Mild Solvothermal Synthesis and Characterisation of Sulfur Rich Cu2-xS Nanocrystals with Varied Sulfur Precursors for Thermoelectrics

Chong Ka Shing1, Tam Teck Lip Dexter2
1Student, River Valley High School
2Institute of Materials Research and Engineering, A*STAR

ABSTRACT

Recent studies have proven that several Cu2-xS compounds display satisfactory figure of merit (ZT values, which evaluate the conversion efficiency of a thermoelectric material. In this study, sulfur rich Cu2-xS nanocrystals were synthesised with dithiooxamide or thiourea as sulfur precursors using a mild solvothermal method. Apart from investigating the two synthetic methods of Cu2-xS, this paper optimises the power factor (PF), where , for Cu2-xS by uncovering the optimum parameters affecting electrical conductivity and the Seebeck coefficient. X-Ray diffraction (XRD) affirmed the identity of the samples as covellite (CuS), while Energy-dispersive X-Ray spectroscopy (EDX) showed that dithiooxamide samples contained a considerable amount of impurities. Due to high purity, thiourea samples showed a high electrical conductivity of 579-1m-1 at room temperature. However, these samples have a moderate Seebeck coefficient and higher thermal conductivity due to their elevated hole concentration. Despite TU2, synthesised from thiourea, having the lowest Seebeck coefficient of 12.3μVK−1, this value is nearly three times greater than that of the single crystalline CuS which has a Seebeck coefficient of 5μVK−1, emphasising the significance of nanostructuring to increase the Seebeck coefficient. Nevertheless, the increment in the figure of merit owing to electrical conductivity is far greater than the deduction in the figure of merit by the Seebeck coefficient and thermal conductivity, especially at lower temperatures. CuS nanocrystals synthesised from the thiourea precursor show great potential as a room temperature thermoelectric device due to their unique temperature dependence of increasing power factor at lower temperatures. TU2 has the highest power factor of 51.4μWm-1K-2 at room temperature, and this power factor was found to be substantially higher than that of previously reported CuS by Narjis et al. with a power factor of 4.5μWm-1K-2 due to the solvothermal synthesis method adopted. This resulted in a ZT value of 0.00220 at room temperature, or 20% higher than that of CuS synthesized by Tarachand et al., which had a ZT of 0.00183. Chemically stable and non-toxic CuS nanocrystal samples synthesised offer a satisfactory thermoelectric performance with low synthetic cost and fill the research gap of a room temperature thermoelectric material.

Background and Purpose of Research

A temperature gradient induces the flow of charge carriers from the hotter to the cooler end, generating an electric field and creating a voltage difference directly proportional to the temperature difference.[1] Utilizing low grade heat from industrial processes or waste heat from electrical appliances is vital to reducing the growing carbon emissions of our society.[2] The Seebeck effect directly recovers electricity from thermal energy, thus receiving much attention as a promising green technology.[3] One notable thermoelectric material is Cu2-xS due to its low cost and non-toxicity (refer to Annex E).[4] Being a p-type semiconductor, Cu vacancies act as charge carriers, making Cu2-xS an excellent electrical conductor.[5] Moreover, Cu2-xS has variable stoichiometric compositions, ranging from sulfur rich CuS to copper rich Cu2S.[6] As a result, it has a Cu:S ratio dependent hole concentration and various band gap values ranging from 1.2 to 2.0 eV, suitable for a wide range of applications.[7] The crystalline structure of Cu2-xS can be subdivided into groups based on the sulfur packing in the lattice (refer to Annex A). Reported Cu2-xS can be further classified into covellite (CuS), yarrowite (Cu1.12S), spionkopite (Cu1.40S), anilite (Cu1.75S), digenite (Cu1.80S), djurleite (Cu1.97S) and chalcocite (Cu2S).[8]

