Next Article in Journal
Reutilization of Reclaimed Asphalt Binder via Co-Pyrolysis with Rice Husk: Thermal Degradation Behaviors and Kinetic Analysis
Next Article in Special Issue
Crucial Role of Ni Point Defects and Sb Doping for Tailoring the Thermoelectric Properties of ZrNiSn Half-Heusler Alloy: An Ab Initio Study
Previous Article in Journal
Exploring the Utilization of PHC Pile Waste Concrete as Filler in Asphalt Mastics
Previous Article in Special Issue
Innovative Design of Bismuth-Telluride-Based Thermoelectric Transistors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Largely Enhanced Thermoelectric Power Factor of Flexible Cu2−xS Film by Doping Mn

Key Laboratory of Advanced Civil Engineering Materials of Ministry of Education, Shanghai Key Laboratory of Development and Application for Metal-Functional Materials, School of Materials Science & Engineering, Tongji University, Shanghai 201804, China
*
Author to whom correspondence should be addressed.
Materials 2023, 16(22), 7159; https://doi.org/10.3390/ma16227159
Submission received: 23 October 2023 / Revised: 7 November 2023 / Accepted: 11 November 2023 / Published: 14 November 2023

Abstract

:
Copper-sulfide-based materials have attracted noteworthy attention as thermoelectric materials due to rich elemental reserves, non-toxicity, low thermal conductivity, and adjustable electrical properties. However, research on the flexible thermoelectrics of copper sulfide has not yet been reported. In this work, we developed a facile method to prepare flexible Mn-doped Cu2−xS films on nylon membranes. First, nano to submicron powders with nominal compositions of Cu2−xMnyS (y = 0, 0.01, 0.03, 0.05, 0.07) were synthesized by a hydrothermal method. Then, the powders were vacuum-filtrated on nylon membranes and finally hot-pressed. Phase composition and microstructure analysis revealed that the films contained both Cu2S and Cu1.96S, and the size of the grains was ~20–300 nm. By Mn doping, there was an increase in carrier concentration and mobility, and ultimately, the electrical properties of Cu2−xS were improved. Eventually, the Cu2−xMn0.05S film showed a maximum power factor of 113.3 μW m−1 K−2 and good flexibility at room temperature. Moreover, an assembled four-leg flexible thermoelectric generator produced a maximum power of 249.48 nW (corresponding power density ~1.23 W m−2) at a temperature difference of 30.1 K, and had good potential for powering low-power-consumption wearable electronics.

1. Introduction

Thermoelectric (TE) materials can directly convert heat into electricity and vice versa. This has attracted great attention for power generation and cooling [1]. With the growing demand for electrical appliances, there is an urge to produce renewable sources using flexible, eco-friendly, and economical materials. Flexible TE films (f-TEFs), which can fulfill this need, can generate electricity by utilizing the temperature difference (ΔT) between the human body and the environment [2]. The TE performance of a material is determined by the dimensionless figure of merit ZT, defined as ZT = α2σT/κ, where α is the Seebeck coefficient, σ is the electrical conductivity, κ is the thermal conductivity, T is the absolute temperature, and α2σ is called power factor (PF) that is related to electrical properties [3].
In recent years, substantial efforts have been mainly devoted to free-standing films [4,5,6] and films on flexible substrates [7,8,9,10,11,12]. For free-standing films, conductive polymers and their based composites have been widely studied due to their excellent flexibility and low thermal conductivity [13,14]. However, their ZT values still cannot be compared to those of inorganic TE materials. Inorganic films on flexible substrates can combine the flexibility of the substrate (such as polyimide, nylon, or paper) with the high ZT of inorganic TE materials, effectively balancing flexibility and TE performance [15]. For instance, a flexible p-type Bi0.4Sb1.6Te3/Te composite film prepared on a Kapton surface by screen printing combined with pressure-less sintering (450 °C, 45 min) achieved a reasonably high PF value of ~3000 μW m−1 K−2 (ZT ~ 1) at 300 K [9]. It showed a 3% increase in resistance after 1000 times bending around a 10 mm-radius rod. In 2019, our group developed a facile method to fabricate flexible Ag2Se films on nylon membranes [7]: Se nanowires (NWs) were synthesized by a wet chemical method and used as templates to prepare Ag2Se NWs, and Ag2Se film was formed on porous nylon membranes by vacuum filtration and then hot-pressing (HP). The Ag2Se films exhibited excellent flexibility and a PF of ~987 μW m −1 K−2 (ZT ~ 0.5) at 300 K. Since then, we have worked to improve the PF [16,17,18]. For example, polypyrrole (PPy) was in situ polymerized to fabricate Ag2Se/Se/PPy composite films on nylon membranes [17], and a PF of ~2240 μW m−1 K−2 at 300 K was achieved. Recently, Lei et al. [19] prepared Ag2Se films by immersing Ag films sputtered on polyimide substrates into Se/Na2S solution for selenization, and the PF at room temperature (RT) was ~2590 ± 414 μW m−1 K−2. Despite the excellent TE properties of Bi2Te3 and Ag2Se at RT, the elements Te and Se are not abundant and are toxic. So, for practical TE applications, it was necessary to search for low-cost inorganic materials without toxic elements.
Copper sulfide (Cu2−xS, 0 ≤ x ≤ 0.25), a typical liquid-like semiconductor [20], which has low lattice thermal conductivity [21], has been considered to be the candidate material to decouple electrical and thermal properties. Meanwhile, the elements Cu and S are abundant, cheap, and nontoxic; thus, Cu2−xS is an economical and environmentally friendly TE material [22,23]. It forms a series of compounds ranging from copper-rich to copper-deficient, such as Cu2S, Cu1.96S, Cu1.92S, and Cu1.8S, whose crystal structures and TE performances vary with copper content. Typically, Cu2S possesses a comparatively high α of ~300 μV K−1 and poor σ (≤10 S cm−1) at RT [24]. The crystal structure of Cu2S undergoes complex change with increasing temperature: it is in a monoclinic-chalcocite phase at RT, which transforms into a hexagonal phase above 370 K and then into a cubic phase near 709 K [9,25]. The current research on copper-sulfide bulks mainly focuses on the medium and high-temperature regions. For example, Tang et al. [26] introduced a 3D graphene heterointerface into the Cu2−xS matrix by mechanical alloying and spark plasma sintering (SPS) and obtained a high PF of 1197 μW m−1 K−2 (ZT ~ 1.56) at 873 K. And Cu2S bulk incorporated Ag nanoparticles prepared by a hydrothermal method realized a high PF of 1698 μW m−1 K−2 (ZT ~ 1.4) at 773 K [27]. Recently, Li et al. [28] synthesized Cu2S-Cu1.96S phase junctions by retaining surface 1-dodeca-nethiol (DDT) ligands and pushed the peak ZT value to 2.1 at 932 K. However, bulk materials are usually costly and not suitable for flexible TE devices [29,30].
To date, there has been little attempt at flexible copper-sulfide-based films. For example, Cu2S/poly (3,4-ethylene dioxythiophene):poly (styrene sulfonate) (PEDOT:PSS) hybrid films were fabricated by screen printing on polyimide. And when the content ratio of Cu2S/PEDOT:PSS was 1:1.2, the prepared film showed a maximum PF of 20.60 µW m−1 K−1 at ~400 K [31]. Self-supporting and flexible Cu2S/PEDOT:PSS composite TE films were prepared by a vacuum filtration method, and the maximum PF was 56.15 μW m−1 K−2 at 393 K [32]. Both of these films had good flexibility, but the TE performances were low.
Recently, our group [33] reported a one-pot method for the synthesis of Ag2Se powders combined with vacuum filtration and HP to prepare flexible Ag2Se film on a nylon membrane, which possessed a PF of ~2042 μW m−1 K−2 at RT. This showed that flexible films can be prepared by the above method as long as the size of the powders is beyond the pore size of the nylon membrane. Herein, we designed a green and facile hydrothermal synthesis method to synthesize Cu2−xS powders without any corrosive chemicals and then prepared the Cu2−xS film on nylon membranes by vacuum filtration and then HP. As mentioned above, the σ of Cu2S is, unfortunately, low. There have been many strategies to improve the σ of TE materials, such as doping to tune carrier concentration (n) [27], introducing a second phase with high σ [23], and band engineering [34]. Copper ions in Cu2S are mobile ions, and doping positive ions with higher electronegativity at the copper site will increase the n by forming a stronger bond with sulfur and is a good strategy to optimize the electrical properties of Cu2S [35].
In this work, to improve the PF of the Cu2−xS film, we chose Mn as the dopant to enhance the σ while maintaining the α. Through the doping of Mn, an optimal film exhibited a PF of 113.3 μW m−1 K−2 at RT, and the output performance of an assembled flexible TE generator (f-TEG) was studied.

