Open Access Article
Rifat Rafiua,
Karim Kriaab,
Md. Sakib Hasan Saikota,
Md. Azizur Rahman
*c,
Imtiaz Ahamed Apon
d,
N. Sfinae,
Mohd Taukeer Khanf,
Noureddine Elboughdiri
g,
S. AlFaifyh and
Imtiaz Ahmmed Talukderc
aDepartment of Material Science and Engineering, Khulna University of Engineering & Technology (KUET), Khulna 9203, Bangladesh
bImam Mohammad Ibn Saud Islamic University (IMSIU), Riyadh 11432, Saudi Arabia
cDepartment of Electrical and Electronic Engineering, Begum Rokeya University, Rangpur, 5400, Bangladesh. E-mail: azizurrahmanatik49@gmail.com
dDepartment of Electrical and Electronic Engineering, Bangladesh Army University of Science and Technology (BAUST), Saidpur 5311, Bangladesh
eApplied College at Mohayel Assir, King Khalid University, Abha 61413, Saudi Arabia
fDepartment of Physics, Faculty of Science, Islamic University of Madinah, Al Jamiah, Madinah, 42351, Saudi Arabia
gChemical Engineering Department, College of Engineering, University of Ha'il, P. O. Box 2440, 81441, Ha'il, Saudi Arabia
hDepartment of Physics, College of Science, King Khalid University, POST BOX 960, Abha, AlQura'a 61421, Saudi Arabia
First published on 17th November 2025
The quest for efficient and durable absorber materials has steered attention toward vacancy-ordered double perovskites, which exhibit tunable band gaps and strong optical absorption, making them promising candidates for next-generation solar cell architectures. In particular, lead-free vacancy-ordered halide double perovskites have emerged as viable alternatives to toxic Pb-based counterparts. In this study, we systematically investigate the structural, electronic, charges density, mechanical, optical, phonon stability, molecular dynamics (MD), population analysis and photovoltaic properties of D2CeX6 (D = Ga, In, Tl; X = Cl, Br) compounds by employing first-principles calculations in conjunction with SCAPS-1D device simulations. Our results reveal that all compounds crystallize in a stable cubic phase with negative formation enthalpies, confirming their thermodynamic stability. Within GGA–PBE, the calculated direct band gaps are 1.733 eV (Ga2CeCl6), 1.276 eV (Ga2CeBr6), 1.555 eV (In2CeCl6), 0.859 eV (In2CeBr6), 1.755 eV (Tl2CeCl6), and 1.364 eV (Tl2CeBr6), placing all but In2CeBr6 within or near the optimal 1.1 to 1.8 eV range for single-junction solar cells. HSE06 hybrid functional results yield wider gaps of 1.776 eV, 2.843 eV, 2.261 eV, 2.170 eV, 3.632 eV, and 1.418 eV, respectively, suggesting suitability for both single-junction and tandem architectures. Optical analyses demonstrate high absorption coefficients (>105 cm−1), strong dielectric responses, and large refractive indices, particularly in In2CeBr6 and Tl2CeBr6. Mechanical evaluations confirm ductile behavior, with Tl2CeBr6 and In2CeBr6 exhibiting superior stiffness and near-isotropic mechanical stability. Molecular dynamics simulations (NPT ensemble, 50 ps) at room temperature confirm the excellent thermal robustness of the studied compounds. The phonon dispersion analysis further indicates full dynamical stability for Ga2CeCl6, In2CeCl6, and Tl2CeCl6, while Ga2CeBr6, In2CeBr6, and Tl2CeBr6 exhibit minor soft or near-zero modes that are likely stabilized by finite-temperature effects. Device-level simulations predict power conversion efficiencies (PCE) up to 25.29% for Ga2CeBr6, with In2CeCl6 and Tl2CeBr6 also exceeding 24%. These findings position the D2CeX6 family, especially bromide-based compounds, as promising candidates for efficient and stable lead-free perovskite solar cells.
