values observed in this study, suggesting that while Bi doping also reduces thermal conductivity, the combined Sn and S doping approach in this work is more effective in achieving lower thermal conductivity, particularly at higher temperatures.
The thermal conductivity is expressed as the sum of these two components determined by carrier-mediated heat transfer and phonon-mediated heat transfer, assuming no bipolar effect [24]. Figure 8 distinguishes between the electronic thermal conductivity and lattice thermal conductivity of Cu3Sb1−xSnxSe4−ySy. In this study, the Lorenz number (L) was calculated by putting the measured Seebeck coefficient values into the formula L = 1.5 + exp(−|α|/116) [25] for thermal conductivity separation, and the L values are presented in Table 1. Values ranging (1.81–2.03) × 10−8 V2K−2 at 323 K were obtained, which fall within the expected theoretical values of (1.45–2.44) × 10−8 V2K−2. These values align with degenerate semiconducting behavior, confirming that the materials maintain efficient electron transport while minimizing lattice thermal conductivity. Pi et al. [20] reported slightly lower Lorenz numbers for undoped Cu3SbSe4, ranging from (1.57–1.56) × 10−8 V2K−2 over the temperature range of 323–623 K. The relatively lower values observed in their study could be attributed to the absence of doping, which typically enhances carrier concentration and can lead to increased electronic contribution to thermal conductivity. In contrast, the higher Lorenz numbers in this study may reflect the impact of Sn and S doping on the electronic structure, leading to a higher electronic thermal conductivity. Ahn and Kim [17] reported Lorenz numbers ranging from (1.80–1.56) × 10−8 V2K−2 at 323 K for Cu3Sb1−x−ySnxBiySe4, with decreasing values as the temperature increased. Their results are similar to our data at 323 K, indicating that Sn doping enhances the electronic thermal conductivity. However, the slight reduction in Lorenz numbers with increasing temperature in their study suggests that phonon scattering becomes more significant at higher temperatures, which reduces the overall electronic contribution to thermal conductivity. Lee and Kim [15] obtained a Lorenz number of 1.54 × 10−8 V2K−2 at 323 K for Cu3SbSe3.2S0.8, noting a decreasing trend in the Lorenz number with increasing S content. This is consistent with the idea that higher S doping reduces carrier concentration, leading to a smaller electronic contribution to thermal conductivity. In this study, where both Sn and S were doped, the Lorenz numbers were slightly higher, suggesting that Sn doping increased the carrier concentration and thus the electronic thermal conductivity, despite the presence of S.
In Figure 8a, almost no temperature dependence of electronic thermal conductivity was found, with Cu3Sb0.92Sn0.02Se3.50S0.50 exhibiting the lowest values of 0.16–0.18 Wm−1K−1 in the temperature range of 323–623 K. The consistent values across the temperature range reflect degenerate semiconducting behavior, where the electronic contribution to thermal conductivity remains stable. Additionally, the relatively low κE values can be attributed to the combined effect of Sn and S doping, which not only enhances carrier concentration but also introduces scattering mechanisms that limit electronic heat transport. Pi et al. [20] reported significantly lower electronic thermal conductivity values for undoped Cu3SbSe4, ranging from 0.02 to 0.04 Wm−1K−1 at 323–623 K. The much lower κE values in their study can be attributed to the lack of doping, which results in a lower carrier concentration and thus a minimal contribution of electrons to thermal conductivity. This highlights the effect of Sn and S doping in this study, where enhanced carrier concentrations lead to higher electronic thermal conductivity. Lee and Kim [15] observed that in Cu3SbSe3.2S0.8, the electronic thermal conductivity approached almost zero across the entire temperature range of 323–623 K. This is consistent with the reduced carrier concentration due to S doping, which decreases the electronic contribution to thermal transport. The absence of significant electronic heat transport in their study contrasts with our results, where Sn doping compensates for the reduction caused by S, resulting in a non-negligible κE. Ahn and Kim [17] reported electronic thermal conductivities of 0.15–0.19 Wm−1K−1 at 323–623 K for Cu3Sb0.94Sn0.02Bi0.04Se4, values that are comparable to this study. The similarity suggests that both Sn and Bi doping at the Sb site maintain a stable electronic contribution to thermal conductivity. However, their results also show that the κE values remain relatively low, similar to this study, indicating that both Sn and Bi introduce scattering mechanisms that suppress the increase in electronic thermal conductivity, despite increased carrier concentrations.