2. Experimental Section

2.1. Sample Synthesis

Mn-doped Cu2−xS powders were synthesized by a hydrothermal method, which was a modified method for Cu2−xS [36]. Typically, 3.6 mmol thiourea (Tu), 7.56 mmol CuCl2·2H2O, and MnCl2·4H2O were dissolved into 20 mL of deionized (DI) water. The nominal doping content of Mn in Cu2−xS was 0, 1, 3, 5, and 7% by molar ratio, marked as Cu2−xMnyS (y = 0, 0.01, 0.03, 0.05, and 0.07). Afterward, they were mixed and sonicated for 30 min and then transferred into a 100 mL Teflon-lined stainless-steel autoclave at 180 °C for 18 h with a heating rate of 2 °C min−1. After cooling naturally, the products were washed with DI water and ethanol 3 times. The corresponding films were prepared by vacuum-assisted filtration of the powder dispersions on porous nylon membranes, then dried at 60 °C in vacuum overnight, and, finally, hot-pressed at 270 °C and 1 MPa for 30 min. The schematic diagram of the preparation process of the Cu2−xMnyS film is shown in Figure S1 (see Supplementary Materials).

2.2. Assembly of the f-TEG

The Cu2−xMn0.05S film was cut into four strips of 2 cm × 0.5 cm. Two ends of each strip were coated with a thin layer of Au by evaporation to reduce contact resistance. Then, the four strips were pasted on a polyimide substrate at intervals of 0.5 cm. Finally, Ag paste (SPI#04998-AB) was used to connect the strips in series.

2.3. Characterization and Property Measurement

X-ray diffraction (XRD, Bruker D8 Advance, Cu Kα radiation, Bruker, Shanghai, China) was performed. Scanning electron microscopy (SEM, Nova NanoSEM 450, Thermo Fisher Scientific, Shanghai, China) and energy-dispersive X-ray spectroscopy (EDS) were used to observe the surface and cross-sectional morphology and composition. X-ray photoelectron spectroscopy (XPS, ESALAB 250Xi Spectrometer Microprobe, Thermo Fisher Scientific, Shanghai, China) was used to analyze the composition and valence states of surface elements. The binding energy was calibrated by setting the standard value of C1s to 284.8 eV. Powders scraped from the Cu2−xMn0.05S film were observed by transmission electron microscopy (TEM, JEM-2100F, JEOL, Shanghai, China). The films were cut into 1.3 cm × 0.3 cm strips to test the temperature dependence of σ and α with a TE test system (CTA Cryoall, Beijing Cryoall, Shanghai, China) under the protection of He. The Hall carrier concentration (nH) and mobility (μH) were measured using the Van der Pauw method with a Hall effect measurement system (HMS-7000, Ecopia, Shanghai, China). To test flexibility, σ was measured before and after bending around a 4 mm-radius rod to test flexibility.
The assembled f-TEG was connected with wires into a circuit, as shown in Figure S2 (see Supplementary Materials) for the output performance test [37,38]. The hot-end temperature was controlled by heating a copper block (T + ΔT) and the other end was put on an adiabatic foam acting as the cold side (T). The temperature at both ends was measured by two thermocouples. The output voltage and current were collected by adjusting the variable resistor box at a specific ΔT, which was varied by setting different heating temperatures.