Previous studies have provided important insights into related systems. For example, A. Jabar et al. investigated Cs2CeCl6 using LSDA + mBJ in Wien2k, reporting a semiconducting bandgap of ∼1.828 eV along with its optical and thermoelectric behavior, though photovoltaic device-level performance was not assessed.10 Earlier, Kaatz and Marcovich (in the year of 1966) established that Cs2CeCl6 crystallizes in a cubic K2PtCl6-type structure (Pm3m), where Ce4+ is octahedrally coordinated by Cl− ions and Cs+ is twelve-fold coordinated, offering a useful structural reference for related Ce-based halide perovskites.11 Meanwhile, hybrid organic–inorganic Pb-halide perovskites have achieved record PCEs of up to 25.5%,12 but their deployment is hampered by inherent toxicity and poor environmental durability. Building on these motivations, we turned our attention to Ce-based vacancy-ordered double perovskites D2CeX6 (D = Ga, In, Tl; X = Cl, Br). Similar investigations, such as that of Redi Kristian Pingak et al., confirmed the structural stability and semiconducting behavior of Tl2SnCl6 and Tl2SnBr6, with Tl2SnBr6 (1.48 eV) identified as a promising absorber due to its strong visible-light response.13 Beyond photovoltaics, Ce-based oxides and halides demonstrate intriguing functionalities. Mehwish K. Butt et al. reported ferromagnetic half-metallicity in CeAlO3 arising from Ce-4f states, suggesting potential in spintronic and optical applications.14 Likewise, Mariya Zhuravleva et al. showed that Ce-halides such as Cs3CeCl6, Cs3CeBr6, CsCe2Cl7, and CsCe2Br7 exhibit high light yields and fast decay times, underlining their promise as scintillators for radiation detection.15 Recent studies have shown that vacancy-ordered CsRbGeCl6 and CsRbGeBr6 possess direct band gaps of 2.60 eV and 1.78 eV, respectively, at the Γ-point, making them highly effective for light absorption across the visible to UV spectrum. Their excellent structural stability, non-toxic composition, and favorable electronic properties make these materials promising candidates for photovoltaic and other optoelectronic applications.16 Similarly, Rb2GeCl6 and Rb2GeBr6 have been identified as lead-free alternatives with direct band gaps, mechanical robustness, and thermodynamic stability, making them suitable for renewable energy applications.17 Further, N. V. Skorodumova et al. demonstrated that Ce-4f states in CeO2 and Ce2O3 play decisive roles in governing electronic structure and bonding.18 Tianyu Tang et al. reported that vacancy-ordered double halide perovskites A2BX6 (A = In, Tl; B = Pd, Pt; X = Cl, Br, I) are structurally stable semiconductors with tunable band gaps, strong UV-IR absorption, and potential applications as solar absorbers (Tl2PdBr6) or transparent conductors (In2PtCl6, Tl2PtCl6), yet their device-level performance and practical applicability remain unexplored.19
Additionally, vacancy-ordered halide perovskites A2BX6 (A = In, Tl; B = Pd, Pt; X = Cl, Br, I) offer extensive tunability of their band gaps (from 0.328 eV to 3.238 eV in HSE functional) and exhibit pronounced optical responses across the visible to infrared spectrum. Compounds such as Tl2PdBr6 and In2PtCl6 are emerging as leading candidates for solar absorber layers and transparent conductors, respectively, owing to their superior absorption coefficients and direct or quasi-direct band structures.20–22 Similarly, Showkat Hassan Mir et al. investigated Cerium-based double perovskites A2CeX6 (A = Cs, K; X = Cl, Br) and found them to be structurally stable, non-magnetic semiconductors with direct band gaps, anisotropic mechanical and electronic properties, and strong visible-light absorption.23 Taken together, these studies suggest that incorporating Ce into vacancy-ordered double perovskites can yield a rich interplay of structural stability, optical absorption, and electronic diversity.
The present investigation focuses on the performance analysis of D2CeX6 (D = Ga, In, Tl; X = Cl, Br) double perovskite compounds, carried out through a combination of first-principles calculations within the CASTEP framework and SCAPS-1D device simulations. The study comprehensively examines their structural, electronic, optical, thermoelectric, and mechanical properties, phonon stability, molecular dynamics, which are critical for energy-harvesting and photovoltaic applications. Structural stability was confirmed under cubic symmetry and mechanical deformation, ensuring robustness for practical implementation. Among the investigated compounds, bromide-based perovskites, particularly Ga2CeBr6 and Tl2CeBr6, demonstrated superior optoelectronic performance, characterized by optimal band gaps, high absorption coefficients, and strong dielectric responses, while In2CeBr6 exhibited notable mechanical robustness and ductility. To the best of current knowledge, no prior theoretical or experimental study has systematically explored the D2CeX6 family in this context. The findings establish Ce-based halide double perovskites as promising lead-free alternatives to toxic Pb-based materials, offering a well-balanced combination of efficiency, stability, and sustainability for next-generation photovoltaic and advanced power-generation technologies.
![]() | (1) |
![]() | (2) |
![]() | (3) |
Drift-diffusion current equations,
![]() | (4) |
![]() | (5) |
![]() | (6) |
Short-circuit current density (JSC),
| JSC = q × A × G × η | (7) |
Fill factor (FF),
![]() | (8) |
Power conversion efficiency (PCE),
![]() | (9) |
![]() | ||
| Fig. 1 X-ray diffraction (XRD) patterns of (a) Ga2CeCl6 and Ga2CeBr6, (b) In2CeCl6 and In2CeBr6, (c) Tl2CeCl6 and Tl2CeBr6 vacancy-ordered perovskites. | ||
m (#225).46 Within this framework, the Ce4+ cations occupy the octahedral coordination sites, where each Ce atom is surrounded by six halide ions (Cl− or Br−), forming corner-sharing CeX6 octahedra that serve as the fundamental building blocks of the lattice. The D-site cations (Ga, In, Tl) are accommodated in the interstitial cavities, stabilizing the perovskite structure through electrostatic interactions. In terms of crystallographic positions, the D-site atoms reside at the face-centered Wyckoff position 3c (0, 0.5, 0.5), the Ce atom is located at the corner site with Wyckoff position 1a (0, 0, 0), and the halide atoms occupy the body-centered Wyckoff position 1b (0.5, 0.5, 0.5).
