This study reports the optoelectronic and thermoelectric properties of antimony trisulfide (Sb2S3) under a hydrostatic pressure of up to 20.4 GPa. The properties were computed based on the full-potential linearized augmented plane wave using the generalized gradient approximation by Perdew, Burke, and Ernzerhof as the exchange-correlation potentials. It was shown that increasing the pressure from 0.00 to 20.4 GPa decreases the calculated bandgap from 1.44 to 0.84 eV. There was a discontinuity in the pressure range of 4.82–6.3 GPa due to the isostructural electronic phase transition. The applied pressure increases the inner electrical polarization. At high pressure, the energy of the negative value of ε1 becomes large, and ε1 itself always remains negative. We observed that the high absorption of Sb2S3 also increases with pressure and the plasmon energy shifts to high energy. The applied pressure increases the static dielectric constant and static refractive index. It was found that the Seebeck coefficients increase with increasing temperature and decrease with increasing pressure. The bipolar effect occurs at low doping levels and high pressure. The optical and thermoelectric properties of Sb2S3 obtained under pressure show that it is suitable for clean energy conversion and optoelectronic applications.

Antimony trisulfide (Sb2S3) is a V–VI semiconductor compound. It is a layered structure composed of [Sb4S6]n atomic chains with van der Waals interaction.1 Among the various metal-sulfides, Sb2S3 has attracted much interest first because of its characteristics such as low-cost, abundance in nature, non-toxicity, low melting point,2–4 high absorption coefficient,5 sufficient bandgap, and other good optical6,7 and thermoelectrical properties8 and second due to the wide range of applications in optoelectronics, microwave devices,9 switching,10 batteries,11 and photovoltaic devices such as solar cells.12 Annealed Sb2S3 thin films are promising materials for developing solar cell technologies.13 It has been shown that n-Sb2S3/p-Si heterojunction solar cells with improved photovoltaic properties have been fabricated through low-cost techniques.14 One density functional theory (DFT) calculation has proved that the two-step doping strategy using O-dopants and introducing other p-type dopants increases the efficiency of Sb2S3-based solar cells.15 

Sb2S3 compounds have been fabricated by various techniques, such as chemical deposition and thermal evaporation.14,16–20 Because of the size and the nature of the bandgap, the electronic band structure of Sb2S3 has remained controversial for decades. Theoretical and experimental studies have presented the direct and indirect bandgap of Sb2S3 in the range of 1.18–2.24 eV.15 Sb2S3 nanoribbons were successfully synthesized, and the optical absorption experiment revealed that Sb2S3 nanoribbons have a bandgap of about 1.15.21 

The electrical and optical investigations have demonstrated that the indirect bandgap energy values of Sb2S3 thin film increase with increasing thickness22 and the presence of Ni-doped atoms decreases the bandgap.6 It has also been shown that Ni-doped-Sb2S3 leads to an increased optical absorption coefficient in the visible region, and it is introduced as a promising candidate for optoelectronic applications. It has been shown that the spin–orbit interaction has no effect on the electronic properties.7 Studies have shown that the thermoelectric (TE) performance of n-type Sb2S3 is better than that of the p-type and the optimal doping concentration is in the order of 1021 cm−3 for all temperatures.8,23 Furthermore, it has become clear that the electronic properties are anisotropic and adding Be atoms to Sb2S3 decreases the Seebeck coefficient to 595 (−635) µV/K for p- (n-)type carriers at 300 K temperature. Bi2S3 is another layered semiconductor with a direct bandgap (Eg ≈ 1.3 eV), which is isostructural to Sb2S3 with an orthorhombic crystal structure. Bi2S3 has been proposed as a good thermoelectric (TE) material with a high Seebeck coefficient (about 500 µV/K) and low thermal conductivity (<1 W/mK). We have also found that optimizing the ball-milling process enhances the thermoelectric properties of polycrystalline Bi2S3.24 The experimental investigation of the thermoelectric properties of (Bi1-xSbx)2S3 has also shown that for x = 0, the Seebeck coefficient is |s| = 180 µV/K and when increasing the concentration of Sb atoms to x = 0.6 and 0.8, the Seebeck coefficient becomes larger, about |s| = 230 and 460 µV/K.25 

Pressure tunes the bandgap and changes the optical properties, and it is able to balance the TE performance such as power factor and figure of merit.24,26–28 The experimental and theoretical investigation shows that Sb2S3 has a stable structure at pressures up to 20 GPa and no first-order phase transition has been observed.4,29–31

So far, numerous studies have been carried out on synthesis of the electronic, optic, and thermoelectric properties and the applications of Sb2S3.14–20 To complete these known aspects, we have tried to investigate the electronic, optic, and thermoelectric of Sb2S3 under pressure. We also investigated the effect of different types and concentrations of doping (from 1018 to 1021 cm−3) and temperature (from 50 to 800 K) on thermoelectric properties. Our calculations show that applied pressure increases the high absorption of Sb2S3 in the extreme ultraviolet region. In addition, the bipolar effect is observable on the Seebeck coefficient curve at high pressures for the 1018 cm−3 doping concertation.

This paper is arranged as follows: The computational details are reviewed in Sec. II. In Sec. III, the obtained results, such as structural properties, band structures, electron density (Subsection III A), optical properties (Subsection III B), and thermoelectric results (Subsection III C), are discussed. Finally, a summary of the results is presented.

The theoretical investigations of the crystal structure and optoelectronic and thermoelectric properties of Sb2S3 under hydrostatic pressure have been performed using the WIEN2k package and the full-potential linearized augmented plane wave method (FP-LAPW).32 The Generalized Gradient Approximation by Perdew–Burke–Ernzerhof (PBE-GGA) functional33 is used to calculate the exchange-correlation potential for all the steps. The basis function is expanded to Kmax × RMT = 7.0, where Kmax is the plane wave cut-off and RMT is the smallest atomic muffin tin sphere radius in the unit cell. The magnitude of the largest vector in the Fourier expansion of charge density (Gmax) is equal to 12 bohrs−1. The number of k-points for the integration in the first Brillouin zone (BZ) is 100, 200, and 2000. The self-consistent calculations are converged when the variation in the total energy is less than 0.0001 Ry. The thermoelectric features are calculated via BoltzTraP,34 which is supported by the WIEN2k package. The thermoelectric calculations are explained in detail in Subsection III C.