3. Results and Discussion

Figure S3 (see Supplementary Materials) shows XRD patterns of the powders. The main phase of undoped copper sulfide can be indexed to tetragonal Cu2S (PDF No. 72-1071) with evident diffraction peaks at 2θ = 31.6°, 32.6°, 39.0°, 39.8°, 45.3°,46.0°, and 48.2°, which correspond to (110), (111), (104), (113), (200), (201), and (202) planes of Cu2S, respectively. There is no obvious shift of XRD peaks and no manganese-related compounds are detected. There are very weak diffraction peaks in Figure S3 when y = 0.07, which belong to digenite Cu1.8S (PDF No. 24-0061). According to [21,36], the related reactions are proposed as follows:
NH 2 CSNH 2 + 3 H 2 O   H 2 S + 2 N H 4 + + CO 3 2−
H 2 S + C u 2 + CuS + 2 H +
CuS Cu 2 - x S
Hydrogen sulfide comes from the decomposition of thiourea in the early stage of the reaction and reacts with copper ions to produce CuS [36], which subsequently transforms into Cu2−xS during the heating process [21]. Figure 1a depicts the XRD patterns for the Cu2−xMnyS films. It can be observed that the main diffraction peaks can be well indexed to monoclinic Cu2S (PDF No. 33-0490). In addition, two peaks at 2θ ~ 32.6° and 39.0° are detected, corresponding to (103) and (104) planes of tetragonal Cu1.96S (PDF No. 29-0578). Hence, the as-prepared films are composed of Cu2S with a small amount of Cu1.96S. When the nominal content of Mn is 7%, the XRD peaks for Cu2S broaden and the peaks for Cu1.96S become stronger.
Figure S4 (see Supplementary Materials) shows SEM images of Cu2−xMnyS (y = 0.01, 0.03, 0.05, and 0.07) powders, with most particles being ~20–300 nm. A representative SEM image of the Cu2−xMn0.05S film is shown in Figure 1b (the other films show similar morphology, see Supplementary Figure S5). The size of grains is <~300 nm. It can be seen from Figure 1b that the film is not very dense (as the sintering temperature was limited by the melting point of the nylon substrate). EDS results (Figure 1c–e) indicate that the elements of Cu, S, and Mn are homogeneously distributed in the Cu2−xMn0.05S film.
Figure 2a shows the XPS survey spectra of the Cu2−xMn0.05S film; the signals of Cu, S, Mn, and C are detected. Two strong peaks at 932.6 eV (Cu 2p3/2) and 952.4 eV (Cu 2p1/2) (see Figure 2b) correspond to Cu+ [39]. The weak split peaks at about 933.8 and 954.0 eV are attributed to Cu2+, and the two satellite peaks located at 944.0 and 962.5 eV also correspond to Cu2+ [27,35], which proves the existence of Cu1.96S [40,41]. We estimated the ratio of Cu+:Cu2+ to be about 4.69:1, which is close to the value estimated by semi-quantitative analysis from the XRD result (5.11:1). Compared with the spectra of Cu2−xMn0.05S powders (see Supplementary Figure S6), the ratio Cu+:Cu2+ becomes higher, indicating the partial conversion of Cu1.96S to Cu2S during HP. In Figure 2c, the characteristic peaks of S 2p3/2 and 2p1/2 are located at 161.6 and 162.8 eV, respectively, and the energy difference between them is about 1.2 eV, indicating the existence of S2− [42,43,44]. Besides, the small resolved peaks with binding energies of 641.2 and 649.2 eV are Mn 2p3/2 and 2p1/2, respectively (see Figure 2d). The peak at 641.2 eV of Mn 2p3/2 is consistent with the previously reported value of MnS, which is attributed to the Mn-S bond [45]. The XPS and EDS results demonstrate the successful incorporation of Mn into copper sulfide. Mn has an initial valence of +2 and no strong oxidant is present in the reaction, so it is considered that Mn exists in the samples in a divalent state. In addition, because Cu+ (0.096 nm) and Cu2+ (0.072 nm) coexist in the Cu2−xS, while the ionic radius of Mn2+ (0.080 nm) is between them, the XRD peaks in Figure S3 and Figure 1a have no obvious shift.
Figure 3a,b,d show the temperature dependence of TE parameters of the Cu2−xMnyS films. In the initial stage of temperature rise, the σ increases concomitantly, exhibiting a typical semiconductor behavior, and then it decreases when the temperature is near 340 K, which is lower than the transition temperature (Tt) of Cu2S (370 K). The leftward shift of the Tt is due to the presence of Cu1.96S, which undergoes a phase transition at 336 K [46]. The undoped Cu2−xS film possessed a high α of 271 μV K−1 and a low σ of 8 S cm−1 at RT. As the Mn doping amount increased, the σ increased, reaching ~70 S cm−1 when y = 0.07, nearly nine times as high as that of the undoped Cu2−xS film. Cu2−xS is a type of p-type semiconductor (see hereinafter). Manganese was doped into copper sulfide in the form of divalent ions, according to the XPS analysis, which should provide additional electrons in the Cu2−xMnyS films to act as a donor. However, the σ at RT exhibited an opposite trend. This phenomenon is consistent with the change in σ of Mn- and Sn-alloyed Cu2S bulk obtained by the melting method [39]. Besides, through molecular orbital theory analysis, Wang et al. [47] revealed that the 3d orbital energies of Mn were similar to those of S 3p orbitals, causing the S 3p orbital to move away from the Cu-S bond, thereby weakening the Cu-S bond. Hence, the introduction of Mn weakens the Cu-S chemical bond, resulting in the formation of more Cu vacancies in the crystal lattice [39,48,49], namely, increasing the hole concentrations, which agrees with the Hall test result: At RT, the nH increased from 1.22 × 1020 for the Cu2−xS and to 3.80 × 1020 cm−3 for the Cu2−xMn0.05S. In addition, the μH also increased from 2.91 for the Cu2−xS to 4.37 cm2 V−1 s−1 for the Cu2−xMn0.05S. Since the σ is proportional to the n and mobility (μ), defined as:
σ = n q μ  
where q is the electron charge, the σ of the Cu2−xMn0.05S film was ultimately improved.