This ordered arrangement reflects the vacancy-ordered nature of the perovskite, where one of the B-site cations in the conventional ABX3 lattice is replaced by a structural vacancy, giving rise to the distinctive A2BX6 stoichiometry. As illustrated in Fig. 2(a), the atomic configurations of analogous compounds such as Ga2CeCl6, Ga2CeBr6, In2CeCl6, In2CeBr6, Tl2CeCl6, and Tl2CeBr6 display a similar cubic symmetry and site occupancy pattern, highlighting the robustness of this structural motif across different compositions. The high degree of symmetry and well-defined cation coordination in these materials play a crucial role in determining their electronic structure, phonon dynamics, and defect tolerance properties that are directly linked to their optoelectronic and photovoltaic performance.
![]() | ||
| Fig. 2 The (a) crystal structure, (b) proposed solar cell device structure of D2CeX6 vacancy-ordered perovskites. | ||
The optimized lattice constants of these vacancy-ordered perovskites are 7.071 Å to 7.784 Å, as shown in Table 1, primarily governed by the ionic radii of both the D-site cations and halide anions. Predictably, substitution of Cl− with the larger Br− ion results in a noticeable increase in lattice parameters. For instance, Ga2CeCl6 possesses a lattice constant of 7.410 Å, which expands to 7.784 Å for Ga2CeBr6. A similar lattice expansion trend is observed for the In- and Tl-based systems. The corresponding unit cell volumes reflect this trend, increasing with the incorporation of the larger halide anion. The highest unit cell volume of 333.524 Å3 is recorded for Ga2CeBr6, while In2CeBr6 and Tl2CeBr6 both display volumes of 249.999 Å3, indicating a consistent cubic framework for these bromide compounds. Furthermore, the calculated volumes obtained using GGA–PBE and HSE06 functionals show negligible differences, indicating reliable structural optimization. Notably, the calculated densities are 2.841 g cm−3 for Ga2CeCl6, 3.277 g cm−3 for In2CeCl6, and 4.233 g cm−3 for Tl2CeCl6, while Ga2CeBr6, In2CeBr6, and Tl2CeBr6 exhibit 3.778, 5.640, and 6.830 g cm−3, respectively. These variations clearly demonstrate that both the nature of the D2-site cation (Ga, In, Tl) and the X6 anion (Cl, Br) strongly influence the structural packing and overall density, with heavier Tl and Br substitutions yielding denser frameworks. In addition to structural analysis, the electronic band gaps were calculated using both GGA–PBE and HSE06 functionals (Table 1). As anticipated, the HSE06 hybrid functional produced consistently higher band gaps owing to its more accurate treatment of exchange–correlation effects. For instance, Ga2CeCl6 exhibited an increase from 1.733 eV (GGA–PBE) to 1.776 eV (HSE06), with similar enhancements observed across all compounds. The bromide analogues displayed the lowest gaps, with In2CeBr6 showing the smallest values of 0.859 eV (GGA–PBE) and 2.170 eV (HSE06), consistent with the established trend of band gap narrowing as the halide ionic radius increases.
| Ref. | Compounds | GGA–PBE | HSE-06 | Formation energy, ΔEf (eV per atom) | ||||
|---|---|---|---|---|---|---|---|---|
| Energy band gap, eV | Lattice constant a0 (Å) | Unit cell volume, V (Å3) | Density, g cm−3 | Energy band gap, eV | Unit cell volume, V (Å3) | |||
| This work | Ga2CeCl6 | 1.733 | 7.410 | 287.748 | 2.841 | 1.776 | 287.775 | −3.597 |
| Ga2CeBr6 | 1.276 | 7.784 | 333.524 | 3.778 | 2.843 | 333.528 | −3.227 | |
| In2CeCl6 | 1.555 | 7.473 | 295.156 | 3.277 | 2.261 | 249.999 | −3.571 | |
| In2CeBr6 | 0.859 | 7.071 | 249.999 | 5.640 | 2.170 | 249.999 | −3.064 | |
| Tl2CeCl6 | 1.755 | 7.503 | 298.718 | 4.233 | 3.632 | 298.732 | −3.699 | |
| Tl2CeBr6 | 1.364 | 7.071 | 249.999 | 6.830 | 1.418 | 249.999 | −3.178 | |
| 48 | K2CeCl6 | 1.742 | 7.522 | 300.945 | 2.378 | — | — | −3.714 |
| Rb2CeCl6 | 1.748 | 7.619 | 312.774 | 2.780 | — | — | −3.720 | |
| Cs2CeCl6 | 1.807 | 7.863 | 343.881 | 2.987 | — | — | −3.746 | |
| Fr2CeCl6 | 1.830 | 8.111 | 377.413 | 3.514 | — | — | −3.705 | |
To further evaluate the thermodynamic stability of these perovskites, their formation energy (ΔEf) was computed,47 as in Table 1.
