Open Access Article
Henry I.
Eya
a,
Robert
Patterson
b and
Nelson Y.
Dzade
*a
aThe Department of Energy and Mineral Engineering, The Pennsylvania State University, USA. E-mail: nxd5313@psu.edu
bSchool of Photovoltaic and Renewable Energy Engineering, University of New South Wales, Australia
First published on 22nd October 2025
Strontium zirconium sulfide (SrZrS3) has garnered significant attention for photovoltaic (PV) applications due to its excellent optoelectronic properties, high chemical and moisture stability, and non-toxicity. However, the bandgaps of both the α- and β-phases lie outside the optimum ranges for both single-junction solar cells (SJSCs) and tandem solar cells (TSCs), thereby limiting their applications in PV technologies. In this study, we employed hybrid density functional theory to engineer the band gaps of α- and β-SrZrS3 through substitutional doping. We considered three dopants (Hf, Sn, and Ti) at the Zr site of SrZrS3 with varying doping concentrations and investigated their effects on the structural, electronic, and optical properties of the materials. We found that Sn and Ti doping effectively lowers the band gaps of both α- and β-SrZrS3, whereas Hf doping increases them. For x values up to 0.25, the band gaps of the α-SrZr1−xSnxS3 and α-SrZr1−xTixS3 are within the optimum range for SJSCs, and those of β-SrZr1−xSnxS3 and β-SrZr1−xTixS3 lie within the optimum range for Si/perovskite as well as perovskite/perovskite TSCs. The three dopants exhibited significant effects on the optical properties of both α- and β-SrZrS3, including the absorption coefficient, energy-loss functions, reflectivity, and refractivity spectra. Thermodynamic stability analysis revealed that for both phases, SrZr1−xHfxS3 can be synthesized via exothermic processes, whereas the formation of SrZr1−xTixS3 and SrZr1−xSnxS3 is endothermic and hence, not thermodynamically favorable. Further analysis showed that SrZr1−xTixS3 in both α- and β-phases are stable under thermodynamic equilibrium conditions, whereas SrZr1−xSnxS3 is prone to dissociation into ternary phases (SrZrS3 and SrSnS3), especially at higher doping concentrations. These results show that Ti doping is effective in tuning the band gaps of α- and β-SrZrS3 toward the optimal values for PV applications.
Recently, chalcogenides perovskites (CPs) with the general formula ABX3 (where A is a group II cation (Ba2+, Ca2+, or Sr2+), B a group IV cation (Zr4+, Ti4+, or Hf4+), and X a chalcogen anion (S2− or Se2−)) have emerged as attractive alternatives to OIHPs owing to their high chemical and thermal stability, nontoxicity, earth-abundant composition, and excellent optoelectronic properties.17–20 Several CPs have been predicted through theoretical studies for thin-film solar cells. For instance, Sun et al. predicted CaTiS3, BaZrS3, CaZrSe3, and CaHfSe3 with distorted perovskite structures and showed that they possess ideal band gaps and high optical absorption for solar cell applications.21 Basera et al. simulated sulfide chalcogenide perovskites ABS3 (A = Ba, Ca, and Sr; B = Hf and Sn) using density functional theory (DFT) methods and reported that they possess exciton binding energies larger than those of three-dimensional (3D) inorganic−organic hybrid perovskites.22 CPs have also been successfully synthesized using various techniques. Niu et al. synthesized α-SrZrS3, β-SrZrS3, BaZrS3, Ba2ZrS4, and Ba3Zr2S7 with several structural types, including distorted perovskite, needle-like, and Ruddlesden–Popper phases, and reported that they exhibit high thermal stability at temperatures up to 550 °C.18
SrZrS3 has great potential for optoelectronic and photovoltaic applications. According to previous studies, this material exists in two crystal phases: α-SrZrS3 and β-SrZrS3, both of which are stable under ambient conditions.18–21 These two phases differ significantly in terms of structural and optoelectronic properties. The NH4CdCl3-type phase (i.e., α-SrZrS3) exhibits a needle-like crystal structure, whereas the GdFeO3-type phase (i.e., β-SrZrS3) possesses a distorted perovskite structure. In addition, α-SrZrS3 has a bandgap of approximately 1.53 eV, which is close to the optimum bandgap for SJSCs (∼1.1–1.4 eV),23,24 whereas that of β-SrZrS3 ranges from 1.98 to 2.13 eV, which is much higher than the ideal value for optimum solar absorption.18,19,25,26 However, according to Sun et al., materials with distorted perovskite structures generally exhibit better optical absorption properties than those with needle-like structures,21 indicating that both α- and β-SrZrS3 are promising for PV applications. In addition, the large bandgap of β-SrZrS3 is attractive for use in tandem solar cells. However, the value exceeds the optimal range for maximum efficiency in both Si/perovskite and perovskite/perovskite TSCs.11,13,15,16 Therefore, to ensure optimal PV performance for both α- and β-SrZrS3, it is necessary to tune their band gaps to the optimum values required for PV applications.