Under ambient conditions, the stable form of Sb2S3 at room temperature is the orthorhombic structure with a Pnma space group.35,36 The primitive unit cell contains four molecules (Z = 4). Sb2S3 has two and three non-equivalent atomic positions for Sb [Sb (1) and Sb (2)] and S [S (1), S (2), and S (3)]. The experimental lattice parameters of Sb2S3 are a = 11.3045 Å, b = 3.8411 Å, c = 11.2386 Å, and α = β = γ = 90° at ambient pressure.31 The applied pressure changes these lattice parameters. We have extracted the values of these lattice parameters under pressure (up to 20.04 GPa) using the experimental results.37 The optimized total energy of Sb2S3 in terms of (% c/a) with the experimental lattice parameters at different pressures was used to obtain the theoretical lattice parameters.

As shown in Fig. 1, the optimized and experimental lattice parameters decrease with increasing pressure. Indeed, we can observe a linear decrease up to 4.82 GPa and an increase at a pressure of 6.34 GPa; then there is another decrease. The optimized lattice parameters at different pressures are in good agreement with the experimental results.37,38

FIG. 1.

(a) Total energy in terms of % c/a under ambient conditions. (b)–(d) Optimized and experimental (Exp. a37 and Exp. b38) lattice parameters of the orthorhombic crystal structure of Sb2S3 as a function of pressure.

FIG. 1.

(a) Total energy in terms of % c/a under ambient conditions. (b)–(d) Optimized and experimental (Exp. a37 and Exp. b38) lattice parameters of the orthorhombic crystal structure of Sb2S3 as a function of pressure.

Close modal

Based on the optimized lattice parameters, we have investigated the band structures of Sb2S3 under different pressures. The band structure of Sb2S3 under pressures of 0.00, 3.29, 4.82, 6.34, 11.79, and 20.04 GPa is shown in Fig. 2.

FIG. 2.

Calculated band structure of Sb2S3 using the PBE-GGA under different pressures: (a) 0.00 GPa, (b) 3.29 GPa, (c) 4.48 GPa, (d) 6.34 GPa, (e) 11.79 GPa, and (f) 20.04 GPa. D refers to the direct bandgap, and I refers to the indirect bandgap.

FIG. 2.

Calculated band structure of Sb2S3 using the PBE-GGA under different pressures: (a) 0.00 GPa, (b) 3.29 GPa, (c) 4.48 GPa, (d) 6.34 GPa, (e) 11.79 GPa, and (f) 20.04 GPa. D refers to the direct bandgap, and I refers to the indirect bandgap.

Close modal

The calculated bandgap of Sb2S3 is given in Table I. It is evident that from 0.00 to 4.82 GPa, the valence band maximum (VBM) and the conduction band minimum (CBM) are located at the Γ point (direct bandgap) while at pressures of 6.34–11.79 GPa, the VBM is shifted to between R and the Γ point.

TABLE I.

The calculated bandgap, refractive indices (n0), and the static dielectric constants (ε0) of Sb2S3 under different pressures. D refers to the direct bandgap, and I refers to the indirect bandgap.

Pressure (GPa)Egn0 − xxn0 − yyn0 − zzε0 − xxε0 − yyε0 − zz
0.00 1.440 D 3.367 05 4.040 03 3.923 79 11.3366 16.3208 15.3952 
0.66 1.426 D 3.372 04 4.059 95 3.936 26 11.3702 16.4823 15.4932 
1.12 1.415 D 3.383 36 4.089 72 3.956 25 11.4467 16.7248 15.651 
2.10 1.397 D 3.393 38 4.133 91 3.968 37 11.5146 17.0882 15.7471 
3.29 1.365 D 3.423 35 4.192 64 4.005 25 11.7189 17.5772 16.0411 
3.99 1.339 D 3.436 82 4.224 17 4.025 6 11.8113 17.8426 16.2045 
4.82 1.330 D 3.441 51 4.230 73 4.036 12 11.8435 17.898 16.2893 
6.34 1.317 I 3.445 39 4.241 62 4.041 82 11.8703 17.9903 16.3353 
7.58 1.281 I 3.465 1 4.278 8 4.069 39 12.0065 18.307 16.5589 
8.83 1.241 I 3.487 04 4.322 91 4.099 07 12.1545 18.6765 16.7923 
10.01 1.188 I 3.505 5 4.366 69 4.129 25 12.2881 19.0668 17.0497 
11.79 1.117 I 3.537 6 4.437 91 4.181 52 12.5141 19.6937 17.484 
13.22 1.103 I 3.559 34 4.464 79 4.209 52 12.6684 19.933 17.7189 
15.28 1.036 I 3.599 96 4.565 02 4.279 76 12.9592 20.8379 18.3151 
17.22 0.960 I 3.633 43 4.638 79 4.340 24 13.2012 21.5167 18.8363 
20.04 0.843 I 3.688 18 4.767 46 4.445 6 13.602 22.7268 19.7616 
Pressure (GPa)Egn0 − xxn0 − yyn0 − zzε0 − xxε0 − yyε0 − zz
0.00 1.440 D 3.367 05 4.040 03 3.923 79 11.3366 16.3208 15.3952 
0.66 1.426 D 3.372 04 4.059 95 3.936 26 11.3702 16.4823 15.4932 
1.12 1.415 D 3.383 36 4.089 72 3.956 25 11.4467 16.7248 15.651 
2.10 1.397 D 3.393 38 4.133 91 3.968 37 11.5146 17.0882 15.7471 
3.29 1.365 D 3.423 35 4.192 64 4.005 25 11.7189 17.5772 16.0411 
3.99 1.339 D 3.436 82 4.224 17 4.025 6 11.8113 17.8426 16.2045 
4.82 1.330 D 3.441 51 4.230 73 4.036 12 11.8435 17.898 16.2893 
6.34 1.317 I 3.445 39 4.241 62 4.041 82 11.8703 17.9903 16.3353 
7.58 1.281 I 3.465 1 4.278 8 4.069 39 12.0065 18.307 16.5589 
8.83 1.241 I 3.487 04 4.322 91 4.099 07 12.1545 18.6765 16.7923 
10.01 1.188 I 3.505 5 4.366 69 4.129 25 12.2881 19.0668 17.0497 
11.79 1.117 I 3.537 6 4.437 91 4.181 52 12.5141 19.6937 17.484 
13.22 1.103 I 3.559 34 4.464 79 4.209 52 12.6684 19.933 17.7189 
15.28 1.036 I 3.599 96 4.565 02 4.279 76 12.9592 20.8379 18.3151 
17.22 0.960 I 3.633 43 4.638 79 4.340 24 13.2012 21.5167 18.8363 
20.04 0.843 I 3.688 18 4.767 46 4.445 6 13.602 22.7268 19.7616 

At pressures more than 13.22 GPa, both VBM and CBM are shifted between R and the Γ point. With increasing pressure, the CBM also is moved to the lower energy, so the bandgap of Sb2S3 becomes small. The variations in the bandgap in terms of pressures are presented in Fig. 3.