In Figure 8b, the lattice thermal conductivity exhibits temperature dependence of T−1, ranging from 1.09 to 1.47 Wm−1K−1 at 323 K and 0.63–0.92 Wm−1K−1 at 623 K. This suggests that Umklapp scattering is the predominant mechanism in phonon transport [15]. Given that the lattice thermal conductivity of each specimen significantly surpasses the electronic thermal conductivity, it can be inferred that the lattice thermal conductivity predominantly affects the total thermal conductivity. The clear influence of doping was observed, where increasing Sn doping (with constant S content) increased the lattice thermal conductivity, while increasing S doping (with constant Sn content) decreased the lattice thermal conductivity. This indicates that S doping is more effective in reducing the lattice thermal conductivity than Sn doping, as shown by the lowest values of κL for Cu3Sb0.98Sn0.02Se3.50S0.50. Pi et al. [20] reported that undoped Cu3SbSe4 exhibited lattice thermal conductivities of 1.17 Wm−1K−1 at 323 K and 0.72 Wm−1K−1 at 623 K, which are close to the κL values for Cu3Sb0.98Sn0.02Se3.50S0.50. However, the slight decrease in the lattice thermal conductivity with S doping in this study indicates the effect of alloy scattering introduced by the S atoms, which further reduces phonon transport compared to the undoped material. Li et al. [12] reported significantly higher initial lattice thermal conductivities of 2.60 Wm−1K−1 at 300 K for Cu3Sb0.98Sn0.02Se4, decreasing to 0.63 Wm−1K−1 at 673 K. Li et al. [16] reported that as Sn doping levels increased in Cu3Sb1−xSnxSe3.5S0.5 (x = 0–0.10), the lattice thermal conductivity was found to decrease, with the lowest values recorded at 0.75–0.22 Wm−1K−1 in the temperature range of 300–700 K. This behavior aligns with our observation that Sn doping can increase the lattice thermal conductivity at lower doping levels, but excessive doping leads to enhanced scattering and a reduction in the lattice thermal conductivity. In agreement with our findings, Li et al. [8] observed that increasing the doping amount of S led to a gradual decrease in the lattice thermal conductivity, reaching very low values of 0.17–0.22 Wm−1K−1 at 673 K for Cu3Sb0.94Sn0.06Se4−ySy (y = 0.5–1.5). Similarly, Lee and Kim [15] found that Cu3SbSe1.6S2.4 exhibited lattice thermal conductivities ranging from 0.84 to 0.56 Wm−1K−1 at 323–623 K, which is lower than the values for undoped Cu3SbSe4. These results confirm the effectiveness of S doping in significantly reducing the lattice thermal conductivity by introducing phonon scattering at higher rates than Sn doping. Ahn and Kim [17] reported a decrease in the lattice thermal conductivity with increasing Sn doping when Bi content was held constant in Cu3Sb1−x−ySnxBiySe4 (x = 0.02–0.06; y = 0.02–0.04). The lowest lattice thermal conductivities (1.10–0.44 Wm−1K−1 at 323–623 K) were observed for Cu3Sb0.92Sn0.06Bi0.02Se4, which compares closely to the results in this study for Cu3Sb0.98Sn0.02Se3.50S0.50. Both studies highlight how combining Sn with another dopant (Bi or S) enhances phonon scattering and leads to further reductions in lattice thermal conductivity.