The optoelectronic properties of various semiconductors, including perovskites, have been successfully tuned by adding impurity atoms into their crystal structures, a technique known as doping. According to He et al., owing to the tunable crystal structure of perovskites (ABX3), the bandgap can be precisely controlled via doping at the A, B, and/or X sites, resulting in high fluorescence quantum yields and emission in the visible and near-infrared regions.27 Specifically, B-site doping can significantly affect the bandgap and the channel for radiative recombination in perovskites. Meng et al. achieved a 16.48% decrease in the bandgap of BaZrS3 (from 1.76 to 1.47 eV) by replacing ∼10% of Zr with Ti.28 Yu et al. also reported Fe- and Mn-doped BaZrS3 (BaZr1−xMnxS3 and BaZr1−xMnxS3) with band gaps of 1.58 and 1.63 eV for a doping concentration of approximately 5%.29 Wei et al. synthesized Ti-doped BaZrS3 (BaZr1−xTixS3) and reported a decrease in bandgap from 1.78 to 1.51 eV for a doping atom percentage of 4%.30 The bandgap of β-SrZrS3 has also been tuned by doping antimony (Sb) into the Zr site; however, the authors recorded an increase in the bandgap.31 The band gaps and other optoelectronic properties of several other perovskite materials have also been successfully tuned through doping.32–35
In this study, we employed first-principles density functional theory (DFT) calculations to investigate the effects of various dopant cations, including Hf, Ti, and Sn, on the band gaps and optical properties of both SrZrS3 phases. We varied the doping concentration (from 0 to 50 atom%) to verify its effects on the material properties. We also investigated the thermodynamic stability of the pristine material by determining the range of chemical potentials in which the material is stable. Furthermore, we studied the formability and stability of the doped materials by calculating their defect formation and dissociation energies, respectively. Our predicted reduction in band gaps and enhanced optical properties indicate that Sn- and Ti-doped SrZrS3 are promising materials for photovoltaic applications.
The linear optical properties of the materials were determined from the frequency-dependent complex dielectric function ε(w), which is expressed as follows:
| ε(w) = ε1(w) + iε2(w), | (1) |
![]() | (2) |
![]() | (3) |
![]() | (4) |
![]() | (5) |
To explore different doping concentrations, supercells of 1 × 2 × 1 and 2 × 1 × 1 were constructed for the α- and β-phases, respectively, resulting in a total of 40 atoms per cell. Geometrical optimization was performed on the supercells using the PBE function with k-point meshes of 5 × 5 × 3 and 3 × 7 × 5 for the α- and β-phases, respectively. Furthermore, 1, 2, 3, and 4 Zr atoms in the supercells were substituted with dopant atoms (Hf, Sn, and Ti), corresponding to doping concentrations of 12.5%, 25%, 37.5%, and 50% atom%, respectively. Considering that all Zr sites are identical in both structures, the doping sites were randomly selected. For cases where more than one Zr atoms were substituted, we tested and found that when the dopant atoms are far away from each other, the structure is 0.16 eV energetically more stable than when they are closer to each other. Therefore, we ensured that the selected sites for multiple dopants were well spread out uniformly throughout the structures (Fig. S1 and S2). The doped structures were optimized using the same settings as the pristine supercells. The electronic and optical properties were also obtained using HSE06 with the same Hartree–Fock exchange fractions as for the unit cells. Considering the computational cost of the hybrid functional, lower Monkhorst–Pack k-point meshes of 3 × 3 × 1 and 1 × 5 × 3 were used for the DOS, band structure, and optical property calculations.
![]() | ||
| Fig. 1 (a) and (d) Optimized crystal structures, (b) and (e) projected density of states (PDOS), and (c) and (f) band structures of α-SrZrS3 and β-SrZrS3 perovskites. | ||
Comparing the three dopants in α-SrZrS3, the Ti-doped systems generally showed a more significant change in the supercell volume, followed by the Sn-doped systems. The percentage change in volume (ΔV, Table S1) for α-SrZr1−xTixS3 varied from −1.04% to −4.20% as x increased from 0.125 to 0.5 (negative values indicate shrinkages), whereas that for α-SrZr1−xHfxS3 only varied from −0.18% to −0.71%. This order is consistent with the decreasing order of the ionic radii of dopants (rZr > rHf > rSn > rTi). Similarly, for the β-phase, the change in volume for the Ti-doped systems is higher than that of Sn- and Hf-doped systems. However, that of the Hf-doped systems is higher than that of the Ti-doped system. Furthermore, considering the supercell angles of the pure and doped systems, there was no significant distortion of the α-phase. In contrast, the β-SrZrS3 structure was appreciably distorted by the addition of the dopants, especially at high doping concentrations.