FIG. 3.

Variations in the calculated electronic bandgap of Sb2S3 with pressures in comparison with the Exp.4 results.

FIG. 3.

Variations in the calculated electronic bandgap of Sb2S3 with pressures in comparison with the Exp.4 results.

Close modal

Two discontinuities are observed: the first one is between the pressures of 4.82 and 6.34 GPa, and the second one is between the pressures of 11.79 and 13.22 GPa. These two discontinuities confirm the changes in the VBM and CBM under pressure and are in good agreement with previous observations.4 The calculated bandgap at ambient pressure is 1.44 eV, which is low compared to the experimental value (1.73 eV). However, the difference between the estimated bandgap and the experimentally obtained bandgap is 0.29 eV. This underestimation (between DFT and the experiment) can be seen in most of the DFT calculations, and it is due to the calculation of the exchange-correlation potential by the GGA.39 

Figure 4 shows the total and partial density of state (DOS) in terms of energy at ambient pressure. As can be observed, the total DOS confirms the semiconductor nature of Sb2S3 with a bandgap of 1.4 eV. The value of the bandgap at 0.00 GPa is in good agreement with other reported results.4,40 In the partial DOS, the s orbitals of S and Sb atoms are considered important in the energy range from −14 to −11.7 eV and from −61 to −9 eV, respectively. At the energy of valence bands of −5 to 0 eV, the p orbitals of S (S-p) and Sb (Sb-p) atoms play a main role so that the contribution of S-p is more than that of Sb-p. This observation is reversed at the conduction energy levels, so the contribution of Sb-p becomes more than that of S-p.

FIG. 4.

Total and partial density of state at ambient pressure using the PBE-GGA.

FIG. 4.

Total and partial density of state at ambient pressure using the PBE-GGA.

Close modal

Considering the DOS and large hybridization of S and Sb atoms at the valence and conduction electron energy levels, the Sb–S bond is a strong covalent bond.15 The variations in the total DOS under applied pressures near the Fermi level (at zero energy) are illustrated in Fig. 5. As seen with increasing pressure, the DOS of conduction electrons shifts to a lower energy while the variations in the valence electron DOS are negligible.

FIG. 5.

Total density of states near the Fermi energy level under different pressures. By increasing the pressure, the bottom of the conduction band shifts to low energies except at 6.34 GPa.

FIG. 5.

Total density of states near the Fermi energy level under different pressures. By increasing the pressure, the bottom of the conduction band shifts to low energies except at 6.34 GPa.

Close modal

Therefore, the increase in pressure decreases the bandgap, and it also predicts the redshift on the absorption spectrum. As mentioned before, an anomaly in the bandgap between pressures 4.82 and 6.34 GPa is observed, which is due to the change in its nature. The result points to the isostructural electronic phase transition.4,37

The study of light interaction and the response of the compound are necessary for optoelectronic applications. The investigation of optical properties begins with the study of the dielectric function that includes real and imaginary parts. The real and imaginary parts of the dielectric function are calculated by41,42
(1)
(2)
(3)
where α and β refer to the x-, y-, and z-directions, δ is the delta function, P represents the Cauchy integral part, ck and vk are the electronic states at the conduction band and the valence band, respectively, and P is a momentum operator. Other optical quantities are related to the real and imaginary parts of the dielectric function and are obtained as follows:
(4)
(5)
(6)
(7)
(8)
where α, n, k, R, and Eloss are the absorption coefficient, refractive index, extinction coefficient, reflectivity, and energy electron loss function, respectively. The orthorhombic crystal structure of Sb2S3 has three independent optical components in the x, y, and z directions.

The real part of the dielectric function shows the electrical polarization of the compound. The imaginary part, which is the sum of all inter-transitions between electronically occupied and unoccupied states, shows a loss.

The real part of the dielectric function at zero energy is called the static dielectric constant ε(0). The calculated static dielectric constants are given in Table I in the three directions. As shown in Table I, the static dielectric constant increases by increasing the pressure. The static dielectric constants at ambient pressure are 11.34, 16.32, and 15.39 in the x, y, and z directions, respectively. These values reach 13.6, 22.73, and 19.76 at 20.04 GPa. The calculated dielectric constants at ambient pressure are also in close agreement with those of the experiment.43 

The real part of the dielectric function is related to the dielectric polarization. As shown in Fig. 6, the calculated spectra show that there is a maximum value at an energy of 2 eV. It means that the light passing through the crystal with the energy in the visible range (1.75–3.11 eV) becomes strongly polarized, and this polarization increases with increasing the pressure, which leads to lattice compression. The lattice parameter in the y direction is smaller than that in the x and z directions. Hence, it can be concluded that the maximum dielectric constant and polarization exist in the y direction. The maximum polarization value in the y direction is 29.27, which reaches 33.14 by increasing the pressure.

FIG. 6.

Real [(a)–(c)] and imaginary [(d)–(f)] parts of the dielectric function in terms of energy in three directions for Sb2S3 under different pressures.

FIG. 6.

Real [(a)–(c)] and imaginary [(d)–(f)] parts of the dielectric function in terms of energy in three directions for Sb2S3 under different pressures.

Close modal

The real dielectric function has negative values after energy values of 5, 3.12, and 3.52 eV in the x, y, and z directions, respectively. At these energy values, Sb2S3 behaves as a metal and reflects the electromagnetic wave completely.

The pressure moves the negative ε1 to higher energy (5.92, 4.12, and 4.50 eV in the x, y, and z directions, respectively). Increasing the pressure widens the energy range for negative values of ε1, and in this energy range, the real dielectric part always remains negative. This feature makes Sb2S3 suitable for supercooling and new photonic applications.44,45

The peaks of the imaginary dielectric function imply transitions between the occupied valence bands and unoccupied conduction bands and present loss. For Sb2S3, energies of 3.41, 2.73, and 3.2 eV in the x, y, and z-directions, respectively, show the maximum peaks of the imaginary dielectric function, which indicates maximum loss. The applied pressure shifts the energy of maximum loss to higher energy in the x direction while decreasing the energy in the y and z directions.