| ΔEf = Etot (A2CeX6) − 2E(A) − E(Ce) − 6E(X) | (10) |
Negative formation energy indicates that a compound is thermodynamically stable and can form spontaneously, with more negative values reflecting stronger bonding and greater intrinsic stability. It's a key factor in assessing whether a material is synthesizable and suitable for applications like photovoltaics or catalysis. All compounds exhibit negative formation energy, confirming their thermodynamic favorability. Tl2CeCl6 displayed the most negative formation energy value of −3.699 eV per atom, indicating it as the most stable composition among the series. In contrast, Ga2CeBr6 showed the least negative value (−3.227 eV per atom), likely due to the combined effect of the lighter Ga cation and the larger, less tightly bound Br− anion.
The calculated structural and electronic properties of Ga2CeX6, In2CeX6, and Tl2CeX6 (X = Cl, Br) exhibit close similarities with the reported results of K2CeCl6, Rb2CeCl6, Cs2CeCl6, and Fr2CeCl6. In both cases, the GGA–PBE predicted band gaps are within the same semiconducting range. At the same time, the lattice constants and unit cell volumes follow a systematic increase with the incorporation of larger A-site cations. Furthermore, the negative formation enthalpies obtained in this work are comparable to those in ref. 48, confirming the thermodynamic stability of these Ce-based halide double perovskites.
Table 2 presents the bond length variations in D2CeX6 (D = Ga, In, Tl; X = Cl, Br), considering D–X, Ce–X, and Ce–D interactions. The D–X bonds range from 3.705 to 3.892 Å, with chlorides showing slightly shorter values (3.705 to 3.751 Å) than bromides, except for In2CeBr6 and Tl2CeBr6, where the D–X length decreases to 3.540 Å. The Ce–X bonds are consistently shorter (2.626 to 2.785 Å), indicating stronger Ce–X interactions compared to D–X; as expected, Ce–Br bonds (≈2.682 to 2.785 Å) are longer than Ce–Cl bonds (≈2.626 to 2.631 Å) owing to the larger ionic radius of Br−. The Ce–D bonds are the longest, ranging between 4.330 and 4.766 Å, with the maximum in Ga2CeBr6 and the minimum in In2CeBr6 and Tl2CeBr6. These variations reveal that the bond lengths are sensitive to both the halide ion and the D-site cation, where bromides generally favor longer separations than chlorides, and the shorter Ce–X bonds suggest stronger Ce-centered bonding, which may play a decisive role in stabilizing the crystal structure and influencing the electronic interactions within these vacancy-ordered double perovskites.
| Compounds | Bond length, L (Å) | ||
|---|---|---|---|
| LD–X | LCe–X | LCe–D | |
| Ga2CeCl6 | 3.705 | 2.626 | 4.538 |
| Ga2CeBr6 | 3.892 | 2.785 | 4.766 |
| In2CeCl6 | 3.736 | 2.629 | 4.576 |
| In2CeBr6 | 3.540 | 2.682 | 4.330 |
| Tl2CeCl6 | 3.751 | 2.631 | 4.594 |
| Tl2CeBr6 | 3.540 | 2.682 | 4.330 |
![]() | (11) |
![]() | (12) |
The tolerance factor generally falls within the range of 0.8 to 1.0, where values close to 1.0 represent an ideal cubic configuration with minimal lattice distortion. Meanwhile, the octahedral factor typically lies between 0.41 and 0.90, indicating that the B-site cation is appropriately sized to fit within the BX6 octahedron.49,50 When both factors fall within these ranges, the perovskite framework achieves a balanced geometric arrangement, leading to enhanced structural stability. Deviations outside these limits may result in lattice distortions, octahedral tilting, or complete structural instability, highlighting the critical role of ionic size matching in the design of stable vacancy-ordered perovskites.