In perovskite materials, structural stability and formability are essential considerations. Hence, when considering dopant choices, it is paramount to explore their effects on the structural stability and formability of the parent material. For an ABX3 perovskite system, the stability and formability can be predicted purely based on geometrical considerations (i.e., from the ionic radii of the constituent elements). The formability of a perovskite structure can be predicted by the Goldschmidt tolerance factor t and the octahedral factor u, which are expressed as follows:50–52
![]() | (6) |
![]() | (7) |
As listed in Table S1, the t and u values for the pure SrZrS3 system are 0.856 and 0.391, respectively, indicating that the material can form a distorted perovskite structure. However, owing to the low u value, there is a high tendency for the needle-like NH4CdCl3 structure (i.e., α-SrZrS3) to form, which explains why SrZrS3 exists in both needle-like and distorted perovskite structures. Because Zr4+ is larger than Hf4+, Sn4+, and Ti4+, the substitution of Zr with Hf, Sn, or Ti atoms results in an increase in the t value, which increases linearly with the doping concentration. The highest increase in t was recorded for the Ti-based systems, followed by the Sn-based systems, which agrees well with the decreasing order of their ionic radii. The u-value, on the other hand, decreased monotonically with increasing doping concentration. The SrZr0.5Ti0.5S3 recorded the lowest u-value of 0.360. According to Tiwari et al., materials with u values in the range of 0.36 to 0.39 can be stabilized using mixed B cations or X anions.50 Therefore, the t and u values of all the doped systems are within the appropriate range for the formation of distorted perovskite structures. However, with increasing doping concentration, the tendency to form needle-like structures increases.
Shown in Fig. 3 and Fig. S3 are the band structures of the SrZr1−xHfxS3, SrZr1−xSnxS3, and SrZr1−xTixS3 systems, and Table 1 lists their band gaps, VBM, and CBM. For both α- and β-SrZrS3, the substitution of Zr with Hf showed a negligible effect on the bandgap. The band gaps of both phases increased monotonically as the doping concentration increased from 0 to 50 atom% (i.e., x = 0.5). With 50% replacement of Zr atoms, the band gaps of α- and β-SrZrS3 increased by only 2.95% and 5.33%, respectively. The band structures show that for all x values, the SrZr1−xHfxS3 systems in both the α- and β-phases maintained a direct bandgap, with the VBM and CBM located at the Γ high symmetry point. Both the VBM and CBM were observed to shift to lower energies as x increased from 0.0 to 0.50 (Table 1). Analyses of the PDOS (Fig. 4 and Fig. S4) revealed that, similar to the pristine systems, the valence band (VB) and conduction band (CB) edges of both α- and β-SrZr1−xHfxS3 are dominated by the S-p and Zr-d orbitals, respectively, with the Hf-d orbital contributing negligibly to both VB and CB, which increased slightly with the increasing doping concentration.
![]() | ||
| Fig. 3 Electronic band structures of (a–f) α- and (g–l) β-phase SrZr1−xHfxS3, SrZr1−xSnxS3, and SrZr1−xTixS3 (x = 0.125 and 0.25) materials. | ||
| System | α-SrZrS3 | β-SrZrS3 | ||||
|---|---|---|---|---|---|---|
| Bandgap (eV) | VBM (eV) | CBM (eV) | Bandgap (eV) | VBM (eV) | CBM (eV) | |
| SrZrS3 | 1.56 | 4.02 | 5.58 | 2.03 | 3.31 | 5.34 |
| SrZr0.875Hf0.125S3 | 1.57 | 3.93 | 5.50 | 2.05 | 3.22 | 5.27 |
| SrZr0.75Hf0.25S3 | 1.58 | 3.83 | 5.41 | 2.07 | 3.13 | 5.20 |
| SrZr0.625Hf0.375S3 | 1.59 | 3.73 | 5.32 | 2.10 | 3.04 | 5.14 |
| SrZr0.50Hf0.50S3 | 1.61 | 3.63 | 5.24 | 2.14 | 2.95 | 5.08 |
| SrZr0.875Sn0.125S3 | 1.48 | 3.98 | 5.46 | 1.83/2.04 | 3.28 | 5.11 |
| SrZr0.75Sn0.25S3 | 1.20 | 3.97 | 5.17 | 1.54/1.78 | 3.22 | 4.77 |
| SrZr0.625Sn0.375S3 | 1.24 | 3.95 | 5.19 | 1.54/1.62 | 3.19 | 4.77 |
| SrZr0.50Sn0.50S3 | 1.30 | 3.87 | 5.17 | 1.44/1.48 | 3.15 | 4.59 |
| SrZr0.875Ti0.125S3 | 1.43 | 4.01 | 5.44 | 1.77 | 3.38 | 5.15 |
| SrZr0.75Ti0.25S3 | 1.41 | 4.12 | 5.52 | 1.59 | 3.44 | 5.03 |
| SrZr0.625Ti0.375S3 | 1.35 | 4.16 | 5.51 | 1.79 | 3.38 | 5.17 |
| SrZr0.50Ti0.50S3 | 1.29 | 4.20 | 5.49 | 1.70 | 3.44 | 5.14 |
![]() | ||
| Fig. 4 Projected density of states (PDOS) of (a–f) α- and (g–l) β-phase SrZr1−xHfxS3, SrZr1−xSnxS3, and SrZr1−xTixS3 (x = 0.125 and 0.25). | ||
For the SrZr1−xSnxS3 systems, a significant decrease in bandgap was observed in both α- and β-phases compared with that of the pristine systems. With the replacement of only 12.5% Zr atoms, the bandgap of α-SrZrS3 decreased by 5.26% (from 1.56 to 1.48 eV), and that of β-SrZrS3 decreased by 9.72% (from 2.03 to 1.83 eV). With a further increase in dopant concentration to 25%, the band gaps of both phases decreased; however, they increased again at higher dopant concentrations.