By comparing the dielectric spectra in the three different directions, it can be seen that Sb2S3 has an anisotropic optical behavior. In fact, the difference in the charge density distribution along the different directions leads to different dielectric functions, resulting in anisotropic optical and photoelectric properties. The anisotropy in the dielectric function of Sb2S3 exhibits ferroelectric phase transition and piezoelectric behavior of this material at the nanoscale.46 It is found that Sb2S3 as an anisotropic semiconductor with a wide optical response to the incident light can be used in polarized photodetection.47 

Figure 7 shows the absorption coefficients in terms of wavelength. The maximum absorption coefficient is the wavelength range of 100–125 nm and with increasing wavelength, the absorption decreases. As seen, the absorption coefficient of Sb2S3 is nonzero in the visible region, and it has a maximum value in the y direction.

FIG. 7.

Absorption coefficient spectrum in three directions—(a) x, (b) y, and (c) z directions—under different pressures.

FIG. 7.

Absorption coefficient spectrum in three directions—(a) x, (b) y, and (c) z directions—under different pressures.

Close modal

Applying the pressure increases the absorption in the extreme ultraviolet region while at wavelengths of 250–300 nm, the pressure causes the absorption to decrease. Next, in the visible region, the pressure increases the absorption slowly. In conclusion, the high absorption of Sb2S3 is 180.56, 159.29, and 173.00 (104 cm−1), and the pressure increases them to values of 216.242, 195.81, and 192.48 (104 cm−1), respectively, in the wavelength range of 100–125 nm in the x, y, and z directions.

The refractive index is shown in Fig. 8. The static refractive indices are given in Table I. These values, under ambient pressure, are 3.37, 4.04, and 3.92 in the x, y, and z directions, which are in agreement with other studies.5–7,22 As seen, the effect of applied pressure is the increase in static refractive indices. The maximum scattering of light is observed in the y direction with an energy of 2.2 eV. After energy values of 10, 9, and 10.5 eV in the x, y, and z directions, refractive values become less than 1, and the superluminal effect occurs.

FIG. 8.

Refractive index spectrum in three directions—(a) x, (b) y, and (c) z directions—under different pressures.

FIG. 8.

Refractive index spectrum in three directions—(a) x, (b) y, and (c) z directions—under different pressures.

Close modal

Figure 9 presents the excitation coefficient of Sb2S3 in terms of energy that determine the light loss inside the compound, and the peaks correspond to the transition between the occupied and unoccupied electron bands. The extinction peaks in the x, y and z directions in the energy range of 3–4 eV decrease with increasing pressure.

FIG. 9.

Extinction coefficient spectrum in three directions—(a) x, (b) y, and (c) z directions—under different pressures.

FIG. 9.

Extinction coefficient spectrum in three directions—(a) x, (b) y, and (c) z directions—under different pressures.

Close modal

Another optical property of Sb2S3 is reflectivity, which is shown in Fig. 10. Our results present that the reflectivity of Sb2S3 in the visible region (gray area) is from 37% to 57% and it increased slightly with pressure.

FIG. 10.

Reflectivity coefficient spectrum in three directions—(a) x, (b) y, and (c) z directions—under different pressures.

FIG. 10.

Reflectivity coefficient spectrum in three directions—(a) x, (b) y, and (c) z directions—under different pressures.

Close modal

It was also found that Sb2S3 has maximum reflectivity of about 60% at an energy of 3.90 eV (ultraviolet region) in the z direction and under ambient pressure. Above energies of 5 eV, a reflectivity increase can be observed with increasing pressure. Our results under ambient pressure are in good agreement with those of other studies.6,22

Plasmon energies can be determined with the energy of remarkable peaks, as illustrated in Fig. 11. Under ambient pressure, the Eloss peaks are located at energies of 18.97, 19.44, and 19.35 eV in the x, y, and z directions, respectively. Our results show that the pressure shifts these peaks to higher energy and causes the loss intensity to increase. The plasmon energies are 21.29, 21.15, and 20.80 eV at 20.04 GPa.

FIG. 11.

Variation in the energy electron loss spectrum Eloss (ω) at three directions—(a) x, (b) y, and (c) z directions—under different pressures.

FIG. 11.

Variation in the energy electron loss spectrum Eloss (ω) at three directions—(a) x, (b) y, and (c) z directions—under different pressures.

Close modal
Concerning semiclassical Boltzmann theory within the rigid band and constant relaxation time approximations as implemented in the BoltzTraP program,34 the transport coefficients such as electrical (σ) and thermal (κ) conductivities, the Seebeck coefficient (S), and the figure of merit (ZT) are calculated based on following relations:
(9)
(10)
(11)
(12)
(13)
(14)
where N is the number of k-points in the k summation; τi,k is the relaxation time; and T, μ, Ω, e, f0, and ε are the temperature, chemical potential, the volume of the unit cell, electron charge, Fermi–Dirac equilibrium distribution function, and energy, respectively. Here, κ is the sum of electrons and lattice thermal conductivities (κe + κlat). It was shown that κlat for Sb2S3 is small8 and κe is calculated as κe = κ0 + σTS2.

To the best of the researchers’ knowledge, the thermoelectric properties of Sb2S3, under hydrostatic pressure, have not been investigated yet. The results are obtained for different doping levels of 1 × 1018, 1 × 1019, 1 × 1020, and 1 × 1021 per cm−3, from 50 to 800 K temperatures, and under applied different pressures from 0 to 20.4 GPa. A dense k-mesh of 22 000 k-points has been used to calculate the transport properties.

First, the temperature dependence of the Seebeck coefficient under applied pressures at different doping levels has been investigated. To simplify, Fig. 12 shows the temperature dependence of the Seebeck coefficient under selected applied pressures (0.00, 7.58, 13.22, and 20.04 GPa) and doping levels of 1018 and 1021 per cm−3. The positive (negative) values of the Seebeck coefficient indicate the p- (n-)type carriers in the Sb2S3 compound. With respect to vertical axes, it is clear that the increasing the doping concentrations leads to a reduction in the Seebeck coefficient.

FIG. 12.

Seebeck coefficient in terms of temperature at doping levels of 1018 (a) and 1021 (b) for p-type and the n-type carriers under different pressures.