Several Cu2-xS compounds produced in recent studies displayed satisfactory figure of merit (ZT values[9], which assess the suitability of a thermoelectric material. , where α is the Seebeck coefficient, is the electrical conductivity, T is the absolute temperature and κ is the thermal conductivity.[10] Most notably, Cu1.97S can display a maximum ZT value of 1.5 at 900K, which represents the highest reported value for p-type copper semiconductors.[11] However, there exists limited reports regarding thermoelectric properties of sulfur rich Cu2-xS because of the compounds’ inferior ZT merit values, with Tarachand et al. reporting the CuS with a highest ZT of 0.0187.[12] Nevertheless, recent nanocrystal engineering demonstrated increased Seebeck coefficient and decreased thermal conductivity due to the reduced dimensionality of nanocrystal quantum dot structures.[13] Hence, this paper studies methods to raise the Seebeck coefficient and electrical conductivity and uncover the potential of CuS through the synthesis of nanocrystals with strong interface phonon-scattering and charge carrier filtering abilities. Despite the current extensive study into Cu2-xS crystals, prior research has not attempted to compare the trends and relationships between the amount of sulfur source and types of sulfur precursors and their corresponding ZT at different temperatures using solvothermal synthesis. Investigating trends of different sulfur rich Cu2-xS nanocrystals using different synthetic conditions explores their potential and lays the foundation for future studies.

Different methods for the synthesis of Cu2-xS with varying oxidation states of Cu have been reported. These include the melting-quenching, solvothermal and microwave assisted synthesis methods.[14] However, this paper focuses solely on the economically efficient solvothermal synthesis of Cu2-xS using CuBr2 with two different sulfur-containing precursors under different conditions: thiourea (TU, SC(NH₂)₂) [15] and dithiooxamide (DTO, NH2(CS)2NH2).[16] Several studies have confirmed that thiourea provides a high reducing power to form a Cu[(NH2)2CS]2+ complex as an intermediate (Annex B) and likely contains less copper (II) oxide impurities in its final product.[17] On the other hand, dithiooxamide forms a stable and easily synthesisable Cu-DTO complex intermediate, hence lowering the activation energy of forming CuS. Both synthetic conditions are predicted to form sulfur rich Cu2-xS nanocrystals.[18] The two reactions consist of sulfur precursors that form a stable intermediate, eliminating the need for a capping ligand. The synthetic media used is N,N-dimethylformamide (DMF), an inexpensive and common solvent for inorganic synthesis.[19] One significant synthetic challenge is the controlling of the stoichiometric ratio of copper and sulfur due to the absence of a detailed mechanism to describe their formation.[20] Apart from investigating the two synthetic methods of Cu2-xS, this paper optimises the power factor ( ) for Cu2-xS by uncovering the optimum parameters affecting electrical conductivity and Seebeck effect through evaluating trends related to choosing between different Cu : S stoichiometric ratios and type of sulfur precursor.

Methodology

Different stoichiometric ratios (refer to Annex C) of copper (II) bromide, thiourea and dithiooxamide were dissolved in DMF and subjected to an oil bath of 160 in an inert atmosphere of nitrogen gas for 24 hours. Stoichiometric ratios were adjusted to produce different phase compositions and morphologies of sulfur rich Cu2-xS.[20] This solvothermal process allows for the creation of grain boundaries among Cu2-xS nanocrystals that had been proven to be effective in phonon scattering and the enriching of defect structures, hence the potential to increase both the Seebeck coefficient and electrical conductivity.[21] Samples were purified through washing them with DMF twice and methanol once after centrifuging at 4000 rpm for five minutes. They were dried in a vacuum oven of 60 overnight and ground into fine powders before pressing into a pellet. Samples were named with their sulfur precursor as the prefix and the copper to precursor concentration ratio as the suffix (refer to Annex C).

Pellets were tested for electrical conductivity (σ) and the Seebeck coefficient (α) using ZEM-3 while hall measurements were taken using BioRad HL5500. Energy-dispersive X-Ray spectroscopy (EDX) and X-Ray Diffraction (XRD) were carried out using FESEM 7600F and Bruker D8 Advance, respectively.