For both phases, the CB and VB of the Sn-doped systems are dominated by the S-p and Zr-d orbitals, respectively (Fig. 4 and Fig. S4). However, the Sn-s orbital contributes significantly to the edge of the CB, increasing band overlap and lowering the CBM, resulting in a decrease in the bandgap. These results are consistent with those reported for Sn-doped BaZrS3, where a decrease in bandgap from 1.55 to 0.96 eV was recorded as 25% of the Zr atoms were replaced with Sn atoms.55 According to the authors, the decrease in the bandgap of the Sn-doped BaZrS3 system is attributed to the higher electronegativity of Sn than Zr, resulting in short covalent bonds between Sn-s and S-p orbitals. In addition, the contribution of the Sn-s orbitals to the CB edge can enhance charge-carrier transport in the material, which is beneficial for PV applications.55 Notably, indirect band gaps were observed for the β-SrZr1−xSnxS3 systems, suggesting that solar cells fabricated with these indirect bandgap materials need to be thicker to provide more opportunities for photoemission and transition to take place. Compared to direct bandgap materials, the formation of an indirect bandgap for the β-SrZr1−xSnxS3 systems could result in lower light absorption efficiency, a large recombination lifetime, and therefore a considerable diffusion length.56,57
Ti doping showed a favorable decrease in the band gaps of both α- and β-SrZrS3. With a doping concentration of 12.5%, the bandgap of α-SrZrS3 decreased to 1.43 eV, which is close to the optimum range for SJSCs. With a further increase in doping concentration, the bandgap decreased monotonically, reaching 1.35 eV at 37.5%. A similar trend was observed for β-SrZrS3. However, beyond 25% doping, the bandgap of β-SrZr1−xTixS3 increased with increasing doping concentration. The overall decrease in band gaps (compared to that of the pristine materials) recorded for α- and β-SrZr1−xTixS3 is attributed to the lower energy of the Ti-3d orbital than that of the Zr-4d orbital.30,48 As a result, the Ti-d orbital dominates the band edge of the CB, as revealed by the PDOS (Fig. 4 and Fig. S4), resulting in the lowering of the CBM (Table 1). These results agree well with previous DFT results reported for Ti-doped BaZrS3.48 Khan et al. also reported a lowering of the CB edge of Ti-doped MAPbBr3, resulting in a decrease in the bandgap energy.58 Therefore, Ti-doping is effective in lowering the bandgap energy of perovskite materials.
Fig. 5 summarizes the band gaps of the pristine and doped SrZrS3 systems. The blue- and green-shaded regions indicate the optimum bandgap ranges for the maximum efficiency of SJSCs (1.10–1.40 eV) and TSCs (1.65–1.90 eV), respectively.13,15,23,24 The band gaps of α-SrZr1−xSnxS3 and α-SrZr1−xTixS3 with x values up to 0.25 lie well within the optimum range for SJSCs. Although the bandgap of α-SrZr0.875Ti0.125S3 lies slightly above this range, the value is sufficient to yield a power conversion efficiency of up to 30% in SJSCs.23 Furthermore, except for an x value of 0.25, the band gaps of β-SrZr1−xSnxS3 and β-SrZr1−xTixS3 lie within the optimum range for the maximum performance of perovskite-based TSCs. According to the theoretical efficiency limit for Si/perovskite TSCs, considering a bandgap of 1.1 eV for Si, the ideal bandgap of a top perovskite cell is approximately 1.73 eV,13 which is very close to 1.70 recorded for β-SrZr0.5Ti0.5S3. For the Hf-doped systems, the band gaps for all doping concentrations lie outside the optimum range for both SJSCs and TSCs. However, at a high doping concentration, the α-SrZr1−xHfxS3 may be considered a top layer for TSCs, especially perovskite/perovskite TSCs.
![]() | ||
| Fig. 5 Variation of the band gaps of α- and β-SrZr1−xHfxS3, SrZr1−xSnxS3, and SrZr1−xTixS3 with the doping concentration x. | ||
Effective masses of holes
and electrons
in the vicinity of the Fermi level are also vital properties of PV materials as they play a key role in determining the charge transport characteristics in the material. The mobility of photogenerated electron–hole pairs, which highly influences the photoactivity of a material, is dependent on
and
, as shown in the following equation:59,60
![]() | (8) |
The hole and electron effective masses are determined by the dispersion of the VB and CB edges of the band structures. They are inversely proportional to the curvature of the corresponding valence and conductance bands and can be expressed as follows:59,61
![]() | (9) |
Table S2 lists the calculated hole and electron effective masses along the Γ–X, Γ–Y, and Γ–Z, and Γ–Z high symmetry directions in the Brillouin zone. Pristine α-SrZrS3 exhibits low
and
in the Γ–X and Γ–Y directions but high
and
in the Γ–Z, indicating higher carrier mobility in Γ–X and Γ–Y directions. The pristine β-SrZrS3, on the other hand, exhibited
and
less than 1 in the three directions. The calculated values are comparable to those reported by Xue et al. for β-SrZrS3.31 For α-SrZrS3, Hf doping reduced
in the Γ–X and Γ–Z directions but increased it in the Γ–Y direction, whereas the
in the Γ–X and Γ–Y directions decreased with increasing doping concentrations, but that in the Γ–Z direction increased. A similar trend was observed for the β-phase, where
and
in all the Γ–X and Γ–Y directions decreased but those in the Γ–Z direction increased with increasing doping concentrations. Considering the Sn-doped systems, both
and
in the three directions generally increased, indicating a decrease in both hole and electron mobility in Sn-doped SrZrS3. Ti doping reduced the
of both α- and β-SrZrS3 in the Γ–X and Γ–Y directions in but increased it in the Γ–Z direction. On the other hand, for both phases,
in the three directions increased with Ti doping. This indicates that Ti doping may enhance hole mobility but reduce electron mobility in α- and β-SrZrS3.