FIG. 12.

Seebeck coefficient in terms of temperature at doping levels of 1018 (a) and 1021 (b) for p-type and the n-type carriers under different pressures.

Close modal

The Seebeck coefficients increase with increasing temperature except for the pressure range of 13.22–20.04 GPa at a doping level of 1018 cm−3 and at 20.04 GPa at a doping level of 1019 cm−3. Indeed, the Seebeck coefficients increase with increasing temperature, and then there is a remarkable decrease. In other words, the bipolar effect occurs at high pressures and low doping concentrations. The bipolar effect that commonly occurs in narrow bandgap semiconductors indicates the effect of both carrier types on electrical transport.48–51 The result of the bipolar effect is the increase in thermal conductivity in comparison to the case where a type of carrier plays a role in transport.

It was found that the applied pressure generally decreases the Seebeck coefficient at different doping concentrations. In the temperature range of 50–250 K and at a doping level of 1021 cm−3, the Seebeck coefficient of n-type carriers first decreases with increasing pressure up to 4 GPa and then increases. At ambient pressure and a temperature of 300 K, the Seebeck coefficient is 612.6 (−536.09), 414.67 (−338.27), 221.20 (−146.73), and 86.82 (−32.13) µV/K for p- (n-)type carriers at doping levels from 1018 to 1021 cm−3, respectively. These values are comparable with the experimental values of Sb2S3 thin film52 and other compounds similar to stibnite, such as Sb2Se3 (about 580 µV/K with a direct bandgap of 1.43 eV),53 Bi2Se3 (−100 µV/K topological insulator),54,55 In2S3 (200–400 µV/K with an indirect bandgap of about 1.04 eV),52 and Bi2S3 (about 500 µV/K with a direct bandgap of 1.3 eV).24 

Figure 13 shows the variations in electrical conductivity with temperature and pressure at different doping levels. In most curves at different doping levels, the variation in electrical conductivity is approximately flat, which means that the conductivity is independent of temperature at ambient pressure.

FIG. 13.

Electrical conductivity in terms of temperature at doping levels of 1018 [(a) and (b)] and 1021 [(c) and (d)] for p-type and the n-type carriers under different pressures.

FIG. 13.

Electrical conductivity in terms of temperature at doping levels of 1018 [(a) and (b)] and 1021 [(c) and (d)] for p-type and the n-type carriers under different pressures.

Close modal

The electrical conductivity decreases slowly because of the temperature at high pressure, except that the increase in temperature leads to an exponential increase in the conductivity at a low doping level of 1018 cm−3. As mentioned before, the bipolar effect results in the increase in electrical conductivity at 1018 cm−3 doping concentration, and increasing doping concentration increases the electrical conductivity. For all pressures, n-type doping of Sb2S3 has greater electronic conductivity than p-type doping. The values of p- (n-)type electrical conductivity at 0.00 GPa and 300 K are 2.21 × 1016 (6.23 × 1016), 2.20 × 1017 (6.22 × 1017), 2.12 × 1018 (6.07 × 1018), and 1.63 × 1019 (3.96 × 1019) for 1018 to 1021 cm−3, respectively.

It was found that the thermal conductivity increases with increasing temperature, doping level, and pressure (Fig. 14) while at a doping level of 1018 cm−3, the increase in thermal conductivity at high temperature and high pressure is exponential.

FIG. 14.

Thermal conductivity in terms of temperature at doping levels of 1018 [(a) and (b)] and 1021 [(c) and (d)] for p-type and the n-type carriers under different pressures.

FIG. 14.

Thermal conductivity in terms of temperature at doping levels of 1018 [(a) and (b)] and 1021 [(c) and (d)] for p-type and the n-type carriers under different pressures.

Close modal

The figure of merit (ZT) in terms of temperature for the p- (n-)type carriers at doping levels of 1018 and 1021 under selected pressures is shown in Fig. 15. As seen, the value of ZT at the same temperature decreases with increasing doping levels, and the increasing temperature leads to continuously increasing ZT at the high doping level. Appling pressure at all doping levels also decreases ZT.

FIG. 15.

Figure of merit in terms of temperature at doping levels of 1018 [(a) and (b)] and 1021 [(c) and (d)] for p-type and the n-type carriers under different pressures.

FIG. 15.

Figure of merit in terms of temperature at doping levels of 1018 [(a) and (b)] and 1021 [(c) and (d)] for p-type and the n-type carriers under different pressures.

Close modal

At a doping level of 1018 cm−3, ZT increases up to 100 K, then it remains constant, and finally, at high temperature (about 800 K), it undergoes a decrease. The increase in pressure higher than 13.22 GPa reduces ZT at temperatures of 700, 550, and 500 K. The bipolar effect occurs at temperatures lower than 800 K for high pressures while it occurs at high temperatures (higher than 800 K) at ambient pressure.

An interesting point is that at doping levels of 1018 and 1019 cm−3, the high value of ZT is independent of temperature and pressure in an interval of temperature. With increasing pressure, this interval of temperature becomes smaller than before. The values of p- (n-)type carrier ZT are 0.95 (0.92), and 0.90 (0.85) at 1018 and 1019 cm−3, respectively. As an applicable point, at these interval temperatures, the uncontrolled and unwanted laboratory conditions such as environmental pressure do not affect the thermoelectric performance of Sb2S3.

In the present study, the optoelectronic and thermoelectric properties of Sb2S3 were studied using DFT based on the FP-LAPW and the PBE-GGA functional as the exchange-correlation potential using the WIEN2k package. The results demonstrated that increasing the pressure from 0.00 to 20.4 GPa changes the calculated bandgap from 1.44 to 0.84 eV. Between pressures of 4.82–6.3 GPa, there was a discontinuity due to the isostructural electronic phase transition. Hence, we can tune the bandgap state (direct vs indirect) and gap energy using the applied pressure.

The results also illustrated that pressure increases the maximum polarization inside the compound. A large absorption coefficient in the order of 104 cm−1 is one of the most interesting properties of Sb2S3. Applying pressure results in a red shift of the optical absorption edge and increases the absorption coefficient. By applying pressure, the maximum reflectivity is about 60% at an energy of 3.90 eV (ultraviolet region), and the intensity of reflectivity is increased at high energy (energy >5 eV). The applied pressure also increases the static dielectric constant and static refractive index.