Results and Discussion

Synthesis of CuS nanocrystals results in the formation of a dark precipitate upon mixing copper (II) bromide and all concentrations of thiourea and dithiooxamide. XRD patterns of the products were recorded using Cu Kα radiation (λ=1.5418Å) in the 2θ range of 10°-85° and plotted in Figure 1. The patterns are in good agreement with CuS of hexagonal structure with cell parameters of a = 3.792Å and c = 16.34Å (JCPDS: 06-0464).

The (006) diffraction peak has a much stronger intensity than the (102) and (103) peaks, suggesting a preferential orientation of the crystals along the (006) plane.[22] This reveals the possibility to obtain anisotropic conductivity to be advantageous for thermoelectrics.[23] Average crystallite sizes of each sample were then calculated and recorded in Table 1 using the Scherrer equation, , where D is crystallite size, K is shape factor (0.9), λ is wavelength of Cu Kα emission, θ is Braggs angle corresponding to the maximum of the diffraction peak and β is full width at half maximum (FWHM) of diffraction peaks.[24] DTO samples have many outliers with massive crystallite sizes due to the presence of impurities

TABLE 1: Average crystallite size calculated from XRD data pattern
Average crystallite Size/nm DTO1 DTO2 TU1 TU1.5 TU2 TU4
35.86 29.05 28.09 19.24 29.42 22.82

The chemical compositions of all samples were evaluated by Energy-dispersive X-Ray spectroscopy. Table 2 revealed that all samples give a Cu : S stoichiometric ratio of about 1:1, which reaffirms the identity of covellite. Powders of DTO1 and DTO2 contain elevated amounts of oxygen and nitrogen due to the presence of unreacted dithiooxamide, thus restricting the flow of electrons and decreasing its conductance.

TABLE 2: EDX analysis of different samples
S.N. Sample Atomic % Total %
Cu S N O Br
1 TU1 33.9 34.7 11.7 13.8 5.9 100
2 TU1.5 48.6 38.5 0 12.9 0 100
3 TU2 54.4 45.6 0 0 0 100
4 TU4 53.4 46.6 0 0 0 100
5 DTO1 34.8 34.1 22.9 7.5 0.7 100
6 DTO2 29.8 31.8 31.8 6.6 0 100

Addition of thiourea shifts the equilibrium of the reaction from CuO to CuS, decreasing the atomic percentage of O as a result of its reducing power.

SEM images in Figure 2[1] illustrate that samples of thiourea precursors contain hexagonal flakes, which exhibit strong thermoelectric properties.[25] Clusters of nanoparticles are present in TU1 and TU1.5, providing the possibility that there exists another phase of CuS in them.


Measurements of electrical conductivity, Seebeck coefficient, power factor, thermal conductivity and figure of merit at different temperatures are recorded in Table 5 (refer to Annex F) and plotted in Figures 3.1, 3.2, 3.3, 3.4 and 3.5, respectively. The measured electrical conductivities of different samples are plotted in Figure 3.1, with values ranging from 7.94-1m-1 to 595.2-1m-1. Exceptional purity and hole concentrations in the order of 1019 cm-3 are believed to contribute to the remarkable electrical conducting properties of samples synthesised from thiourea (refer to Annex D). This hole concentration is obtained mainly due to the low to minimal band gap value of CuS, which is around 1.2 eV, as well as the presence of grain boundaries, which are shown to create abundant defect states.[26][27] Since electrical conductivity is given by:

where n, e and μ are the carrier concentration, charge and charge carrier mobility, respectively, the trends among the samples can be evaluated using their grain size and hole concentrations.[28] Small grain sizes induce a considerable number of grain boundaries which greatly reduce hole mobility by hole scattering.[29] Hence, electrical conductivity generally rises along with increased grain size, with the anomaly of TU1 due to its exceptionally poor hole concentration. Moreover, the presence of a sizeable amount of impurities reduces the electrical conductivity of TU1 further to a low of 97.1 -1m-1 at room temperature. Among all samples, TU2 has the highest ever reported Cu2-xS electrical conductivity of 595.2 -1m-1 at room temperature, a factor of 102 higher than that of TU1, labelling it as a suitable thermoelectric material to recover electricity even at room temperature.[30][31]