To gain insight into the recombination rate of charge carriers, we calculated the ratios of
to
(denoted here as D in Table S2) in the Γ–X, Γ–Y, and Γ–Z directions. A high value of D indicates a greater difference between the hole and electron mobilities, resulting in the low recombination of electron–hole pairs.62 For both the pristine and doped α-SrZrS3, low D values were generally recorded in the Γ–X direction, whereas the relatively high D values were observed in the Γ–Y and Γ–Z directions. Conversely, for the β-phase, low D values were observed in the Γ–Z direction, whereas higher values were obtained in the Γ–X and Γ–Y directions. These results show that high charge-carrier recombination is expected in the Γ–X and Γ–Z high symmetry directions of α- and β-SrZrS3, respectively.
The dielectric function ε(ω), which comprises the real (ε1) and imaginary (ε2) parts, is a vital property because other optical properties of a material are determined from it. The imaginary part of the dielectric function indicates the radiation absorbed by the material, and the real part indicates the energy gain due to the scattering of the incident light.55 The imaginary part also describes the absorptive behavior of a material, whereas the real part describes the degree of polarization.63,64 A high dielectric constant is desired in PV applications as it promotes effective charge screening and decreases charge-carrier recombination.60Fig. 6 shows the real and imaginary parts of the dielectric function of the pristine and doped α- and β-SrZrS3, and Fig. S6 shows the variation of their static dielectric functions with x values. The calculated static dielectric constant of the pristine α-SrZrS3 is approximately 6.42. Substituting one Zr atom with Hf, this value first increased to approximately 7.13 (Fig. 6a). However, with a further increase in the doping concentration, the dielectric constant slightly decreased, with SrZr0.5Hf0.5S3 exhibiting a dielectric constant of approximately 7.04. Sn-doping generally increased the static dielectric function of pristine α-SrZrS3; however, the value tends to decrease at higher doping concentration (Fig. 6b). For Ti-doped α-SrZrS3, with the substitution of 12.5% of the Zr atoms with Ti (i.e., x = 0.125), the static dielectric function first slightly decreased (from 6.42 to 6.40). However, as the doping concentration further increased, the static dielectric function increased monotonically, reaching a value of 7.92 at an x value of 0.50 (Fig. 6c). Considering the imaginary part of the dielectric function, the main peak for the SrZr1−xHfxS3 system increased significantly at a photon energy of approximately 4 eV (Fig. 6d), indicating increased absorption of radiation. For the SrZr1−xSnxS3 and SrZr1−xTixS3 systems, in addition to the increase in the peak intensity, the peaks were red-shifted, which corresponds to the decrease in the bandgap energies (Fig. 6(e) and (f)).
![]() | ||
| Fig. 6 Real and imaginary parts of the dielectric functions of (a–f) α- and (g–l) β-phase SrZr1−xHfxS3, SrZr1−xSnxS3, and SrZr1−xTixS3. | ||
For the β-phase, the static dielectric function of SrZr1−xHfxS3 decreased monotonically from 7.78 to approximately 7.28 as the doping concentration increased from 0% to 50% (Fig. 6g). A similar result was observed for the β-SrZr1−xSnxS3 system, with the static dielectric function decreasing from 7.78 to 7.24 as x increased from 0 to 0.5 (Fig. 6h). Doping with Ti also generally increased the dielectric constant of β-SrZr1−xTixS3, except for the case of x = 0.25, where a slight decrease in the dielectric constant was observed (Fig. 6i). The reduction in the dielectric constant of β-SrZr1−xSnxS3 is consistent with the results reported for Sn-doped BaZrS3 in the distorted perovskite structure.48,55 Considering the imaginary part of the dielectric function, Hf and Sn doping generally shifted the region at which the material starts to absorb radiation to lower energies (red-shift) and significantly decreased the intensity of the main peak (Fig. 6j and k). The imaginary dielectric function of the β-SrZr1−xTixS3 were also red-shifted, but the intensities of the peaks were higher than that of the pristine materials, except in the energy range of approximately 3.5–4.5 eV, where the intensity of the pristine material is significantly higher than that of the doped systems.