Under ambient conditions and at a doping level of 1018 cm−3, the obtained Seebeck coefficient was 612.6 (−536.09) µV/K, for p- (n-)type carriers. These values are comparable with the experimental values of Sb2S3 and other compounds similar to stibnite. By increasing the doping concentration and pressure, the Seebeck coefficients decrease. It was found that the bipolar effect occurs at low doping concentrations and high pressure. The variation in electrical conductivity is approximately independent of temperature at ambient pressure. At high pressure and 1018 cm−3 doping concentration, the bipolar effect leads to the increase in electrical conductivity. It was found that the values of electrical conductivity for n-type doping carriers were greater than those of the p-type doping carriers. The results also show that the thermal conductivity increases with increasing temperature, doping level, and pressure. The maximum value of ZT is 0.95 (0.92) for p- (n-)type carriers at 1018 cm−3 and under ambient conditions. As an interesting point, it was found that the high value of ZT is independent of temperature and pressure at doping levels of 1018 and 1019 cm−3 and in the mid-range of temperature.

The authors are grateful to P. Blaha (Vienna University of Technology, Austria) for his technical assistance in using the WIEN2k code.

The authors have no conflicts to disclose.

All authors contributed to the study conception and design. Material preparation, data collection, and analysis were performed by M.A., H.A.R.A., and E.G.Ö. The first draft of the manuscript was written by M.A., and all authors commented on previous versions of the manuscript. All authors read and approved the final manuscript.

Maliheh Azadparvar: Conceptualization (equal); Software (equal); Writing – original draft (equal). H. A. Rahnamaye Aliabad: Conceptualization (equal); Supervision (equal); Writing – original draft (equal). Evren Görkem Özdemir: Conceptualization (equal); Investigation (equal); Writing – review & editing (equal).