The calculated tolerance factors, as summarized in Table 2, are 0.7867, 0.7867, 0.8267, 0.8205, 0.8648, and 0.8567 for Ga2CeCl6, Ga2CeBr6, In2CeCl6, In2CeBr6, Tl2CeCl6, and Tl2CeBr6, respectively. All compounds exhibit values within the structural stability range, except Ga2CeCl6 and Ga2CeBr6, which fall slightly below the lower limit but remain close to the stability threshold. This indicates that most of these halide double perovskites possess geometrically favorable ionic packing. At the same time, the Ga-based variants, though marginally less ideal, are still likely to form stable frameworks with only minor octahedral tilting or distortion. The calculated octahedral factors, presented in Table 3, are 0.5359, 0.4949, 0.5359, 0.4949, 0.5359, and 0.4949 for Ga2CeCl6, Ga2CeBr6, In2CeCl6, In2CeBr6, Tl2CeCl6, and Tl2CeBr6, respectively. All values fall within the established structural stability window (0.41 to 0.90), confirming that the B-site cations are appropriately accommodated within the BX6 octahedral framework. This suggests that these compounds are geometrically favorable for forming stable vacancy-ordered perovskite structures.
| Materials | rA | rCe | rX | Tolerance factor (t) | Octahedral factor (µ) |
|---|---|---|---|---|---|
| Ga2CeCl6 | 1.3 | 0.97 | 1.81 | 0.791 | 0.5359 |
| Ga2CeBr6 | 1.3 | 0.97 | 1.96 | 0.7867 | 0.4949 |
| In2CeCl6 | 1.44 | 0.97 | 1.81 | 0.8267 | 0.5359 |
| In2CeBr6 | 1.44 | 0.97 | 1.96 | 0.8205 | 0.4949 |
| Tl2CeCl6 | 1.59 | 0.97 | 1.81 | 0.8648 | 0.5359 |
| Tl2CeBr6 | 1.59 | 0.97 | 1.96 | 0.8567 | 0.4949 |
![]() | ||
| Fig. 3 Energy band gap of (a and b) Ga2CeX6, (c and d) In2CeX6 and (e and f) Tl2CeX6 double perovskite materials using GGA–PBE and HSE06 function. | ||
The band gap forms between the valence-band maximum (VBM) and conduction-band minimum (CBM). As shown in Fig. 3(a–f), DFT reveals a clear dependence on both the D-site cation (D = Ga, In, Tl) and halide (X = Cl, Br). In Fig. 3(a, c and e), within GGA–PBE, the calculated gaps are: Ga2CeCl6 of 1.733 eV, Ga2CeBr6 of 1.276 eV, In2CeCl6 of 1.555 eV, In2CeBr6 of 0.859 eV, Tl2CeCl6 of 1.755 eV, and Tl2CeBr6 of 1.364 eV, placing all but In2CeBr6 (too narrow) within or near the 1.1 to 1.8 eV single-junction “sweet spot.” A clear anion dependence is observed: replacing Cl with Br consistently reduces the band gap due to the higher energy of Br-4p orbitals, which raises the VBM and narrows the gap. As expected, HSE06 increases the gaps because of its improved treatment of exchange: Ga2CeCl6 of 1.776 eV, Ga2CeBr6 of 2.843 eV, In2CeCl6 of 2.261 eV, In2CeBr6 of 2.170 eV, Tl2CeCl6 of 3.632 eV, and Tl2CeBr6 of 1.418 eV, as shown in Fig. 3(b, d and f). Under HSE06, Tl2CeBr6 (1.418 eV) and Ga2CeCl6 (1.776 eV) remain suitable for single-junction devices, whereas the wider-gap Ga2CeBr6, In2CeCl6, In2CeBr6, and Tl2CeCl6 (>2.0 eV) are better suited as top cells in tandem architectures. In contrast, In2CeBr6 at the PBE level (0.859 eV) exemplifies the “narrow-gap” regime (<1.1 eV), which enhances IR absorption but suffers from higher thermalization losses.
The Partial Density of States (PDOS) is a powerful tool in materials science for identifying the orbital and atomic contributions to the electronic structure of compounds.54 Such analysis provides key insights into the electronic behavior relevant for solar cells and light-emitting devices. For the D2CeX6 vacancy-ordered perovskites, the valence band (VB) extends from −6 to 0 eV, while the conduction band (CB) spans from 0 to +18 eV, as illustrated in Fig. 4. In the valence band region, the dominant contributions originate from halogen p-states hybridized with the D-site cation s-states: Cl-3p and Ga-4s in Ga2CeCl6 (Fig. 4(a)), Cl-3p and In-5s in In2CeCl6 (Fig. 4(b)), Cl-3p and Tl-6s in Tl2CeCl6 (Fig. 4(c)), Br-4p and Ga-4s in Ga2CeBr6 (Fig. 4(d)), Br-4p and In-5s in In2CeBr6 (Fig. 4(e)), and Br-4p and Tl-6s in Tl2CeBr6 (Fig. 4(f)). In all cases, Ce-4d states provide a minor but noticeable contribution to the VB. In the conduction band region, the primary states arise from Ce-5f orbitals, strongly mixed with the p-states of the D-site cations: Ce-5f and Ga-4p in Ga2CeCl6 (Fig. 4(a)), Ce-5f and In-5p in In2CeCl6 (Fig. 4(b)), Ce-5f and Tl-6p in Tl2CeCl6 (Fig. 4(c)), Ce-5f and Ga-4p in Ga2CeBr6 (Fig. 4(d)), Ce-5f and In-5p in In2CeBr6 (Fig. 4(e)), and Ce-5f and Tl-6p in Tl2CeBr6 (Fig. 4(f)).