The α increased with rising temperature (Figure 3b). And the positive α values also indicate that holes were the dominant charge carriers. As the Mn doping amount increased, the α decreased from 271 to 109.9 μV K−1 at RT, which is opposite to the trend of σ with the amount of doping. The α is proportional to n−2/3, expressed as follows:
α = 8 π 2 k B 2 3 e h 2 m * T ( π 3 n ) 2 / 3
where kB is the Boltzmann constant, h is the Planck constant, and m* is the effective mass of carriers. Therefore, an increase in n will lead to a decrease in α. Figure 3c gives the α as a function of n based on a single parabolic band (SPB) model and assuming a dominated scattering by acoustic phonons, which is called the Pisarenko curve. The m* of Cu2−xS in this work (the red line) is 6.5 me. It is much higher than that of the reported data for Cu2S (0.5 me) bulk [50]. In ref. [50], the α of the Cu2S is 280 μV K−1 with nH ~ 2.3 × 1018 cm−3, whereas the α of the present Cu2−xS is 271 μV K−1 with nH ~ 1.22 × 1020 cm−3. According to Equation (5), the Cu2−xS will have a much higher m*. Therefore, the Cu2−xS with a large m* is due to it having a high nH. In the present case, the Cu2−xS consisted of two phases: Cu2S and Cu1.96S. The Cu1.96S, which possesses more copper vacancies, would have increased the nH [28,52]. And manganese doping (y = 0.05) effectively reduced the m* to 4.2 me, which is close to the value of Cu2S1−xTex [51] (m* = 4.5 me). Consequently, a higher n and reduced m* adjusted the α to a moderate value. Additionally, the decreased m* also had an impact on μ according to the equation:
μ = q τ m *
where τ is the carrier relaxation time. The decrease in m* caused the enhancement of μ after doping. As a result, the Cu2−xMnyS (y = 0.05) film possessed a higher PF of ~113.3 μW m−1 K−2 at RT (~152.1 μW m−1 K−2 at 413 K), nearly twice that of the Cu2−xS film (58.5 μW m−1 K−2). Table 1 shows a comparison of TE performance between this work and the reported copper-sulfide-based bulks and flexible films. The PF value of the Cu2−xMn0.05S film was superior to those of previously reported Cu2S-based flexible films and most bulks. However, it was lower than the PF value in ref. [21]. This may have been due to the fact that the Cu2−xMn0.05S film was relatively porous compared to the SPS sintered bulk.
To further understand the internal microstructure, TEM was applied for observation of the grains scraped from the optimal film, and the result is shown in Figure 4. The grain sizes were ~20–300 nm. In the HRTEM image (Figure 4d), the measured lattice spacings of 0.305, 0.266, 0.315, and 0.194 nm are in good agreement with those of the (132), (042), (−114), and (630) planes of monoclinic Cu2S, respectively. And there are edge dislocations in the grains, which may be the lattice distortion caused by Mn doping. In addition, Cu1.96S is observed. Figure 4g shows an FFT image of the grain circled in a yellow dotted line, and the interplanar spacing is about 0.170 nm, corresponding to that of the (212) plane of Cu1.96S [52]. Figure 4e,f,h,i demonstrate that the lattice spacings are 0.188 and 0.318 nm, which correspond to the (−136) and (222) planes of Cu2S. Thus, the TEM results also confirm that the Cu2−xMn0.05S film contained two the phases, Cu2S and Cu1.96S.
Flexibility is also a key factor in the practical application of f-TEFs. The thickness of the Cu2−xMn0.05S film was approximately 8.96 μm (see Supplementary Figure S7). And Figure 5a shows the corresponding flexibility test result: the σ maintained at 94.4%, 93.3%, and 89.6% of the original after being bent around a 4 mm-radius rod 500, 1000, and 1500 times, respectively. This is better than the flexibility of the reported Cu2S/PEDOT:PSS composite films [32] (the resistance rose 10% after bending 1000 cycles under a bending radius of 4 mm). The main reasons for the good flexibility are as follows: (1) the nylon membrane possessed excellent flexibility and (2) the combination between the porous film and the nylon membrane was good (see Figure 5b).
Figure 6a gives the variation of open-circuit voltage (Voc) with ΔT. When the ΔT values were 10.3, 21.3, and 30.1 K, the Voc values of the f-TEG were 5.74, 11.86, and 16.65 mV, respectively, which is close to the values (see Figure 6c) calculated by the equation: V o c = α · N · T (N is the number of f-TEG legs). Figure 6b shows the output properties of the f-TEG by adjusting the load resistance (Rload) under different ΔT. Output voltage (Vout) and output current (Iout) show a negative correlation. The output power (Pout) of the f-TEG can be calculated by the equation below:
P o u t = V o u t 2 R i n + R e x
where Rin is the internal resistance of the f-TEG, Rex includes Rload and Rbox+ammeter (~15.7 Ω, the internal resistance of the variable resistance box and the ammeter). Rin was measured to be 264 Ω, and the resistance of the four legs (R1) was calculated by the formula R 1 = N · l / σ · A is 249 Ω (l and A are the length and cross-sectional area of one leg). Their difference comes from the contact resistance of the f-TEG. When Rin was equal to Rex, the Pout reached its maximum value (Pmax). At ΔT values of 10.3, 21.3, and 30.1 K, the Pmax values were about 29.92, 123.87, and 249.48 nW, respectively, corresponding to Rload of ~251 Ω. Therefore, Rex = Rload + Rbox+ammeter = ~251 Ω + 15.7 Ω = ~266.7 Ω, close to the Rin (264 Ω). The measured Pmax value was close to the value estimated by the equation: P m a x = V o c 2 / 4 R i n   (see Figure 6d; more details are shown in Supplementary Table S1). The maximum power density (PDmax) can be obtained by dividing Pmax by the total cross-sectional area of the f-TEG. It was 1.23 W m−2 at a ΔT of 30.1 K, which indicates that this f-TEG has a good potential for powering low-power consumption wearable electronics.