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
F.
Caruso
,
M. R.
Filip
, and
F.
Giustino
, “
Excitons in one-dimensional van der Waals materials: Sb2S3 nanoribbons
,”
Phys. Rev. B
92
(
12
),
125134
(
2015
).
2.
M.
Ishaq
,
S.
Chen
,
U.
Farooq
,
M.
Azam
,
H.
Deng
,
Z.-H.
Su
,
Z.-H.
Zheng
,
P.
Fan
,
H.-S.
Song
, and
G.-X.
Liang
, “
High open‐circuit voltage in full‐inorganic Sb2S3 solar cell via modified Zn‐doped TiO2 electron transport layer
,”
Sol. RRL
4
(
12
),
2000551
(
2020
).
3.
V.
Janošević
,
M.
Mitrić
,
N.
Bundaleski
,
Z.
Rakočević
, and
I. L.
Validžić
, “
High‐efficiency Sb2S3‐based hybrid solar cell at low light intensity: Cell made of synthesized Cu and Se‐doped Sb2S3
,”
Prog. Photovoltaics
24
(
5
),
704
715
(
2016
).
4.
Z.
Cui
,
K.
Bu
,
Y.
Zhuang
,
M. E.
Donnelly
,
D.
Zhang
,
P.
Dalladay-Simpson
,
R. T.
Howie
,
J.
Zhang
,
X.
, and
Q.
Hu
, “
Phase transition mechanism and bandgap engineering of Sb2S3 at gigapascal pressures
,”
Commun. Chem.
4
(
1
),
125
(
2021
).
5.
A.
Radzwan
,
R.
Ahmed
,
A.
Shaari
, and
A.
Lawal
, “
Ab initio calculations of optoelectronic properties of antimony sulfide nano-thin film for solar cell applications
,”
Results Phys.
15
,
102762
(
2019
).
6.
A.
Radzwan
,
A.
Lawal
,
A.
Shaari
,
I. M.
Chiromawa
,
S. T.
Ahams
, and
R.
Ahmed
, “
First-principles calculations of structural, electronic, and optical properties for Ni-doped Sb2S3
,”
Comput. Condens. Matter
24
,
e00477
(
2020
).
7.
T.
Ben Nasr
,
H.
Maghraoui-Meherzi
,
H.
Ben Abdallah
, and
R.
Bennaceur
, “
Electronic structure and optical properties of Sb2S3 crystal
,”
Physica B
406
(
2
),
287
292
(
2011
).
8.
H.
Yang
,
P.
Boulet
, and
M.-C.
Record
, “
Thermoelectric properties of Sb-S system compounds from DFT calculations
,”
Materials
13
(
21
),
4707
(
2020
).
9.
J.
Grigas
,
J.
Meshkauska
, and
A.
Orliukas
, “
Dielectric properties of Sb2S3 at microwave frequencies
,”
Phys. Status Solidi A
37
(
1
),
K39
K41
(
1976
).
10.
M. S.
Ablova
,
A. A.
Andreev
,
T. T.
Dedegkaev
,
B. T.
Melekh
,
A. B.
Pevtsov
,
N. S.
Shendel
, and
L. N.
Shumilova
, “
Switching effect in Sb2S3
,”
Sov. Phys. Semicond.-USSR
10
(
6
),
629
631
(
1976
).
11.
D. Y.
Yu
,
P. V.
Prikhodchenko
,
C. W.
Mason
,
S. K.
Batabyal
,
J.
Gun
,
S.
Sladkevich
,
A. G.
Medvedev
, and
O.
Lev
, “
High-capacity antimony sulphide nanoparticle-decorated graphene composite as anode for sodium-ion batteries
,”
Nat. Commun.
4
(
1
),
2922
(
2013
).
12.
R.
Kondrotas
,
C.
Chen
, and
J.
Tang
, “
Sb2S3 solar cells
,”
Joule
2
(
5
),
857
878
(
2018
).
13.
S. S.
Chiad
,
A. M. E.
Ibrahim
, and
N. F.
Habubi
, “
Annealing effect on the optical properties of Sb2S3 thin films
,”
J. Univ. Anbar Pure Sci.
5
(
3
),
18
(
2011
).
14.
O.
Savadogo
and
K. C.
Mandal
, “
Low‐cost technique for preparing n‐Sb2S3/p‐Si heterojunction solar cells
,”
Appl. Phys. Lett.
63
(
2
),
228
230
(
1993
).
15.
Z.
Cai
,
C.-M.
Dai
, and
S.
Chen
, “
Intrinsic defect limit to the electrical conductivity and a two‐step p‐type doping strategy for overcoming the efficiency bottleneck of Sb2S3‐based solar cells
,”
Sol. RRL
4
(
4
),
1900503
(
2020
).
16.
O.
Savadogo
and
K. C.
Mandal
, “
Studies on new chemically deposited photoconducting antimony trisulphide thin films
,”
Sol. Energy Mater. Sol. Cells
26
(
1–2
),
117
136
(
1992
).
17.
J.
Escorcia-García
,
D.
Becerra
,
M. T. S.
Nair
, and
P. K.
Nair
, “
Heterojunction CdS/Sb2S3 solar cells using antimony sulfide thin films prepared by thermal evaporation
,”
Thin Solid Films
569
,
28
34
(
2014
).
18.
H.
Maghraoui-Meherzi
,
T.
Ben Nasr
,
N.
Kamoun
, and
M.
Dachraoui
, “
Structural, morphology and optical properties of chemically deposited Sb2S3 thin films
,”
Physica B
405
(
15
),
3101
3105
(
2010
).
19.
N. S.
Yesugade
,
C. D.
Lokhande
, and
C. H.
Bhosale
, “
Structural and optical properties of electrodeposited Bi2S3, Sb2S3 and As2S3 thin films
,”
Thin Solid Films
263
(
2
),
145
149
(
1995
).
20.
B.
Krishnan
,
A.
Arato
,
E.
Cardenas
,
T. K. D.
Roy
, and
G. A.
Castillo
, “
On the structure, morphology, and optical properties of chemical bath deposited Sb2S3 thin films
,”
Appl. Surf. Sci.
254
(
10
),
3200
3206
(
2008
).
21.
Y.
Yu
,
R. H.
Wang
,
Q.
Chen
, and
L.-M.
Peng
, “
High-quality ultralong Sb2Se3 and Sb2S3 nanoribbons on a large scale via a simple chemical route
,”
J. Phys. Chem. B
110
(
27
),
13415
13419
(
2006
).
22.
A.
Radzwan
,
R.
Ahmed
,
A.
Shaari
,
A.
Lawal
, and
Y. X.
Ng
, “
First-principles calculations of antimony sulphide Sb2S3
,”
Malays. J. Fundam. Appl. Sci.
13
(
3
),
285
289
(
2017
).
23.
T.
Ben Nasr
,
H.
Maghraoui-Meherzi
, and
N.
Kamoun-Turki
, “
First-principles study of electronic, thermoelectric and thermal properties of Sb2S3
,”
J. Alloys Compd.
663
,
123
127
(
2016
).
24.
Z.-H.
Ge
,
B.-P.
Zhang
,
P.-P.
Shang
,
Y.-Q.
Yu
,
C.
Chen
, and
J.-F.
Li
, “
Enhancing thermoelectric properties of polycrystalline Bi2S3 by optimizing a ball-milling process
,”
J. Electron. Mater.
40
(
5
),
1087
1094
(
2011
).
25.
Y.
Kawamoto
and
H.
Iwasaki
, “
Thermoelectric properties of (Bi1−xSbx)2S3 with orthorhombic structure
,”
J. Electron. Mater.
43
(
6
),
1475
1479
(
2014
).
26.
S. V.
Ovsyannikov
and
V. V.
Shchennikov
, “
High-pressure routes in the thermoelectricity or how one can improve a performance of thermoelectrics
,”
Chem. Mater.
22
(
3
),
635
647
(
2010
).
27.
N. V.
Morozova
,
I. V.
Korobeinikov
, and
S. V.
Ovsyannikov
, “
Strategies and challenges of high-pressure methods applied to thermoelectric materials
,”
J. Appl. Phys.
125
(
22
),
220901
(
2019
).
28.
N. M.
Alsaleh
,
E.
Shoko
,
M.
Arsalan
, and
U.
Schwingenschlögl
, “
Thermoelectric materials under pressure
,”
Phys. Status Solidi RRL
12
(
8
),
1800083
(
2018
).
29.
J.
Ibáñez
,
J. A.
Sans
,
C.
Popescu
,
J.
López-Vidrier
,
J. J.
Elvira-Betanzos
,
V. P.
Cuenca-Gotor
,
O.
Gomis
,
F. J.
Manjón
,
P.
Rodríguez-Hernández
, and
A.
Muñoz
, “