![]() | ||
| Fig. 4 The partial density of states of (a) Ga2CeCl6, (b) Ga2CeBr6, (c) In2CeCl6, (d) In2CeBr6, (e) Tl2CeCl6, and (f) Tl2CeBr6 vacancy-ordered double perovskites. | ||
Notably, localized peaks appear just above the Fermi level (EX), confirming that the conduction-band minimum (CBM) is dominated by Ce-derived states. Overall, the PDOS indicates that the VBM is mainly governed by halogen p-states hybridized with D-site s-states, while the CBM is strongly influenced by Ce-5f states, a characteristic feature of these vacancy-ordered double perovskites. It is worth noting that a slight mismatch between the band gap values obtained from the band structure (Fig. 3) and the PDOS (Fig. 4) is observed. This minor deviation originates from the intrinsic limitations of DFT within the PBE framework, differences in k-point sampling, and the broadening applied in DOS plotting.55 Nevertheless, all data have been carefully rechecked, and the overall electronic trends remain fully consistent, confirming the reliability of the presented results.
Table 4 presents the HOMO–LUMO energy values of six halide double perovskite compounds, namely Ga2CeCl6, Ga2CeBr6, In2CeCl6, In2CeBr6, Tl2CeCl6, and Tl2CeBr6, which help evaluate their electronic properties and chemical reactivity. The eigenvalues of both HOMO and LUMO orbitals are negative, indicating bound electronic states, while the difference between them represents the HOMO–LUMO energy gap (ΔE), which reflects the ease of electronic excitation. Among these, bromine-based compounds (Br6) exhibit slightly smaller energy gaps than their chlorine-based counterparts (Cl6), implying that Br-based systems are more reactive and less stable, with better potential for charge transfer and optical activity.
| Compounds | Type | N | s | Eigenvalue | Field |
|---|---|---|---|---|---|
| Ga2CeCl6 | HOMO | 111 | + | −0.263252 | Yes |
| LUMO | 112 | + | −0.219414 | Yes | |
| Ga2CeBr6 | HOMO | 165 | + | −0.244130 | Yes |
| LUMO | 166 | + | −0.212826 | Yes | |
| In2CeCl6 | HOMO | 129 | + | −0.266703 | Yes |
| LUMO | 130 | + | −0.222910 | Yes | |
| In2CeBr6 | HOMO | 183 | + | −0.227664 | Yes |
| LUMO | 184 | + | −0.185877 | Yes | |
| Tl2CeCl6 | HOMO | 161 | + | −0.267885 | Yes |
| LUMO | 162 | + | −0.224097 | Yes | |
| Tl2CeBr6 | HOMO | 215 | + | −0.229307 | Yes |
| LUMO | 216 | + | −0.186823 | Yes |
Conversely, Cl-based compounds show larger gaps, indicating greater chemical stability and lower electronic polarizability. Thus, the trend reveals that the HOMO–LUMO gap decreases in the order Cl6 > Br6, suggesting that substitution of Cl by Br enhances the electronic softness and conductivity of the materials, which is beneficial for optoelectronic and semiconductor applications.
Fig. 5 presents the spatial distributions of the HOMO and LUMO for Ga2CeCl6 and Ga2CeBr6 double perovskites, illustrating their electronic charge density and orbital hybridization characteristics. In both compounds, the HOMO states are primarily localized around the halogen atoms (Cl or Br) and partially extend toward the Ce atoms, indicating that the valence band maximum (VBM) mainly originates from the p-orbitals of the halogens with minor Ce contributions.
Conversely, the LUMO orbitals are more delocalized over the Ce and Ga atoms, suggesting that the conduction band minimum (CBM) is predominantly governed by Ce 4f and Ga 4s/4p orbitals. This behavior highlights a charge transfer from the halogen sites to the metal centers during electronic excitation. Comparatively, Ga2CeCl6 exhibits a more compact and symmetric orbital distribution owing to the smaller ionic radius and stronger bonding of Cl atoms, whereas Ga2CeBr6 shows broader and slightly distorted orbital features due to the weaker Ga–Br and Ce–Br interactions associated with the larger Br ions. Overall, both materials demonstrate a similar electronic transition pattern—from halogen-dominated HOMO to metal-centered LUMO—but Ga2CeCl6 displays stronger orbital overlap and higher electronic localization, resulting in a wider band gap and enhanced bonding strength. In contrast, Ga2CeBr6 exhibits narrower band gaps and improved charge transfer characteristics, making it potentially more favorable for optoelectronic applications.