Seebeck coefficients are shown in Figure 3.2 in the order of μVK−1. The Seebeck coefficients of CuS nanocrystals are optimistic because of the existence of grain boundaries. Despite increasing interface phonon-scattering[32] and suppressing the phonon drag effect, preferential scattering of low-energy electrons increases the number of high-energy electrons in CuS, which makes the system more conducive to the movement of charges to lower energy states, increasing Seebeck coefficient. This effect is apparent in TU1.5, which has the most grain boundaries. Seebeck coefficient can be determined using the Mott formula[33], , where T is absolute temperature, n is the carrier concentration and m* is the effective mass of the carrier. Positive values of Seebeck coefficient indicate the charge carriers as holes.[34] Generally, p-type materials demonstrate a poor Seebeck coefficient due to the lower m* of holes relative to electrons.[35] However, because of low hole concentrations, DTO2 reached the highest Seebeck coefficient of 21.3 μVK−1 at 573K, while TU2 had the lowest Seebeck coefficient of 12.3 μVK−1. Regardless, this lowest value is nearly 3 times greater than that of a single crystalline CuS with α = 5 μVK−1, which emphasises the significance of nanostructuring in improving α. Nevertheless, the α2 of samples increases less than proportionately than decreases in temperatures higher than 300K, shifting the focus from optimising power factor at room temperature to synthesising samples with higher electrical conductivity in the range where Seebeck coefficient remains similar. The power factor of all samples is determined using . Despite having the lowest Seebeck coefficient, TU2 had the highest power factor of 51.4 μWm-1K-2 at room temperature and reached a peak of 55.7 μWm-1K-2 at 523K. These values are substantially higher than the previously reported power factors of CuS at 4.5 μWm-1K-2 and Cu1.8S at 1.93 μWm-1K-2 at room temperature.[30][31]

As shown in Figure 3.3, CuS is a potential candidate for a room temperature thermoelectric, as it has a synthetic cost that is relatively inexpensive and a unique power factor dependence on temperature. Owing to its metallic behavior of decreasing electrical conductivity and increasing Seebeck coefficient as temperature rises, there exists a reasonably high power factor of 51.4 μWm-1K-2 for CuS at low temperatures, as shown by sample TU2. DTO1 and DTO2 had comparatively lower power factors of 1.33 μWm-1K-2 and 2.82 μWm-1K-2, respectively, despite their higher Seebeck coefficients. Hence, having pure samples is an indispensable requirement for obtaining the optimum power factor.

Thermal conductivity results from two contributions. , where κL is the lattice thermal conductivity, and κC is the carrier thermal conductivity. Due to no commercially available device to test for parallel thermal conductivity and the anisotropic properties of all samples as evident by their preferred orientation, κC and κL were obtained from existing literature and calculations. κC was determined using the equation ,[34] where L is the Lorenz number obtained by the following equation to minimize deviation from experimental thermal conductivity values:[36] , where L is in 10−8WΩK−2 and in µVK-1. κL values were taken from Tarachand et al., who synthesised similar CuS nanocrystals with analogous lattice thermal conductivity[37] and extrapolated to the desired range. Thermal conductivity was then determined by the sum of κC and κL. Samples with a high conductivity allow for the transporting of heat with holes and reveal a greater dependence on κC than on κL, with TU2 having the highest thermal conductivity of 7.10 Wm-1K-1 at room temperature. Hence, the key to obtain the optimal power factor is to strive for a good balance among thermal conductivity, electrical conductivity and Seebeck coefficient by varying hole concentration.[38] This can be engineered by adjusting concentrations of sulfur precursors, which demonstrated a positive relationship with hole concentration in samples.