Structural, vibrational, and electronic study of Sb2S3 at high pressure
,”
J. Phys. Chem. C
120
(
19
),
10547
10558
(
2016
).
30.
E. L.
Da Silva
,
M. C.
Santos
,
P.
Rodríguez-Hernández
,
A.
Muñoz
, and
F. J.
Manjón
, “
Theoretical Study of Pressure-Induced Phase Transitions in Sb2S3, Bi2S3, and Sb2Se3
,”
Crystals
13
(
3
),
498
(
2023
).
31.
I.
Efthimiopoulos
,
C.
Buchan
, and
Y.
Wang
, “
Structural properties of Sb2S3 under pressure: Evidence of an electronic topological transition
,”
Sci. Rep.
6
(
1
),
24246
(
2016
).
32.
P.
Blaha
,
K.
Schwarz
,
G. K. H.
Madsen
,
D.
Kvasnicka
, and
J.
Luitz
, WIEN2k. An augmented plane wave + local orbitals program for calculating crystal properties,
Institute of Materials Chemistry, TU Vienna
,
2011
, http://www.wien2k.at/.
33.
M.
Städele
,
J. A.
Majewski
,
P.
Vogl
, and
A.
Görling
, “
Exact Kohn-Sham exchange potential in semiconductors
,”
Phys. Rev. Lett.
79
(
11
),
2089
(
1997
).
34.
G. K. H.
Madsen
and
D. J.
Singh
, “
BoltzTraP. A code for calculating band-structure dependent quantities
,”
Comput. Phys. Commun.
175
(
1
),
67
71
(
2006
).
35.
P.
Bayliss
and
W.
Nowacki
, “
Refinement of the crystal structure of stibnite, Sb2S3
,”
Z. Kristallogr. - Cryst. Mater.
135
(
1–6
),
308
315
(
1972
).
36.
A.
Kyono
and
M.
Kimata
, “
Structural variations induced by difference of the inert pair effect in the stibnite-bismuthinite solid solution series (Sb, Bi)2S3
,”
Am. Mineral.
89
(
7
),
932
940
(
2004
).
37.
Y.
Wang
,
Y.
Ma
,
G.
Liu
,
J.
Wang
,
Y.
Li
,
Q.
Li
,
J.
Zhang
,
Y.
Ma
, and
G.
Zou
, “
Experimental observation of the high pressure induced substitutional solid solution and phase transformation in Sb2S3
,”
Sci. Rep.
8
(
1
),
14795
(
2018
).
38.
L. F.
Lundegaard
,
R.
Miletich
,
T.
Balic-Zunic
, and
E.
Makovicky
, “
Equation of state and crystal structure of Sb2S3 between 0 and 10 GPa
,”
Phys. Chem. Miner.
30
,
463
468
(
2003
).
39.
J. P.
Perdew
and
M.
Levy
, “
Physical content of the exact Kohn-Sham orbital energies: Band gaps and derivative discontinuities
,”
Phys. Rev. Lett.
51
(
20
),
1884
(
1983
).
40.
H.
Zhang
,
M.
Ge
,
L.
Yang
,
Z.
Zhou
,
W.
Chen
,
Q.
Li
, and
L.
Liu
, “
Synthesis and catalytic properties of Sb2S3 nanowire bundles as counter electrodes for dye-sensitized solar cells
,”
J. Phys. Chem. C
117
(
20
),
10285
10290
(
2013
).
41.
M.
Azadparvar
,
M.
Kheirabadi
, and
H. A.
Rahnamaye Aliabad
, “
Fluorinated derivatives of tetrahydroaltersolanol molecule on COVID-19, HIV, and HTLV protease by DFT and molecular docking approaches
,”
J. Mol. Model.
28
(
11
),
350
(
2022
).
42.
M.
Azadparvar
,
H. A.
Rahnamaye Aliabad
,
E.
Rezaei-Seresht
, and
M.
Mirzaei
, “
Effect of fluorine substitution on the photobiological and electronic properties of resveratrol crystal structure: A first-principles study
,”
J. Photochem. Photobiol., A
429
,
113941
(
2022
).
43.
M.
Schubert
,
T.
Hofmann
,
C. M.
Herzinger
, and
W.
Dollase
, “
Generalized ellipsometry for orthorhombic, absorbing materials: Dielectric functions, phonon modes and band-to-band transitions of Sb2S3
,”
Thin Solid Films
455-456
,
619
623
(
2004
).
44.
Y.
Li
and
N.
Engheta
, “
Supercoupling of surface waves with ε-near-zero metastructures
,”
Phys. Rev. B
90
(
20
),
201107(R)
(
2014
).
45.
S.
Savoia
,
G.
Castaldi
,
V.
Galdi
,
A.
Alù
, and
N.
Engheta
, “
PT-symmetry-induced wave confinement and guiding in ε-near-zero metamaterials
,”
Phys. Rev. B
91
(
11
),
115114
(
2015
).
46.
K.
Zhao
,
J.
Yang
,
M.
Zhong
,
Q.
Gao
,
Y.
Wang
,
X.
Wang
,
W.
Shen
et al, “
Direct polarimetric image sensor and wide spectral response based on quasi‐1D Sb2S3 nanowire
,”
Adv. Funct. Mater.
31
(
6
),
2006601
(
2021
).
47.
J.
Varghese
,
S.
Barth
,
L.
Keeney
,
R. W.
Whatmore
, and
J. D.
Holmes
, “
Nanoscale ferroelectric and piezoelectric properties of Sb2S3 nanowire arrays
,”
Nano Lett.
12
(
2
),
868
872
(
2012
).
48.
J. J.
Gong
,
A. J.
Hong
,
J.
Shuai
,
L.
Li
,
Z. B.
Yan
,
Z. F.
Ren
, and
J.-M.
Liu
, “
Investigation of the bipolar effect in the thermoelectric material CaMg2Bi2 using a first-principles study
,”
Phys. Chem. Chem. Phys.
18
(
24
),
16566
16574
(
2016
).
49.
H.-S.
Kim
,
K. H.
Lee
, and
S.-i.
Kim
, “
Understanding bipolar thermal conductivity in terms of concentration ratio of minority to majority carriers
,”
J. Mater. Res. Technol.
14
,
639
646
(
2021
).
50.
Z.
Chen
,
X.
Zhang
,
J.
Ren
,
Z.
Zeng
,
Y.
Chen
,
J.
He
,
L.
Chen
, and
Y.
Pei
, “
Leveraging bipolar effect to enhance transverse thermoelectricity in semimetal Mg2Pb for cryogenic heat pumping
,”
Nat. Commun.
12
(
1
),
3837
(
2021
).
51.
Z.
Zamanipour
,
E.
Salahinejad
,
P.
Norouzzadeh
,
J. S.
Krasinski
,
L.
Tayebi
, and
D.
Vashaee
, “
The effect of phase heterogeneity on thermoelectric properties of nanostructured silicon germanium alloy
,”
J. Appl. Phys.
114
(
2
),
023705
(
2013
).
52.
B. B.
Nayak
and
H. N.
Acharya
, “
Electrical and thermoelectric properties of Sb2S3 thin films prepared by the dip-dry method
,”
Thin Solid Films
122
(
2
),
93
103
(
1984
).
53.
H. A.
Rahnamaye Aliabad
and
F.
Asadi Rad
, “
Structural, electronic and thermoelectric properties of bulk and monolayer of Sb2Se3 under high pressure: By GGA and mBJ approaches
,”
Physica B
545
,
275
284
(
2018
).
54.
Q.
Chen
,
J.
Chen
,
X.
Xu
,
Z.
Wang
,
Y.
Ding
,
L.
Xiong
,
B.
Zhu
, and
L.
Tao
, “
Morphology optimization of Bi2Se3 thin films for enhanced thermoelectric performance
,”
Cryst. Growth Des.
21
(
12
),
6737
6743
(
2021
).
55.
M.
Guo
,
Z.
Wang
,
Y.
Xu
,
H.
Huang
,
Y.
Zang
,
C.
Liu
,
W.
Duan
,
Z.
Gan
,
S.-C.
Zhang
,
K.
He
,
X.
Ma
,
Q.
Xue
, and
Y.
Wang
, “
Tuning thermoelectricity in a Bi2Se3 topological insulator via varied film thickness
,”
New J. Phys.
18
(
1
),
015008
(
2016
).
56.
Y. X.
Chen
,
K.
Kitahara
, and
T.
Takeuchi
, “
Thermoelectric properties of β-indium sulfide with sulphur deficiencies
,”
J. Appl. Phys.
118
(
24
),
245103
(
2015
).