Fig. 6 illustrates the HOMO and LUMO charge density distributions of In2CeCl6 and In2CeBr6 double perovskites, providing a detailed visualization of their electronic interactions, orbital hybridization, and charge transfer mechanisms. In both compounds, the HOMO orbitals are primarily localized around the halogen atoms (Cl or Br), with noticeable overlap extending toward the Ce atoms. This indicates that the VBM mainly originates from the halogen p-orbitals, slightly hybridized with Ce f-orbitals, suggesting a strong halogen–Ce bonding contribution in the occupied states. Conversely, the LUMO orbitals are predominantly distributed over the Ce and In atoms, revealing that the CBM is mainly derived from Ce 4f and In 5s/5p orbitals. This spatial separation between the HOMO and LUMO charge densities confirms a halogen-to-metal charge transfer (HMCT) mechanism during electronic excitation, a typical feature of Ce-based halide perovskites. Comparatively, In2CeCl6 exhibits more compact and symmetric charge density lobes around the Ce–Cl–In framework, indicating stronger covalent bonding and tighter electronic confinement. This results in a larger band gap and enhanced structural rigidity. The smaller ionic radius and higher electronegativity of Cl promote stronger orbital overlap and increased charge localization. In contrast, In2CeBr6 displays more extended and diffuse charge distributions around the Br atoms, reflecting weaker In–Br and Ce–Br interactions due to the larger atomic radius and lower electronegativity of Br. This delocalization facilitates higher electron mobility and reduces the band gap, making the bromide compound more suitable for optoelectronic and photovoltaic applications that demand efficient charge transport. Overall, while both In2CeCl6 and In2CeBr6 exhibit a similar electronic transition pattern (halogen p → Ce/In f/s), In2CeCl6 is characterized by stronger bonding and greater electronic localization—ideal for structural and electronic stability—whereas In2CeBr6 shows enhanced delocalization and charge transport properties, making it a promising candidate for devices requiring improved conductivity and light absorption efficiency.
Fig. 7 presents the spatial charge density distributions of the frontier molecular orbitals HOMO and LUMO for Tl2CeCl6 and Tl2CeBr6, offering crucial insight into their electronic behavior and bonding nature. In Tl2CeCl6, the HOMO (top left) shows electron density predominantly concentrated around the Cl and Ce atoms, suggesting strong Cl-p and Ce-f orbital hybridization within the valence band. This localization near the halide atoms indicates that the valence band maximum (VBM) is largely halogen-derived, with only a minor contribution from Ce orbitals.
Conversely, in the LUMO (top right), the charge density becomes more delocalized toward Ce and Tl centers, signifying that the conduction band minimum (CBM) arises mainly from Ce-d and Tl-p orbitals. This spatial transition of electron density from halogen to metal centers across HOMO–LUMO levels highlight a characteristic charge-transfer excitation process within the Tl–Ce–Cl framework. For Tl2CeBr6 (bottom row), a similar orbital distribution pattern is observed, yet with notable differences in the extent and nature of charge delocalization. The HOMO charge density around Br and Ce atoms appears more diffused and extensive than in Tl2CeCl6, reflecting the larger size and lower electronegativity of Br compared to Cl. This results in weaker bonding and reduced orbital confinement, which can lower the band gap energy. In the LUMO of Tl2CeBr6, electron density remains focused on Ce and Tl sites but extends more toward the Br orbitals, implying enhanced orbital overlap and electronic coupling between the halide and metal centers. Overall, both compounds share similar electronic topologies, where the valence band is dominated by halide-Ce interactions and the conduction band by Ce–Tl hybridization. However, the Br-based compound exhibits greater orbital delocalization, smoother charge transitions, and stronger metal-halide interaction continuity, all of which contribute to narrower band gaps and potentially improved optical absorption and charge transport. These electronic distinctions underline the role of halide substitution in tuning the band-edge characteristics and optoelectronic performance of Tl–Ce halide perovskites.