4. Conclusions

In summary, we synthesized a series of Mn-doped Cu2−xS powders by a green and facile hydrothermal method and successfully prepared Mn-doped Cu2−xS/nylon flexible films. The doping of Mn increased the n and μ, which ultimately improved the σ of the Cu2−xMnyS films. The optimal film (Cu2−xMn0.05S) exhibited a high PF of 113.3 μW m−1 K−2 at RT, with an enhanced σ of 60.1 S cm−1 and a proper α of 137.3 μV K−1. At the same time, this film possessed good flexibility: the σ was maintained at ~93.3% after bending 1000 times around a rod with a radius of 4 mm. An assembled four-leg f-TEG produced a maximum power of 249.48 nW (corresponding power density ~1.23 W m−2) at a ΔT of 30.1 K. Our results indicate that Mn doping is an effective and convenient to improve the electrical properties of Cu2−xS and provides the possibility of fabricating low-cost and high-flexibility copper-sulfide-based films for powering wearable devices.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ma16227159/s1. Figure S1: Schematic diagram of the preparation process of the Cu2−xMnyS film; Figure S2: A schematic diagram for the output performance measurement of the TE generator; Figure S3: XRD patterns of the Cu2−xMnyS powders with varying doping content of Mn; Figure S4: SEM images of Cu2−xMnyS powders. (a) y = 0.01, (b) y = 0.03, (c) y = 0.05, (d) y = 0.07; Figure S5: SEM images of the Cu2−xMnyS films. (a) y = 0.01, (b) y = 0.03, (c) y = 0.05, (d) y = 0.07; Figure S6: XPS spectra of the Cu2−xMn0.05S powders. (a) survey scan. (b–d) high-resolution scans for Cu 2p, S 2p and Mn 2p, respectively; Figure S7: Cross-sectional SEM image of the Cu2−xMn0.05S film; Table S1. The measured and estimated values of the open-circuit voltage (Voc) and maximum output power (Pmax) of the f-TEG at different ΔT.

Author Contributions

Conceptualization, K.C.; Methodology, X.Z.; Formal analysis, X.Z., X.H., Y.L. (Yiming Lu), Y.L. (Ying Liu), Z.W. and J.L.; Investigation, X.Z.; Writing—original draft, X.Z.; Writing—review & editing, X.H., Y.L. (Yiming Lu), Y.L. (Ying Liu), Z.W., J.L. and K.C.; Supervision, K.C.; Project administration, K.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (51972234 and 92163118).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article and Supplementary Materials.

Conflicts of Interest

The authors declare that they have no conflict of interest.