From the dielectric constants, we simulated other optical properties of the pristine and doped SrZrS3 systems, including the absorption coefficient, energy loss function, reflectivity, and refractivity spectra in the visible range. The absorption coefficient measures how far light of a specific wavelength can penetrate a material before being absorbed, thereby determining the light-harvesting power of a material.26,65 A high absorption coefficient results in a small diffusion length, making it possible to achieve high power conversion efficiency using a thin absorber layer (<1 μm).21Fig. 7 shows the absorption coefficients of the pristine and doped α- and β-SrZrS3. Hf doping blue-shifted the absorption edge of both materials. It also increased the absorption of α-SrZrS3 significantly in the energy range of approximately 3.5 to 4.3 eV (Fig. 7a). The main absorption peak of α-SrZrS3 reached a maximum intensity of approximately 6.5 × 105 cm−1 at 4.2 eV, whereas those of α-SrZr1−xHfxS3 reached a maximum value of 8.64 × 105 cm−1 (for x = 0.5). However, further into the UV range (above photon energy of 4.5 eV), the absorption coefficients of the α-SrZr1−xHfxS3 decreased below that of the pristine material. The absorption spectra of the α-SrZr1−xSnxS3 and α-SrZr1−xTixS3 systems also increased, but less significantly than that of α-SrZr1−xHfxS3. The two systems, especially at x values up to 0.25, start to absorb photons at energies much lower than that of the pristine α-SrZrS3, which corresponds to their reduced band gaps. α-SrZr0.875Sn0.125S3 and α-SrZr0.75Ti0.25S3 recorded maximum absorbance of 6.9 × 105 and 7.3 × 105 cm−1 at photon energies of approximately 4.1 and 3.9 eV, respectively.
![]() | ||
| Fig. 7 Absorption coefficient and energy loss functions of (a–f) α- and (g–l) β-phase SrZr1−xHfxS3, SrZr1−xSnxS3, and SrZr1−xTixS3. | ||
For the β-SrZrS3 systems, the three dopants generally reduced the absorption spectra of the material. In agreement with their increased band gaps, the commencement of absorption for β-SrZr1−xHfxS3 blue-shifted, whereas that for SrZr1−xSnxS3 and SrZr1−xTixS3 red-shifted, which is consistent with their lower band gaps. For all x values, the intensity of the main peak of the pristine material remained higher than that SrZr1−xHfxS3 and SrZr1−xSnxS3 in the lower energy range. On the other hand, the intensity of the Ti-doped system was higher than that of the pristine material in the lower energy range. These results show that Hf and Sn doping improve the absorption spectra of α-SrZrS3 but reduce that of β-SrZrS3, whereas Ti doping improves the absorption spectra of both phases.
Fig. 7(d)–(f) and (j)–(l) show the energy loss functions of the α- and β-SrZrS3 systems with and without dopants. The energy loss function represents the energy loss of fast electrons as they traverse a material.63 For the α-phase, Hf doping significantly increased the energy loss spectra in the energy range of 4.4–5.0 eV, indicating an increase in energy loss of fast electrons in the Hf-doped material. On the other hand, Sn and Ti doping only slightly increased the energy loss spectra in the entire energy range, indicating that both dopants slightly increase the energy loss of fast electrons traversing through the material. For the β-phase, the three dopants, especially Hf and Sn, reduced the energy loss spectra of the material, thereby improving its performance in PV applications.
The reflectivity of a material is also an essential property to consider when evaluating it for PV applications. It measures the amount of incident photons lost to reflection. Lower reflectivity indicates a lower loss of photons, which is desirable for PV absorber layers. Fig. S5 shows the reflectivity of the pristine and doped α- and β-SrZrS3. Hf and Sn doping generally increased the reflectivity of the α-phase in the low photon-energy range (Fig. S5a and b). The static reflectivity of α-SrZr1−xHfxS3 increased from approximately 0.188 to 0.206 as the x increased from 0 to 0.125 (Fig. S5a). With a further increase in x, the static reflectivity decreased slightly, reaching a value of 0.204 at x = 0.50. However, for all x values, the reflectivity of α-SrZr1−xHfxS3 remained higher than that of the pristine material in the infrared and visible regions. The reflectivity significantly increased in the ultraviolet region (3.8–4.4 eV), reaching a maximum value of approximately 0.39 at an energy value of 4.17 eV (for SrZr0.875Hf0.125S3). For the Sn-doped system, α-SrZr0.75Sn0.25S3 recorded the highest static reflectivity of 0.22, which is 0.032 higher than that of the pure α-SrZrS3 (Fig. S5b). The Ti-doped material showed a slight decrease in reflectivity in the entire plotted energy range when one Zr atom was replaced with a Ti atom (i.e., x = 0.125); however, as doping concentration increased from 0.125 to 0.50, the reflectivity increased monotonously, SrZr0.5Ti0.5S3 exhibiting the highest static reflectivity of 0.22 (Fig. S5c). These systems maintained high reflectivity across the entire infrared and visible regions; however, their reflectivity decreased below that of the pristine material at energy values exceeding 3.8 eV. For the β-phase, Hf and Sn doping showed a monotone decrease in reflectivity in the infrared and visible regions, SrZr0.5Hf0.5S3 and SrZr0.5Sn0.5S3 recording static reflectivity values of 0.206 and 0.210 (Fig. S5g and h), which are 0.017 and 0.013 lower than that of the pristine material (0.223), respectively. On the other hand, similar to the case of α-SrZrS3, the reflectivity of β-SrZrS3 first slightly decreased as the Ti-doping concentration increased from 0 to 0.125, after which it increased with further increase in x (Fig. S5i). These results show that Hf and Sn doping generally reduce the reflectivity of β-SrZrS3 but increase that of α-SrZrS3, whereas Ti reduces the reflectivity of both SrZrS3 phases at low doping concentrations but increases it at high doping concentrations.