Thermoelectric figure of merit values were calculated for all samples across a range of temperatures. TU2 had the highest figure of merit of 0.00220 at room temperature, while TU4 had the maximum ZT of 0.00409 at 564K. While DTO samples have a lower thermal conductivity due to limited carrier concentrations, they had an extremely low electrical conductivity resulting in DTO1 having a low ZT of 0.000146 at room temperature. Figure 3.5 supports the hypothesis that power factor is a good comparison of thermoelectric efficiency, as it follows the trend of the figure of merit. Thiourea is evident to be a better sulfur precursor for future exploration of Cu2-xS thermoelectrics.

Conclusion, Limitations and Suggestions for Further Research

Both synthesis by thiourea or dithiooxamide sulfur precursors result in the formation of sulfur rich Cu2-xS. Synthesis by dithiooxamide led to the presence of a large amount of impurities, while synthesis with high thiourea concentration produced comparatively pure covellite. Due to the presence of impurities, CuS prepared with dithiooxamide had low electrical conductivity, and hence an unsatisfactory power factor. Alternatively, despite having the lowest Seebeck coefficient, TU2 has a ZT of 0.00220 and power factor of 51.4 μWm-1K-2 at room temperature, which exceed those of CuS reported in Tarachand et al with a ZT of 0.00187 and in Najis et al. with a power factor of 4.5 μWm-1K-2. Thus, thiourea is a better candidate for the solvothermal synthesis of pure CuS than dithiooxamide, which is also more expensive. A Cu:S stoichiometric ratio of 1:2 should be adopted as evident by the superior power factor of TU2.

Regardless of a relatively low maximum figure of merit, chemically stable CuS nanocrystals overcome the lapse in knowledge of room temperature thermoelectrics by offering satisfactory thermoelectric performance with a low synthetic cost coupled with their non-toxic nature. Despite the lack of thermal conductivity measurements, the lattice and carrier thermal conductivity of samples were determined from existing literature and theoretical equations, respectively. These predicted values allow for a rough comparison of its ZT with other thermoelectric materials despite its limitations. Furthermore, there also exists a possibility in thiourea samples to induce anisotropic conductivity to be beneficial for thermoelectric devices due to the presence of preferred orientation. In addition, there is an absence of mechanisms describing the formation of different grain sizes. Further exploration into these factors will allow for the optimisation of grain boundaries, and thus the optimising of the ZT of CuS. Since α2/κ increases less proportionately than electrical conductivity decreases at room temperature, this study suggests that the basis of synthesising room temperature thermoelectric material in future studies is to enhance electrical conductivity further in the range where there are little fluctuations in the Seebeck coefficient. This can be achieved through tuning grain size to look for the optimal point of power factor at room temperature. Further research conducted on CuS nanocrystals could possibly uncover more of their potential as a room temperature thermoelectric due to their optimised figure of merit at room temperature.