Figure 9 shows the ZT values of Cu3Sb1−xSnxSe4−ySy. At elevated temperatures, the ZT value experienced substantial enhancement due to the double doping of Sn and S. Specifically, Cu3Sb0.86Sn0.02Se3.50S0.50 recorded a peak ZT of 0.68 at 623 K. The elevated ZT at higher temperatures reflects the balance between improved electrical conductivity and reduced lattice thermal conductivity due to the phonon scattering induced by the double dopants. Compared to undoped Cu3SbSe4 prepared by Pi et al. [20] using MA–HP, which showed relatively low ZT values of 0.11–0.39 in the temperature range of 323–623 K, the Sn/S double-doped permingeatite in this study exhibited a remarkable improvement in ZT. Skoug et al. [11] achieved a slightly higher peak ZT of 0.72 at 630 K for Cu3Sb0.98Sn0.02Se4, prepared using melting–quenching–annealing–HP. This result is comparable to the ZT obtained in this study for the Sn/S double-doped sample, demonstrating the beneficial effect of Sn doping, though the addition of S leads to a more balanced performance, especially in reducing thermal conductivity. Prasad and Rao [13] reported a much lower ZT of 0.127 at 374 K for Cu3Sb0.99Sn0.01Se4 synthesized via solid-state reaction and vacuum heating. The lower ZT values in their study may be attributed to the synthesis method, which may not have optimized the material’s microstructure as effectively as the HP or spark plasma sintering (SPS) methods used in other studies. Bhardwaj et al. [14] achieved a ZT of 1.08 at 623 K for Cu3Sb0.985Sn0.015Se4 synthesized using conventional fusion followed by SPS. This ZT value is notably higher than the value obtained in our study. The increased ZT can likely be attributed to the SPS technique, which often results in superior densification and enhanced electrical properties. However, the presence of S doping in this study might have contributed to better thermal performance, even though the overall ZT remains slightly lower. Li et al. [16] reported a peak ZT of 1.1 at 700 K for Sn/S double-doped Cu3Sb0.94Sn0.06Se3.5S0.5, synthesized using a co-precipitation method combined with HP. Their results demonstrate the highest ZT among the studies, likely due to the higher Sn doping content and the synergistic effect of both dopants on reducing the thermal conductivity and enhancing the Seebeck coefficient at higher temperatures. The co-precipitation method used in their study also likely contributed to improved material homogeneity, which might explain the higher ZT values compared to those obtained in this study. Ahn and Kim [17] reported a ZT of 0.75 at 623 K for Sn/Bi double-doped Cu3Sb0.92Sn0.06Bi0.02Se4, synthesized using MA–HP. Their result is slightly higher than the ZT value in our study, showing that Bi doping, in combination with Sn, can also significantly enhance thermoelectric performance. This suggests that Bi may serve as a potent dopant alongside Sn, although in our case, S appears to have more effectively reduced thermal conductivity.

4. Conclusions

Double-doped Cu3Sb1−xSnxSe4−ySy (x = 0.02–0.08 and y = 0.25–0.50) permingeatites were synthesized through mechanical alloying combined with hot pressing. The tetragonal permingeatite phase could be synthesized, and all elements were uniformly distributed. Double doping of Sn and S resulted in changes in lattice parameters, implying the successful substitutions of Sn and S into the permingeatite structure. Sn doping increased carrier concentration, enhancing electrical conductivity but negatively impacting mobility. In contrast, S doping adversely affected both carrier concentration and mobility, reducing electrical conductivity when used alone. However, the combination of Sn and S doping maintained high electrical conductivity while effectively suppressing thermal conductivity. Ultimately, Sn doping lowered the Seebeck coefficient, while S doping counteracted this effect, highlighting the importance of carefully tuning both doping levels to optimize thermoelectric performance. Carrier concentration was controlled by double doping of Sn and S, and thus significant enhancement in the power factor was observed, with Cu3Sb0.98Sn0.02Se3.75S0.25 showing a maximum value of 1.12 mWm−1K−2 at 623 K. However, Cu3Sb0.98Sn0.02Se3.50S0.50 recorded the highest ZT value of 0.68, interpreted as having the lowest thermal conductivity. The introduction of double doping with Sn and S into permingeatite, prepared via the solid-state mechanical alloying–hot pressing (MA–HP) process, contributed to the enhancement of thermoelectric performance.