References

  1. Bell, L.E. Cooling, heating, generating power, and recovering waste heat with thermoelectric systems. Science 2008, 321, 1457–1461. [Google Scholar] [CrossRef]
  2. Wang, Y.; Shi, Y.; Mei, D.; Chen, Z. Wearable thermoelectric generator to harvest body heat for powering a miniaturized accelerometer. Appl. Energy 2018, 215, 690–698. [Google Scholar] [CrossRef]
  3. Zebarjadi, M.; Esfarjani, K.; Dresselhaus, M.S.; Ren, Z.F.; Chen, G. Perspectives on thermoelectrics: From fundamentals to device applications. Energy Environ. Sci. 2012, 5, 5147–5162. [Google Scholar] [CrossRef]
  4. Fan, J.; Huang, X.; Liu, F.; Deng, L.; Chen, G. Feasibility of using chemically exfoliated SnSe nanobelts in constructing flexible SWCNTs-based composite films for high-performance thermoelectric applications. Compos. Commun. 2021, 24, 100612. [Google Scholar] [CrossRef]
  5. Park, W.; Park, S.; Mun, Y.; Lee, D.; Jang, K.-S. Effects of thickness on flexibility and thermoelectric performance of free-standing Ag2Se films. J. Ind. Eng. Chem. 2023, 121, 142–148. [Google Scholar] [CrossRef]
  6. Wang, L.; Zhang, Z.; Liu, Y.; Wang, B.; Fang, L.; Qiu, J.; Zhang, K.; Wang, S. Exceptional thermoelectric properties of flexible organic−inorganic hybrids with monodispersed and periodic nanophase. Nat. Commun. 2018, 9, 3817. [Google Scholar] [CrossRef]
  7. Ding, Y.; Qiu, Y.; Cai, K.; Yao, Q.; Chen, S.; Chen, L.; He, J. High performance n-type Ag2Se film on nylon membrane for flexible thermoelectric power generator. Nat. Commun. 2019, 10, 841. [Google Scholar] [CrossRef]
  8. Howlader, S.; Vasudevan, R.; Jarwal, B.; Gupta, S.; Chen, K.H.; Sachdev, K.; Banerjee, M.K. Microstructure and mechanical stability of Bi doped Mg2Si0.4Sn0.6 thermoelectric material. J. Alloys Compd. 2020, 818, 15288. [Google Scholar] [CrossRef]
  9. Varghese, T.; Dun, C.; Kempf, N.; Saeidi-Javash, M.; Karthik, C.; Richardson, J.; Hollar, C.; Estrada, D.; Zhang, Y. Flexible Thermoelectric Devices of Ultrahigh Power Factor by Scalable Printing and Interface Engineering. Adv. Funct. Mater. 2019, 30, 5796. [Google Scholar] [CrossRef]
  10. Shang, H.; Dun, C.; Deng, Y.; Li, T.; Gao, Z.; Xiao, L.; Gu, H.; Singh, D.J.; Ren, Z.; Ding, F. Bi0.5Sb1.5Te3-based films for flexible thermoelectric devices. J. Mater. Chem. A 2020, 8, 4552–4561. [Google Scholar] [CrossRef]
  11. Gao, J.; Miao, L.; Liu, C.; Wang, X.; Peng, Y.; Wei, X.; Zhou, J.; Chen, Y.; Hashimoto, R.; Asaka, T.; et al. A novel glass-fiber-aided cold-press method for fabrication of n-type Ag2Te nanowires thermoelectric film on flexible copy-paper substrate. J. Mater. Chem. A 2017, 5, 24740–24748. [Google Scholar] [CrossRef]
  12. Li, X.; Cai, K.; Gao, M.; Du, Y.; Shen, S. Recent advances in flexible thermoelectric films and devices. Nano Energy 2021, 89, 106309. [Google Scholar] [CrossRef]
  13. Fan, Z.; Du, D.; Guan, X.; Ouyang, J. Polymer films with ultrahigh thermoelectric properties arising from significant seebeck coefficient enhancement by ion accumulation on surface. Nano Energy 2018, 51, 481–488. [Google Scholar] [CrossRef]
  14. Wang, H.; Hsu, J.H.; Yi, S.I.; Kim, S.L.; Choi, K.; Yang, G.; Yu, C. Thermally Driven Large N-Type Voltage Responses from Hybrids of Carbon Nanotubes and Poly(3,4-ethylenedioxythiophene) with Tetrakis(dimethylamino)ethylene. Adv. Mater. 2015, 27, 6855–6861. [Google Scholar] [CrossRef] [PubMed]
  15. Fan, W.; Guo, C.; Chen, G. Flexible films of poly(3,4-ethylenedioxythiophene)/carbon nanotube thermoelectric composites prepared by dynamic 3-phase interfacial electropolymerization and subsequent physical mixing. J. Mater. Chem. A 2018, 6, 12275–12280. [Google Scholar] [CrossRef]
  16. Jiang, C.; Wei, P.; Ding, Y.; Cai, K.; Tong, L.; Gao, Q.; Lu, Y.; Zhao, W.; Chen, S. Ultrahigh performance polyvinylpyrrolidone/Ag2Se composite thermoelectric film for flexible energy harvesting. Nano Energy 2021, 80, 105488. [Google Scholar] [CrossRef]
  17. Li, Y.; Lou, Q.; Yang, J.; Cai, K.; Liu, Y.; Lu, Y.; Qiu, Y.; Lu, Y.; Wang, Z.; Wu, M.; et al. Exceptionally High Power Factor Ag2Se/Se/Polypyrrole Composite Films for Flexible Thermoelectric Generators. Adv. Funct. Mater. 2021, 32, 216902. [Google Scholar] [CrossRef]
  18. Wang, Z.; Gao, Q.; Wang, W.; Lu, Y.; Cai, K.; Li, Y.; Wu, M.; He, J. High performance Ag2Se/Ag/PEDOT composite films for wearable thermoelectric power generators. Mater. Today Phys. 2021, 21, 100553. [Google Scholar] [CrossRef]
  19. Lei, Y.; Qi, R.; Chen, M.; Chen, H.; Xing, C.; Sui, F.; Gu, L.; He, W.; Zhang, Y.; Baba, T.; et al. Microstructurally Tailored Thin beta-Ag2Se Films toward Commercial Flexible Thermoelectrics. Adv. Mater. 2022, 34, 2104786. [Google Scholar] [CrossRef]
  20. Mulla, R.; Rabinal, M.H.K. Copper Sulfides: Earth-Abundant and Low-Cost Thermoelectric Materials. Energy Technol. 2019, 7, 1800850. [Google Scholar] [CrossRef]
  21. Yue, Z.; Zhou, W.; Ji, X.; Wang, Y.; Guo, F. Thermoelectric performance of hydrothermally synthesized micro/nano Cu2-xS. Chem. Eng. J. 2022, 449, 137748. [Google Scholar] [CrossRef]
  22. 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. Adv. Mater. 2014, 26, 3974–3978. [Google Scholar] [CrossRef] [PubMed]
  23. Zhao, L.; Wang, X.; Fei, F.Y.; Wang, J.; Cheng, Z.; Dou, S.; Wang, J.; Snyder, G.J. High thermoelectric and mechanical performance in highly dense Cu2−xS bulks prepared by a melt-solidification technique. J. Mater. Chem. A 2015, 3, 9432–9437. [Google Scholar] [CrossRef]
  24. Yao, Y.; Zhang, B.P.; Pei, J.; Liu, Y.C.; Li, J.F. Thermoelectric performance enhancement of Cu2S by Se doping leading to a simultaneous power factor increase and thermal conductivity reduction. J. Mater. Chem. C 2017, 5, 7845–7852. [Google Scholar] [CrossRef]
  25. Chakrabarti, D.J.; Laughlin, D.E. The Cu-S (Copper-Sulfur) System. Bull. Alloys Phase Diagr. 1983, 4, 254. [Google Scholar] [CrossRef]