Reference List

  1. Ge, Z. H., Zhao, L. D., Wu, D., Liu, X., Zhang, B. P., Li, J. F., & He, J. “Low-cost, abundant binary sulfides as promising thermoelectric materials.” Materials Today, 19(4) (2016): 227-239.
  2. Chen, G, M S. Dresselhaus, G Dresselhaus and J P. Fleurial. “Recent developments in thermoelectric materials.” T. Int. Mater. Rev. 48 (2003): 45–66.
  3. Dresselhaus, M. S.New directions for low-dimensional thermoelectric materials.: Adv. Mater.19 (2007): 1043–1053.
  4. Han, J., Zhou, Y., Tian, Y., Huang, Z., Wang, X., Zhong, J., … & Tang, J. “Hydrazine processed Cu 2 SnS 3 thin film and their application for photovoltaic devices.” Frontiers of Optoelectronics, 7(1) (2014): 37-45
  5. Rabinal, M. H., & Mulla, R. “Copper Sulfides: Earth Abundant and Low Cost Thermoelectric Materials.” Energy Technology (2018).
  6. Grozdanov, I., & Najdoski, M. “Optical and electrical properties of copper sulfide films of variable composition.” Journal of Solid State Chemistry, 114(2) (1995): 469-475.
  7. Zhang, Y., Wang, Y., Xi, L., Qiu, R., Shi, X., Zhang, P., & Zhang, W. “Electronic structure of antifluorite Cu2X (X = S, Se, Te) within the modified Becke-Johnson potential plus an on-site Coulomb U.” The Journal of Chemical Physics, 140(7) (2014): 074702.
  8. Sun, S., Li, P., Liang, S., & Yang, Z. “Diversified copper sulfide (Cu2−xS) micro-/nanostructures: A comprehensive review on synthesis, modifications and applications.” Nanoscale, 9(32) (2017): 11357-11404.
  9. Zhao, L., Wang, X., Fei, F. Y., Wang, J., Cheng, Z., Dou, S., … & Snyder, G. J. “High thermoelectric and mechanical performance in highly dense Cu 2− x S bulks prepared by a melt-solidification technique.” Journal of Materials Chemistry A, 3(18) (2015): 9432-9437.
  10. Goldsmid, H. J. “The thermoelectric figure of merit”. The Physics of Thermoelectric Energy Conversion (2017)
  11. He, Y., Day, T., Zhang, T., Liu, H., Shi, X., Chen, L., & Snyder, G. J. “High Thermoelectric Performance in Non-Toxic Earth-Abundant Copper Sulfide.” Advanced Materials, 26(23) (2014): 3974-3978.
  12. Tarachand, Hussain, S., Lalla, N. P., Kuo, Y. K., Lakhani, A., Sathe, V. G., Deshpande, U., & Okram, G. S. “Thermoelectric properties of Ag-doped CuS nanocomposites synthesized by a facile polyol method.” Physical Chemistry Chemical Physics, 20(8) (2018): 5926-5935.
  13. Martin, J., Wang, L., Chen, L., & Nolas, G. S. “Enhanced Seebeck coefficient through energy-barrier scattering in PbTe nanocomposites.” Physical review B, 79(11) (2009): 115311.
  14. Liao, X. H., Chen, N. Y., Xu, S., Yang, S. B., & Zhu, J. J. “A microwave assisted heating method for the preparation of copper sulfide nanorods.” Journal of Crystal Growth, 252(4) (2003): 593-598.
  15. Rahmani, A., Rahmani, H., & Zonouzi, A. “Synthesis of copper sulfides with different morphologies in DMF and water: Catalytic activity for methyl orange reduction.” Materials Research Express, 4(12) (2017): 125024.
  16. Roy, P., & Srivastava, S. K. “Hydrothermal Growth of CuS Nanowires from Cu−Dithiooxamide, a Novel Single-Source Precursor.” Crystal Growth & Design, 6(8) (2006): 1921-1926.
  17. Madarász, J., Bombicz, P., Okuya, M., & Kaneko, S. “Thermal decomposition of thiourea complexes of Cu (I), Zn (II), and Sn (II) chlorides as precursors for the spray pyrolysis deposition of sulfide thin films.” Solid State Ionics, 141 (2001): 439-446.
  18. Zhao, L., Tao, F., Quan, Z., Zhou, X., Yuan, Y., & Hu, J. “Bubble template synthesis of copper sulfide hollow spheres and their applications in lithium ion battery.” Materials Letters, 68 (2012): 28-31.
  19. Adler, A D., F R. Longo, F Kampas and J Kim. \”On the preparation of metalloporphyrins.\” Journal of Inorganic and Nuclear Chemistry 32 (1970): 2443-2445.