Fig. S5(d–f) and (j–l) show the refractive spectra of the undoped and doped SrZrS3, which measure the degree to which photons are slowed down in the materials due to their interactions with electrons and other factors.63 Low refractive indices indicate low photon diffraction in a material, which is beneficial in PV applications. For both phases and the three dopants, the refractive spectra followed the same trend as the reflectivity. Hf and Sn increased the refractivity of the α-phase but reduced that of the β-phase, whereas Ti doping reduced the refractivity of both phases at low concentrations but increased it at higher concentrations. These results indicate that Hf and Sn doping generally enhance the absorption coefficient of α-SrZrS3; however, they also increase its energy loss, reflectivity, and refractivity. In contrast, the two dopants reduce the energy loss, reflectivity, and refractivity of β-SrZrS3 as well as the absorption coefficient. On the other hand, Ti doping increases the absorption coefficient and energy function of α-SrZrS3 and slightly lowers those of β-SrZrS3. However, for both phases, it lowers the reflectivity and refractive spectra only at a low doping concentration.
During the synthesis of ternary systems like SrZrS3, the formation of elemental (Sr, Zr, and S) and binary (such as SrS, SrS2, and ZrS) phases always competes with the desired ternary phase. In addition to these, ternary and possibly quaternary phases could compete with the formation of quaternary systems, such as SrZr1−xHfxS3, SrZr1−xSnxS3, and SrZr1−xTixS3. This competition can be quantitatively described by the limit to the chemical potentials of the component elements.67 Chemical potentials represent the energy of the reservoirs with which atoms are exchanged.68 They can be defined as the partial molar derivative of the Gibbs free energy with respect to elemental species. Under constant temperature and pressure, the chemical potential determines the stability of substances and their tendency to chemically react to form new substances, transform into new physical states, or migrate from one spatial location to another.69 It is, therefore, necessary to predict the range of chemical potentials of the constituent elemental species of a material over which the target phase, rather than the elemental species or competing phases, is stable. This can be achieved by comparing the free energy of the material with that of all competing phases, including those consisting of subsets of the elemental species in the material.70
To synthesize a ternary-phase material with the chemical formula AxByCz under equilibrium conditions, the chemical potential of the constituent elements (i.e., μA, μB, and μC) in the synthesis environment must satisfy the following conditions:67,71
| xμA + yμB + zμC = ΔHf(AxByCz), | (10) |
| mμA + nμB < ΔHf(AmBn). | (11) |
This applies to ternary or quaternary phases.
Herein, we considered the thermodynamic stability of SrZrS3. We screened the Materials Project database to identify SrZrS3 competing phases whose energies above Hull are less than 0.01 eV per atoms, which include the elemental phases (Sr, Zr, and S) and binary phases (SrS, SrS3, ZrS, ZrS2, ZrS3, Zr2S, Zr3S4, and Zr9S2). Therefore, based on eqn (11), stable SrZrS3 can be synthesized without the formation of the competing phases if the following constraints are satisfied:
| μSr < 0 |
| μZr < 0 |
| μS < 0 |
| μSr + μS < ΔHf(SrS) |
| μSr + 3μS < ΔHf(SrS3) |
| μZr + μS < ΔHf(ZrS) |
| μZr − 2μS < ΔHf(ZrS2) |
| μZr + 3μS < ΔHf(ZrS3) |
| 2μZr + μS < ΔHf(Zr2S) |
| 3μZr + 4μS < ΔHf(Zr3S4) |
| 9μZr + 2μS < ΔHf(Zr9S2) |
![]() | ||
| Fig. 8 (a) Thermodynamically stable region of SrZrS3 in the Sr–Zr–S system. (b) Substitutional defect formation energies of α- and β-SrZr1−xHfxS3, SrZr1−xSnxS3, and SrZr1−xTixS3. | ||
As shown in Fig. 8(a), the chemical stability window for SrZrS3 is a trapezium bound by four coexistence curves AB, AC, BD, and CD, where SrZrS3 coexists with Zr3S4, ZrS2, SrS, and ZrS3, respectively. This is consistent with the experimental study reported by Lee et al., in which small amounts of ZrS3, ZrS2, and SrS were observed to coexist with β-SrZrS3 synthesized via solid-state reactions.20 Bystrický et al. also reported the coexistence of ZrS2 and ZrO2 with β-SrZrS3 synthesized via the solid-state sulfurization of Sr and ZrO2.73 The vertices B and D represent Zr-rich/S-poor and S-poor/Zr-rich regions, respectively. Lines AB and CD also represent Zr-rich/S-poor and S-poor/Zr-rich regions, indicating that Zr3S4 and ZrS3 could be precipitated under Zr-rich/S-poor and S-poor/Zr-rich conditions, respectively. On the other hand, lines AC and AC extend from Zr-rich/S-poor to S-poor/Zr-rich regions, indicating that the precipitation of SrS and ZrS2 is not an indicator of the chemical potential limits of Zr and S.74 Overall, this chemical potential diagram shows that SrZrS3 is stable, which is consistent with previous experimental studies, where both SrZrS3 phases have been successfully synthesized.19,20,73,75 However, the stability window is narrow, indicating the need for adequate control of the synthesis environment to prevent the synthesis of the competing phases.