Author Contributions

Conceptualization, B.-K.H. and I.-H.K.; methodology, B.-K.H. and I.-H.K.; software, B.-K.H.; validation, I.-H.K.; formal analysis, B.-K.H.; investigation, B.-K.H.; resources, B.-K.H.; data curation, B.-K.H.; writing—original draft preparation, B.-K.H.; writing—review and editing, I.-H.K.; visualization, B.-K.H.; supervision, I.-H.K.; project administration, I.-H.K.; funding acquisition, I.-H.K. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the Basic Science Research Capacity Enhancement Project (National Research Facilities and Equipment Center) through the Korea Basic Science Institute funded by the Ministry of Education (grant No. 2019R1A6C1010047).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the article. Further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Wei, T.R.; Qin, Y.; Deng, T.; Song, Q.; Jiang, B.; Liu, R.; Qiu, P.; Shi, X.; Chen, L. Copper chalcogenide thermoelectric materials. Sci. China Mater. 2019, 62, 8–24. [Google Scholar] [CrossRef]
  2. Li, J.M.; Li, D.; Song, C.J.; Wang, L.; Xin, H.X.; Zhang, J.; Qin, X.Y. Realized high power factor and thermoelectric performance in Cu3SbSe4. Intermetallics 2019, 109, 68–73. [Google Scholar] [CrossRef]
  3. García, G.; Palacios, P.; Cabot, A.; Wahnon, P. Thermoelectric properties of doped-Cu3SbSe4 compounds: A first-principles insight. Inorg. Chem. 2018, 57, 7321–7333. [Google Scholar] [CrossRef] [PubMed]
  4. Wei, T.R.; Li, F.; Li, J.F. Enhanced thermoelectric performance of nonstoichiometric compounds Cu3−xSbSe4 by Cu deficiencies. J. Electron. Mater. 2014, 43, 2229–2238. [Google Scholar] [CrossRef]
  5. Kumar, A.; Dhama, P.; Saini, D.S.; Banerji, P. Effect of Zn substitution at a Cu site on the transport behavior and thermoelectric properties in Cu3SbSe4. RSC Adv. 2016, 6, 5528–5534. [Google Scholar] [CrossRef]
  6. Kumar, A.; Dhama, P.; Das, A.; Sarkar, K.J.; Banerji, P. Enhanced electrical transport and thermoelectric properties in Ni doped Cu3SbSe4. AIP Conf. Proc. 2018, 1953, 050030. [Google Scholar]
  7. Zhao, D.; Wu, D.; Bo, L. Enhanced thermoelectric properties of Cu3SbSe4 compounds via gallium doping. Energies 2017, 10, 1524. [Google Scholar] [CrossRef]
  8. Li, D.; Ming, H.W.; Li, J.M.; Jabar, B.; Xu, W.; Zhang, J.; Qin, X.Y. Ultralow thermal conductivity and extraordinary thermoelectric performance realized in codoped Cu3SbSe4 by plasma spark sintering. ACS Appl. Mater. Interf. 2020, 12, 3886–3892. [Google Scholar] [CrossRef]
  9. Wu, Y.; Qiao, X.; Fan, X.; Zhang, X.; Cui, S.; Wan, J. Facile synthesis of monodisperse Cu3SbSe4 nanoparticles and thermoelectric performance of Cu3SbSe4 nanoparticle-based materials. J. Nanopart. Res. 2015, 17, 285. [Google Scholar] [CrossRef]
  10. Dou, Y.; Zhu, Q.; Du, Y.; Xu, J.; Li, D. Enhanced thermoelectric performance of Cu3SbSe4 doped with alkali-ion (Na and K). Electron. Mater. Lett. 2020, 16, 99–105. [Google Scholar] [CrossRef]
  11. Skoug, E.J.; Cain, J.D.; Majsztrik, P.; Kirkham, M.; Lara-Curzio, E.; Morelli, D.T. Doping effects on the thermoelectric properties of Cu3SbSe4. Sci. Adv. Mater. 2011, 3, 602–606. [Google Scholar] [CrossRef]