  20. Kumar, P., Gusain, M., & Nagarajan, R. “Synthesis of Cu1. 8S and CuS from copper-thiourea containing precursors; anionic (Cl−, NO3−, SO42−) influence on the product stoichiometry.” Inorganic chemistry, 50(7) (2011): 3065-3070
  21. Zhu, G. H., Lee, H., Lan, Y. C., Wang, X. W., Joshi, G., Wang, D. Z., … & Dresselhaus, M. S. “Increased phonon scattering by nanograins and point defects in nanostructured silicon with a low concentration of germanium.” Physical review letters102(19) (2009): 196803.
  22. Chang, J Y. and C Y. Cheng. \”Facile one-pot synthesis of copper sulfide–metal chalcogenide anisotropic heteronanostructures in a noncoordinating solvent.\” Chemical Communications 47 (2011): 9089-9091.
  23. Liang, W., & Whangbo, M. H. “Conductivity anisotropy and structural phase transition in covellite CuS.” Solid state communications, 85(5) (1993): 405-408.
  24. Holzwarth, U., & Gibson, N. “The Scherrer equation versus the \’Debye-Scherrer equation\’.” Nature nanotechnology, 6(9) (2011): 534.
  25. Du, W., Qian, X., Ma, X., Gong, Q., Cao, H., & Yin, J. “Shape‐Controlled Synthesis and Self‐Assembly of Hexagonal Covellite (CuS) Nanoplatelets.” Chemistry–A European Journal, 13(11) (2007): 3241-3247
  26. Itoh, K., Kuzuya, T., & Sumiyama, K. “Morphology and composition-controls of cuxs nanocrystals using alkyl-amine ligands.” Materials transactions47(8) (2006):1953-1956
  27. Feng, W., Fang, Z., Wang, B., Zhang, L., Zhang, Y., Yang, Y., … & Liu, P. “Grain boundary engineering in organic–inorganic hybrid semiconductor ZnS (en) 0.5 for visible-light photocatalytic hydrogen production.” Journal of Materials Chemistry A5(4) (2017): 1387-1393
  28. Nolas, G. S., Sharp, J., & Goldsmid, J. “Thermoelectrics: basic principles and new materials developments” (Vol. 45). Springer Science & Business Media (2013).
  29. Yamada, T., Makino, H., Yamamoto, N., & Yamamoto, T. “Ingrain and grain boundary scattering effects on electron mobility of transparent conducting polycrystalline Ga-doped ZnO films.” Journal of Applied Physics107(12) (2010): 123534.
  30. Narjis, A., Outzourhit, A., Aberkouks, A., El Hasnaoui, M., & Nkhaili, L. “Structural and thermoelectric properties of copper sulphide powders.” (2018).
  31. Narjis, A., Outzourhit, A., Aberkouks, A., El Hasnaoui, M., & Nkhaili, L. “Spectroscopic study and thermoelectric properties of a mixed phase copper sulfide lamellas.” Journal of Alloys and Compounds762 (2018): 46-48.
  32. Lan, Y., Minnich, A. J., Chen, G., & Ren, Z. “Enhancement of thermoelectric figure‐of‐merit by a bulk nanostructuring approach.” Advanced Functional Materials20(3) (2010): 357-376
  33. Mott, N. F. “Conduction in glasses containing transition metal ions.” Journal of Non-Crystalline Solids1(1) (1968): 1-17.
  34. Snyder, G. J., & Toberer, E. S. “Complex thermoelectric materials.” In Materials For Sustainable Energy: A Collection of Peer-Reviewed Research and Review Articles from Nature Publishing Group (2011): 101-110.
  35. Hosono, H., Mishima, Y., Takezoe, H., & MacKenzie, K. J. (Eds.). “Nanomaterials: Research Towards Applications” (Vol. 161). Elsevier (2006)
  36. Kim, H. S., Gibbs, Z. M., Tang, Y., Wang, H., & Snyder, G. J. “Characterization of Lorenz number with Seebeck coefficient measurement.” APL materials3(4) (2015): 041506.
  37. Tritt, T. M. (Ed.). “Thermal conductivity: theory, properties, and applications.” Springer Science & Business Media (2005).
  38. Terasaki, I. “1.09 Thermal Conductivity and Thermoelectric Power of Semiconductors.” Comprehensive Semiconductor Science and Technology, Six-Volume Set, 1 (2011): 326.