Next, to determine whether the doped systems can be synthesized, we calculated the defect formation of the dopants in SrZrS3 using the supercell approach.68 Based on this approach, the formation energy Ef of an intrinsic or extrinsic point defect X in a material can be calculated using the following formula:76,77
![]() | (12) |
Fig. 8(b) and Table S3 show the formation energies of the SrZr1−xHfxS3, SrZr1−xSnxS3, and SrZr1−xTixS3 systems in the α- and β-phases. For both phases, the formation energies of SrZr1−xHfxS3 were less than zero and decreased linearly with increasing x value, whereas those of SrZr1−xSnxS3 and SrZr1−xTixS3 were greater than zero and increased linearly with x.
These results show that SrZr1−xHfxS3 can be synthesized through an exothermic process, whereas SrZr1−xSnxS3 or SrZr1−xTixS3 can only be formed through endothermic processes. For SrZr1−xHfxS3, the defect formation energies of the β-phase are slightly lower than those of the α-phase for all x values, indicating that it is easier to dope Hf into β-SrZrS3 than into α-SrZrS3. In contrast, for all values of x, the calculated defect formation energies of SrZr1−xSnxS3 and SrZr1−xTixS3 in the α-phase are lower than those in the β-phase, indicating than Sn and Ti doping of α-SrZrS3 is more energetically favorable than that of β-SrZrS3. The very high defect formation energies of the SrZr1−xTixS3 systems indicate that high energy is required to dope SrZrS3 with Sn at the Zr site.
To further determine the stability of the doped systems, we calculated the energies required for the systems to dissociate into secondary binary phases, such as SrS, ZrS3, and MS2 (M = Hf, Sn, or Ti), or ternary phases, such as SrZrS3 and SrMS3. The energy required to dissociate SrZr1−xMxS3 into binary and ternary phases can be calculated as follows:55
| Ebd = xET(SrZrx−1MxS3) − ET(SrS) − (1 − x)ET(ZrS2) − xET(MS2), | (13) |
| ETd = ET(SrZrx−1MxS3) − (1 − x)ET(SrZrS3) − xET(SrMS3), | (14) |
Fig. 9 and Table S3 show the calculated dissociation energies of SrZr1−xHfxS3, SrZr1−xSnxS3, and SrZr1−xTixS3 in both the α- and β-phases. For both phases, Ebd for all the systems is less than zero, except for β-SrZr0.5Sn0.5S3, where Ebd is slightly greater than zero (0.0399 eV). This indicates that the dissociation of the doped systems into binary secondary phases is not thermodynamically favorable. As shown in Fig. 9(a) and (b), Ebd increases nearly linearly with x, indicating that at higher doping concentrations, the systems become more susceptible to dissociation into binary phases. Considering the ternary dissociation (Fig. 9(a) and (b)), for all values of x, the ETd values of SrZr1−xTixS3 in both phases are negative and decrease significantly with increasing x value, indicating that under equilibrium conditions, SrZr1−xTixS3 is more stable than the competing ternary phases. In contrast, the dissociation energy of both α- and β-SrZr1−xSnxS3 is positive and increases with increasing x value, indicating its tendency to dissociate into the ternary phases under thermodynamic equilibrium conditions. For SrZr1−xHfxS3, the ETd values in both α- and β-phases are very close to zero. In the α-phase, it decreases from 0.0004 to −0.0007 eV as x increases from 0.125 to 0.50, indicating that the system would dissociate into ternary phases at low Hf doping concentrations. For the β-phase, the ETd values are all negative and slightly increase with the x value, indicating an increasing tendency to dissociate into ternary phases. These results show that Ti-doped SrZrS3 is more stable than Hf- and Sn-doped SrZrS3 under thermodynamic equilibrium conditions.
Supplementary information (SI) contains the optimized structures, band structures, projected density of states (PDOS), optical spectra (reflectivity and refractivity), and a summary of structural parameters, effective masses, and formation energies of the α- and β-phases of SrZr1−xHfxS3, SrZr1−xSnxS3, and SrZr1−xTixS3. See DOI: https://doi.org/10.1039/d5cp02689j.
| This journal is © the Owner Societies 2025 |