  12. Li, D.; Li, R.; Qin, X.Y.; Song, C.J.; Xin, H.X.; Wang, L.; Zhang, J.; Guo, G.L.; Zou, T.H.; Liu, Y.F.; et al. Co-precipitation synthesis of nanostructured Cu3SbSe4 and its Sn-doped sample with high thermoelectric performance. Dalton Trans. 2014, 43, 1888–1896. [Google Scholar] [CrossRef] [PubMed]
  13. Prasad, K.S.; Rao, A. Enhancement in the thermoelectric properties of Cu3SbSe4 by Sn doping. J. Mater. Sci. Mater. Electron. 2019, 30, 16596–16605. [Google Scholar] [CrossRef]
  14. Bhardwaj, R.; Bhattacharya, A.; Tyagi, K.; Ghatori, B.; Chauhan, N.S.; Bathula, S.; Auluck, S.; Dhar, A. Tin doped Cu3SbSe4: A stable thermoelectric analogue for the mid-temperature applications. Mater. Res. Bull. 2019, 113, 38–44. [Google Scholar] [CrossRef]
  15. Lee, G.E.; Kim, I.H. Preparation and investigation of thermoelectric properties of Cu3SbS4-Cu3SbSe4 solid solutions. Korean J. Met. Mater. 2022, 60, 384–390. [Google Scholar] [CrossRef]
  16. Li, D.; Li, R.; Qin, X.Y.; Zhang, J.; Song, C.J.; Wang, L.; Xin, H.X. Co-precipitation synthesis of Sn and/or S doped nanostructured Cu3Sb1−xSnxSe4−ySy with a high thermoelectric performance. CrystEngComm 2013, 15, 7166–7170. [Google Scholar] [CrossRef]
  17. Ahn, H.J.; Kim, I.H. Thermoelectric performance of Sn and Bi double-doped permingeatite. Korean J. Met. Mater. 2022, 60, 593–601. [Google Scholar] [CrossRef]
  18. Liu, Y.; Garcia, G.; Ortega, S.; Cadavid, D.; Palacios, P.; Lu, J.; Ibanez, M.; Xi, L.; Roo, J.D.; Lopez, A.M.; et al. Solution-based synthesis and processing of Sn- and Bi-doped Cu3SbSe4 nanocrystals, nanomaterials and ring-shaped thermoelectric generators. J. Mater. Chem. A 2017, 5, 2592–2602. [Google Scholar] [CrossRef]
  19. Lee, G.E. Thermoelectric Properties of Cu3Sb(S/Se)4 and Cu3Sb(S/Se)3 Prepared by Solid-State Synthesis. Ph.D. Thesis, Korea National University of Transportation, Chungju, Republic of Korea, 2021. [Google Scholar]
  20. Pi, J.H.; Lee, G.E.; Kim, I.H. Effects of Ge doping on the charge transport and thermoelectric properties of permingeatites Cu3Sb1−yGeySe4. Korean J. Met. Mater. 2021, 59, 422–429. [Google Scholar] [CrossRef]
  21. Anthony, J.W.; Bideaux, K.W.; Nichols, M.C. Handbook of Mineralogy, Vol. 1: Elements, Sulfides, Sulfosalts; Mineral Data Publishing: Tucson, AZ, USA, 2003; p. 1613. [Google Scholar]
  22. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Cryst. A 1976, 32, 751–767. [Google Scholar] [CrossRef]
  23. Min, L.; Xia, Y.; Ying, P.; Cui, J. Hybrid structure responsible for improved thermoelectric performance of Sn-incorporated Cu3SbSe4 with a second phase CuSe. J. Appl. Phys. 2020, 127, 235104. [Google Scholar] [CrossRef]
  24. Li, Y.; Qin, X.; Li, D.; Li, X.; Liu, Y.; Zhang, J.; Song, C.; Xin, H. Transport properties and enhanced thermoelectric performance of aluminum doped Cu3SbSe4. RSC Adv. 2015, 5, 31399–31403. [Google Scholar] [CrossRef]
  25. 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, 041506. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of Cu3Sb1−xSnxSe4−ySy prepared by the MA–HP process.
Figure 1. XRD patterns of Cu3Sb1−xSnxSe4−ySy prepared by the MA–HP process.
Materials 17 04859 g001
Figure 2. BSE–SEM micrographs of Cu3Sb1−xSnxSe4−ySy.
Figure 2. BSE–SEM micrographs of Cu3Sb1−xSnxSe4−ySy.
Materials 17 04859 g002
Figure 3. Elemental analysis for the Cu3Sb0.98Sn0.02Se3.50S0.50 specimen.