Annex B
Formation of copper sulfide through the intermediate of a copper-thiourea complex.[15]
Annex C
TABLE 3: Different stoichiometric ratios of precursors added for samples

Sample Copper (II) bromide Dithiooxamide Thiourea Dimethylformamide
DTO1 1 1 0 10ml
DTO2 1 2 0 10ml
TU1 1 0 1 10ml
TU1.5 1 0 1.5 10ml
TU2 1 0 2 10ml
TU4 1 0 4 10ml

Annex D
TABLE 4: Carrier concentrations of samples

Carrier Concentration (cm-3) DTO1 DTO2 TU1 TU1.5 TU2 TU4
1.6×1017 3.7×1017 1.5×1018 1.6×1018 5.7x 1018 9.3×1018

Annex E


Annex F
TABLE 5: Different samples and their electrical conductivity, Seebeck coefficient, power factor, thermal conductivity and figure of merit at different temperatures

Sample Temperature/K Electrical Conductivity/-1m-1 Seebeck Coefficient/ μVK−1 Power Factor/μWm-1K-2 Thermal Conductivity/ Wm-1K-1 Figure of Merit*103
DTO1 305 10.582 11.2 1.33 2.785 0.1459
322 10.384 11.7 1.41 2.818 0.1613
369 9.803 12.9 1.63 2.902 0.2074
418 9.433 14.8 2.08 2.971 0.2927
467 9.090 16.4 2.44 3.029 0.3762
515 8.474 17.5 2.59 3.078 0.4334
564 7.936 20.7 3.42 3.105 0.6213
DTO2 324 16.469 13.1 2.82 2.866 0.3187
370 15.406 14.2 3.12 2.951 0.3912
419 14.487 15.8 3.62 3.021 0.5021
468 13.946 17.6 4.33 3.082 0.6575
517 13.312 19.3 4.96 3.137 0.8176
566 12.673 21.3 5.72 3.167 1.0222
TU1 305 97.087 9.79 9.33 3.422 0.8302
324 95.238 10.6 10.7 3.482 0.9949
369 102.040 10.5 11.2 3.724 1.1099
382 103.305 11.2 13 3.827 1.2975
466 65.359 12.5 10.2 3.658 1.2993
515 134.770 12.4 20.8 4.640 2.3087
563 273.224 11.7 37.4 6.698 3.1436
TU1.5 305 8.620 10.1 0.872 2.771 0.0960
323 8.403 9.98 0.836 2.804 0.0962
370 23.419 11.3 2.97 3.023 0.3633
419 31.055 11.6 4.22 3.190 0.5543
468 19.685 13.6 3.65 3.148 0.5426
516 28.248 16.8 7.95 3.320 1.2357
565 158.730 14.6 33.7 5.136 3.7073
TU2 304 595.238 9.28 51.4 7.095 2.2041
322 571.238 9.31 49.6 7.197 2.2203
369 515.463 9.87 50.1 7.419 2.4941
418 460.829 10.1 47.2 7.532 2.6194
467 413.223 11.4 53.3 7.572 3.2874
516 378.787 12.1 55.7 7.668 3.7483
564 348.432 12.3 52.3 7.715 3.8233
TU4 304 375.939 9.05 30.8 5.477 1.7087
322 359.712 8.91 28.6 5.550 1.6605
369 327.868 9.36 28.8 5.749 1.8504
418 299.401 10.1 30.6 5.901 2.1674
467 272.479 11.5 36.3 5.989 2.8305
515 248.756 12.8 40.6 6.044 3.4595
564 233.644 13.8 44.6 6.147 4.0924
  1. SEM images of samples synthesized from dithiooxamide are in an amorphous state, hence not shown.

About the Author

 Ka Shing is a Singaporean student from River Valley High School who is interested in Materials Science since 15 years old. He participates in various research attachments to gain insights of how a scientist works.

Leave a Comment

Your email address will not be published. Required fields are marked *