  26. Tang, H.C.; Sun, F.H.; Dong, J.F.; Asfandiyar; Zhuang, H.L.; Pan, Y.; Li, J.F. Graphene network in copper sulfide leading to enhanced thermoelectric properties and thermal stability. Nano Energy 2018, 49, 267–273. [Google Scholar] [CrossRef]
  27. Yue, Z.; Zhou, W.; Ji, X.; Zhang, F.; Guo, F. Enhanced thermoelectric properties of Ag doped Cu2S by using hydrothermal method. J. Alloys Compd. 2022, 919, 165830. [Google Scholar] [CrossRef]
  28. Li, X.; Lou, Y.; Jin, K.; Fu, L.; Xu, P.; Shi, Z.; Feng, T.; Xu, B. Realizing zT>2 in Environment-Friendly Monoclinic Cu2S—Tetragonal Cu1.96S Nano-Phase Junctions for Thermoelectrics. Angew. Chem. Int. Ed. 2022, 61, e202212885. [Google Scholar] [CrossRef]
  29. Zhang, Q.H.; Huang, X.Y.; Bai, S.Q.; Shi, X.; Uher, C.; Chen, L.D. Thermoelectric Devices for Power Generation: Recent Progress and Future Challenges. Adv. Eng. Mater. 2016, 18, 194–213. [Google Scholar] [CrossRef]
  30. Jaziri, N.; Boughamoura, A.; Müller, J.; Mezghani, B.; Tounsi, F.; Ismail, M. A comprehensive review of Thermoelectric Generators: Technologies and common applications. Energy Rep. 2020, 6, 264–287. [Google Scholar] [CrossRef]
  31. Zhao, J.; Zhao, X.; Guo, R.; Zhao, Y.; Yang, C.; Zhang, L.; Liu, D.; Ren, Y. Preparation and Characterization of Screen-Printed Cu(2)S/PEDOT:PSS Hybrid Films for Flexible Thermoelectric Power Generator. Nanomaterials 2022, 12, 2430. [Google Scholar] [CrossRef] [PubMed]
  32. Liu, D.; Yan, Z.; Zhao, Y.; Zhang, Z.; Zhang, B.; Shi, P.; Xue, C. Facile self-supporting and flexible Cu2S/PEDOT:PSS composite thermoelectric film with high thermoelectric properties for body energy harvesting. Results Phys. 2021, 31. [Google Scholar] [CrossRef]
  33. Liu, Y.; Lu, Y.; Wang, Z.; Li, J.; Wei, P.; Zhao, W.; Chen, L.; Cai, K. High performance Ag2Se films by a one-pot method for a flexible thermoelectric generator. J. Mater. Chem. A 2022, 10, 25644–25651. [Google Scholar] [CrossRef]
  34. Lee, K.H.; Kim, S.-i.; Kim, H.-S.; Kim, S.W. Band Convergence in Thermoelectric Materials: Theoretical Background and Consideration on Bi–Sb–Te Alloys. ACS Appl. Energy Mater. 2020, 3, 2214–2223. [Google Scholar] [CrossRef]
  35. Peng, Q.; Zhang, S.; Yang, H.; Sheng, B.; Xu, R.; Wang, Q.; Yu, Y. Boosting Potassium Storage Performance of the Cu(2)S Anode via Morphology Engineering and Electrolyte Chemistry. ACS Nano 2020, 14, 6024–6033. [Google Scholar] [CrossRef]
  36. Tao, F.; Zhang, Y.; Zhang, F.; An, Y.; Dong, L.; Yin, Y. Structural evolution from CuS nanoflowers to Cu9S5 nanosheets and their applications in environmental pollution removal and photothermal conversion. RSC Adv. 2016, 6, 63820–63826. [Google Scholar] [CrossRef]
  37. Nandihalli, N. A short account of thermoelectric film characterization techniques. Mater. Today Phys. 2023, 36, 101173. [Google Scholar] [CrossRef]
  38. Wang, C.; Chen, F.; Sun, K.; Chen, R.; Li, M.; Zhou, X.; Sun, Y.; Chen, D.; Wang, G. Contributed Review: Instruments for measuring Seebeck coefficient of thin film thermoelectric materials: A mini-review. Rev. Sci. Instrum. 2018, 89, 5038406. [Google Scholar] [CrossRef]
  39. Nkemeni, D.S.; Yang, Z.; Lou, S.; Li, G.; Zhou, S. Achievement of extra-high thermoelectric performance in doped copper (I) sulfide. J. Alloys Compd. 2021, 878, 160128. [Google Scholar] [CrossRef]
  40. Wu, R.; Wang, D.P.; Kumar, V.; Zhou, K.; Law, A.W.; Lee, P.S.; Lou, J.; Chen, Z. MOFs-derived copper sulfides embedded within porous carbon octahedra for electrochemical capacitor applications. Chem. Commun. 2015, 51, 3109–3112. [Google Scholar] [CrossRef]
  41. Jia, X.; Zhang, E.; Yu, X.; Lu, B. Facile Synthesis of Copper Sulfide Nanosheet@Graphene Oxide for the Anode of Potassium-Ion Batteries. Energy Technol. 2019, 8, 1900987. [Google Scholar] [CrossRef]
  42. Geng, X.; Jiao, Y.; Han, Y.; Mukhopadhyay, A.; Yang, L.; Zhu, H. Freestanding Metallic 1T MoS2 with Dual Ion Diffusion Paths as High Rate Anode for Sodium-Ion Batteries. Adv. Funct. Mater. 2017, 27, 1702998. [Google Scholar] [CrossRef]
  43. Kutty, T.R.N. A controlled copper-coating method for the preparation of ZnS:Mn DC electroluminescent powder phosphors. Mater. Res. Bull. 1991, 26, 399–406. [Google Scholar] [CrossRef]
  44. Chen, T.W.; Rajaji, U.; Chen, S.M.; Li, Y.L.; Ramalingam, R.J. Ultrasound-assisted synthesis of alpha-MnS (alabandite) nanoparticles decorated reduced graphene oxide hybrids: Enhanced electrocatalyst for electrochemical detection of Parkinson’s disease biomarker. Ultrason. Sonochem. 2019, 56, 378–385. [Google Scholar] [CrossRef]
  45. Xin, Y.; Cao, L.; Huang, J.; Liu, J.; Fei, J.; Yao, C. Influence of S/Mn molar ratio on the morphology and optical property of γ-MnS thin films prepared by microwave hydrothermal. J. Alloys Compd. 2013, 549, 1–5. [Google Scholar]
  46. Rastogi, A.C.; Salkalachen, S.; Bhide, V.G. Electrical conduction and phase transitions in vacuum-deposited Cu2-xS films. Thin Solid Films 1978, 52, 1–10. [Google Scholar] [CrossRef]
  47. Wang, Y.; Long, Z.; Cheng, Y.; Zhou, M.; Chen, H.; Zhao, K.; Shi, X. Chemical bonding engineering for high-symmetry Cu2S-based materials with high thermoelectric performance. Mater. Today Phys. 2023, 32, 101028. [Google Scholar] [CrossRef]
  48. Liang, X.; Jin, D.; Dai, F. Phase Transition Engineering of Cu2S to Widen the Temperature Window of Improved Thermoelectric Performance. Adv. Electron. Mater. 2019, 5, 1900486. [Google Scholar] [CrossRef]
  49. Evans, H.T. Djurleite (Cu1.94S) and low chalcocite (Cu2S): New crystal structure studies. Science 1979, 203, 356–358. [Google Scholar] [CrossRef]
  50. He, Y.; Lu, P.; Shi, X.; Xu, F.; Zhang, T.; Snyder, G.J.; Uher, C.; Chen, L. Ultrahigh Thermoelectric Performance in Mosaic Crystals. Adv. Mater. 2015, 27, 3639–3644. [Google Scholar] [CrossRef]
  51. Yao, Y.; Zhang, B.P.; Pei, J.; Sun, Q.; Nie, G.; Zhang, W.Z.; Zhuo, Z.T.; Zhou, W. High Thermoelectric Figure of Merit Achieved in Cu(2)S(1- x)Te (x) Alloys Synthesized by Mechanical Alloying and Spark Plasma Sintering. ACS Appl. Mater. Interfaces 2018, 10, 32201–32211. [Google Scholar] [CrossRef] [PubMed]