Figure 3. Elemental analysis for the Cu3Sb0.98Sn0.02Se3.50S0.50 specimen.
Materials 17 04859 g003
Figure 4. Electrical conductivity of Cu3Sb1−xSnxSe4−ySy.
Figure 4. Electrical conductivity of Cu3Sb1−xSnxSe4−ySy.
Materials 17 04859 g004
Figure 5. Seebeck coefficient of Cu3Sb1−xSnxSe4−ySy.
Figure 5. Seebeck coefficient of Cu3Sb1−xSnxSe4−ySy.
Materials 17 04859 g005
Figure 6. Power factor of Cu3Sb1−xSnxSe4−ySy. For comparison, the data for Cu3SbSe4 was obtained from reference [20].
Figure 6. Power factor of Cu3Sb1−xSnxSe4−ySy. For comparison, the data for Cu3SbSe4 was obtained from reference [20].
Materials 17 04859 g006
Figure 7. Thermal conductivity of Cu3Sb1−xSnxSe4−ySy.
Figure 7. Thermal conductivity of Cu3Sb1−xSnxSe4−ySy.
Materials 17 04859 g007
Figure 8. Separation of (a) electronic thermal conductivity and (b) lattice thermal conductivity for Cu3Sb1−xSnxSe4−ySy.
Figure 8. Separation of (a) electronic thermal conductivity and (b) lattice thermal conductivity for Cu3Sb1−xSnxSe4−ySy.
Materials 17 04859 g008
Figure 9. Dimensionless figure of merit for Cu3Sb1-xSnxSe4-ySy. For comparison, the data for Cu3SbSe4 was obtained from reference [20].
Figure 9. Dimensionless figure of merit for Cu3Sb1-xSnxSe4-ySy. For comparison, the data for Cu3SbSe4 was obtained from reference [20].
Materials 17 04859 g009
Table 1. Relative densities, lattice parameters, and charge-transport characteristics of Cu3Sb1−xSnxSe4−ySy.
Table 1. Relative densities, lattice parameters, and charge-transport characteristics of Cu3Sb1−xSnxSe4−ySy.
SpecimenRelative
Density
[%]
Lattice ParameterCarrier
Concentration
[1019 cm−3]
Mobility
[cm2V−1s−1]
Lorenz
Number
[10−8 V2K−2]
a [nm]c [nm]c/a
Cu3SbSe4 [20]98.10.56491.12431.99030.52501.57
Cu3Sb0.98Sn0.02Se3.75S0.2597.00.5636(1)1.1223(5)1.99121.172092.03
Cu3Sb0.96Sn0.04Se3.75S0.2599.70.5635(2)1.1221(8)1.99131.382852.00
Cu3Sb0.94Sn0.06Se3.75S0.2597.60.5636(2)1.1221(6)1.9910--1.95
Cu3Sb0.92Sn0.08Se3.75S0.2597.10.5636(2)1.1219(7)1.9906--1.83
Cu3Sb0.98Sn0.02Se3.50S0.5096.10.5618(2)1.1904(6)2.11891.081881.95
Cu3Sb0.96Sn0.04Se3.50S0.5095.50.5618(2)1.1891(5)2.11661.292771.92
Cu3Sb0.94Sn0.06Se3.50S0.5097.60.5619(3)1.1887(9)2.1155--1.91
Cu3Sb0.92Sn0.08Se3.50S0.5095.20.5619(2)1.1874(7)2.1132--1.81
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Hong, B.-K.; Kim, I.-H. Thermoelectric Characteristics of Permingeatite Compounds Double-Doped with Sn and S. Materials 2024, 17, 4859. https://doi.org/10.3390/ma17194859

AMA Style

Hong B-K, Kim I-H. Thermoelectric Characteristics of Permingeatite Compounds Double-Doped with Sn and S. Materials. 2024; 17(19):4859. https://doi.org/10.3390/ma17194859

Chicago/Turabian Style

Hong, Bong-Ki, and Il-Ho Kim. 2024. "Thermoelectric Characteristics of Permingeatite Compounds Double-Doped with Sn and S" Materials 17, no. 19: 4859. https://doi.org/10.3390/ma17194859

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