  52. Yang, M.; Liu, X.; Zhang, B.; Chen, Y.; Wang, H.; Yu, J.; Wang, G.; Xu, J.; Zhou, X.; Han, G. Phase Tuning for Enhancing the Thermoelectric Performance of Solution-Synthesized Cu2–xS. ACS Appl. Mater. Interfaces 2021, 13, 39541–39549. [Google Scholar] [CrossRef]
  53. Chen, X.Q.; Fan, S.J.; Han, C.; Wu, T.; Wang, L.J.; Jiang, W.; Dai, W.; Yang, J.P. Multiscale architectures boosting thermoelectric performance of copper sulfide compound. Rare Met. 2021, 40, 2017–2025. [Google Scholar] [CrossRef] [PubMed]
  54. Zhang, R.; Pei, J.; Han, Z.J.; Wu, Y.; Zhao, Z.; Zhang, B.P. Optimal performance of Cu1.8S1−xTex thermoelectric materials fabricated via high-pressure process at room temperature. J. Adv. Ceram. 2020, 9, 535–543. [Google Scholar] [CrossRef]
Figure 1. (a) XRD patterns of the Cu2−xMnyS films with varying doping content of Mn. (b) A typical SEM image at low magnification, inset at high magnification. (ce) Corresponding element mappings of the Cu2−xMn0.05S film.
Figure 1. (a) XRD patterns of the Cu2−xMnyS films with varying doping content of Mn. (b) A typical SEM image at low magnification, inset at high magnification. (ce) Corresponding element mappings of the Cu2−xMn0.05S film.
Materials 16 07159 g001
Figure 2. XPS spectra of the Cu2−xMn0.05S film. (a) Survey scan. (bd) High-resolution scans for Cu 2p, S 2p, and Mn 2p, respectively.
Figure 2. XPS spectra of the Cu2−xMn0.05S film. (a) Survey scan. (bd) High-resolution scans for Cu 2p, S 2p, and Mn 2p, respectively.
Materials 16 07159 g002
Figure 3. The temperature-dependent (a) electrical conductivities and (b) Seebeck coefficients. (c) Seebeck coefficient as a function of carrier concentration for Cu2−xS-, Cu2−xMn0.05S-, and reported Cu2−xS-based bulks [21,50,51] at 300 K. (d) Temperature-dependent power factors of the Cu2−xMnyS films.
Figure 3. The temperature-dependent (a) electrical conductivities and (b) Seebeck coefficients. (c) Seebeck coefficient as a function of carrier concentration for Cu2−xS-, Cu2−xMn0.05S-, and reported Cu2−xS-based bulks [21,50,51] at 300 K. (d) Temperature-dependent power factors of the Cu2−xMnyS films.
Materials 16 07159 g003
Figure 4. Microstructure characterization of grains scraped from the Cu2−xMn0.05S film. (ac) TEM images. (df) Enlarged images of the blue, red, and orange squares marked in (ac), respectively. The “” in (d) denotes edge dislocation. (gi) FFT images corresponding to the grains marked with green, purple, and pink squares in (df), respectively.
Figure 4. Microstructure characterization of grains scraped from the Cu2−xMn0.05S film. (ac) TEM images. (df) Enlarged images of the blue, red, and orange squares marked in (ac), respectively. The “” in (d) denotes edge dislocation. (gi) FFT images corresponding to the grains marked with green, purple, and pink squares in (df), respectively.
Materials 16 07159 g004
Figure 5. (a) Flexibility of the Cu2−xMn0.05S film: The relative electrical conductivity changes with bending around a bending radius of 4 mm. (b) An HRTEM image of the Cu2−xMn0.05S film showing a good combination between the film and the nylon membrane; the inset is the corresponding IFFT image.
Figure 5. (a) Flexibility of the Cu2−xMn0.05S film: The relative electrical conductivity changes with bending around a bending radius of 4 mm. (b) An HRTEM image of the Cu2−xMn0.05S film showing a good combination between the film and the nylon membrane; the inset is the corresponding IFFT image.
Materials 16 07159 g005
Figure 6. Output performance of the f-TEG assembled with the Cu2−xMn0.05S film. (a) The open-circuit voltage at different ΔT (the inset is a schematic diagram of the four-leg f-TEG). (b) Output voltage and power versus current at different ΔT. The straight lines and curves in (b) correspond to V-I and P-I relations, respectively. Comparison of the estimated and measured values of the (c) open circuit voltage (Voc) and (d) maximum output power (Pmax) under different ΔT.
Figure 6. Output performance of the f-TEG assembled with the Cu2−xMn0.05S film. (a) The open-circuit voltage at different ΔT (the inset is a schematic diagram of the four-leg f-TEG). (b) Output voltage and power versus current at different ΔT. The straight lines and curves in (b) correspond to V-I and P-I relations, respectively. Comparison of the estimated and measured values of the (c) open circuit voltage (Voc) and (d) maximum output power (Pmax) under different ΔT.
Materials 16 07159 g006
Table 1. TE performance of the reported copper-sulfide-based bulks and flexible films.
Table 1. TE performance of the reported copper-sulfide-based bulks and flexible films.
MaterialsPF (μW m−1K−2)Temperature (K)MethodRef. No
Cu2−xS56300Solvothermal and HP[52]
Cu2−xS75300Melting and SPS[53]
Cu1.98−2xMnxS0.985Se0.015100325Melting and SPS[47]
Cu2S1−xSex122325Ball milling and SPS[24]
Micro/nano Cu2−xS250320Hydrothermal and SPS[21]
Cu2S1−xTex145325Ball milling and SPS[54]
Cu2S hybrid films20393Screen printing[31]
Cu2S/PEDOT:PSS films56393Vacuum filtration[32]
Cu2−xMnyS/nylon films113300Hydrothermal and HPThis work
150393
Note: The data related to copper-sulfide-based bulks are estimated from the relevant reported graphs.
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

Zuo, X.; Han, X.; Lu, Y.; Liu, Y.; Wang, Z.; Li, J.; Cai, K. Largely Enhanced Thermoelectric Power Factor of Flexible Cu2−xS Film by Doping Mn. Materials 2023, 16, 7159. https://doi.org/10.3390/ma16227159

AMA Style

Zuo X, Han X, Lu Y, Liu Y, Wang Z, Li J, Cai K. Largely Enhanced Thermoelectric Power Factor of Flexible Cu2−xS Film by Doping Mn. Materials. 2023; 16(22):7159. https://doi.org/10.3390/ma16227159

Chicago/Turabian Style

Zuo, Xinru, Xiaowen Han, Yiming Lu, Ying Liu, Zixing Wang, Jiajia Li, and Kefeng Cai. 2023. "Largely Enhanced Thermoelectric Power Factor of Flexible Cu2−xS Film by Doping Mn" Materials 16, no. 22: 7159. https://doi.org/10.3390/ma16227159

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