Tl2CeBr6 shows a slightly lower maximum of 1.61 × 101 a.u. with a minimum of 1.29 × 10−2 a.u., reflecting substantial accumulation around Br but limited interstitial density. In2CeCl6 exhibits a lower maximum of 1.09 × 101 a.u. and a very small minimum of 7.61 × 10−4 a.u., while In2CeBr6 shows a comparable maximum with a slightly higher minimum (1.36 × 10−2 a.u.), suggesting more diffuse electron density around Br. The Ga-based systems display the strongest localization: Ga2CeCl6 reaches the highest maximum among all studied compounds at 3.95 × 101 a.u. with a near-zero minimum of 2.92 × 10−3 a.u., while Ga2CeBr6 retains similarly high values (3.94 × 101 a.u.; minimum 3.76 × 10−3 a.u.). Overall, the Ga-containing systems demonstrate the highest peak charge densities, whereas the In-based compounds show the lowest. A systematic trend emerges with A-site substitution, where electron density becomes more diffuse from Ga → In → Tl, consistent with increasing ionic size and reduced covalent interactions. Furthermore, Cl-based perovskites consistently exhibit stronger electron localization than their Br-based counterparts. These results highlight the direct connection between charge density features, bonding character, and the electronic stability of the D2CeX6 family. Notable studies supporting these findings include work by Faizan et al., Rifat et al., and Saikot et al., which collectively highlight the role of charge density analysis in advancing halide double perovskite research for next-generation optoelectronic applications.7,48,61
In this study, Fig. 9 presents the calculated optical properties of the D2CeX6 (D = Ga, In, Tl; X = Cl, Br) vacancy-ordered double perovskite compounds. The refractive indices of conventional double perovskites such as (Cs2, K2, Rb2)PbCl6 in the visible and ultraviolet regions offer significant advantages for solar cell applications, as their wide dielectric constants enable strong photon absorption within the 1 to 5 eV energy range, making them suitable for efficient solar energy harvesting.63 Similarly, Cs2PbX6 (X = Cl, Br, I) double perovskites demonstrate excellent optoelectronic properties, with direct band gaps ranging from 0.45 to 2.54 eV and high absorption coefficients of approximately 5.95 × 105 cm−1, which highlight their potential for photovoltaic and optical applications.64 However, the inherent toxicity and poor environmental stability of lead severely limit its large-scale utilization. On the other hand, Although Si-based halide double perovskites A2SiI6 (A = K, Rb, Cs; X = Cl, Br, I) show suitable band gaps (0.84 to 1.15 eV), strong visible absorption, and good thermal stability, conventional silicon materials still suffer from an indirect 1.12 eV band gap, weak light absorption, high-temperature processing requirements, and limited flexibility, making them less efficient.65 In contrast, our investigated D2CeX6 compounds exhibit comparable or even superior optical characteristics, such as strong visible-light absorption, high dielectric constants, and large refractive indices, while eliminating the environmental and health hazards associated with lead. Moreover, their tunable band gaps within the optimal range for solar energy conversion, coupled with robust structural and mechanical stability, establish D2CeX6 materials as advanced, sustainable, and highly promising candidates for next-generation lead-free optoelectronic and photovoltaic devices.
![]() | ||
| Fig. 9 Optical functions (a) dielectric function, (b) absorption coefficient, (c) conductivity, (d) refractive index, (e) reflectivity, (f) loss function of D2CeX6 double perovskites. | ||
![]() | (13) |
![]() | (14) |
Among the studied compounds, the static (0 eV) real part of the dielectric function (ε1) is found to be 4.48 for Ga2CeCl6, 5.41 for Ga2CeBr6, 4.56 for In2CeCl6, 7.65 for In2CeBr6, 4.28 for Tl2CeCl6, and 7.216 for Tl2CeBr6, as presented in Fig. 9(a). The ε1 parameter reflects a material's capacity to store electric energy when subjected to an external electromagnetic field. With increasing photon energy, ε1 initially increases due to enhanced polarization arising from interband electronic transitions, reaching a peak value before gradually declining. The relatively high ε1 values suggest strong polarization capability and effective screening of internal electric fields, which are beneficial for lowering exciton binding energy and improving charge carrier separation in optoelectronic applications.
For all the investigated compounds, the imaginary part of the dielectric function (ε2), which represents the material's ability to absorb electromagnetic energy, begins to increase from photon energies between 0.31 and 0.87 eV. As photon energy rises, ε2 increases and reaches its maximum within the visible region (1.5–4 eV), with peak values of 2.83 for Ga2CeCl6, 4.74 for Ga2CeBr6, 2.67 for In2CeCl6, 8.15 for In2CeBr6, 1.59 (at 3.31 eV) for Tl2CeCl6, and 5.067 for Tl2CeBr6. These maximum values generally occur around 4.0 eV. The highest ε2 peak observed in In2CeBr6 indicates significant energy loss due to strong absorption of electromagnetic radiation, whereas Tl2CeCl6 exhibits the lowest ε2 value, suggesting comparatively weaker absorption behavior.
The comparison between chloride and bromide analogs reveals that Br-based compounds generally exhibit enhanced imaginary dielectric components due to narrower band gaps, which facilitate easier electronic excitations. Conversely, Cl-based compounds show higher real parts, indicating stronger dielectric screening. The cation variation (Ga → In → Tl) affects both amplitude and energy dispersion, with heavier cations (Tl) contributing to stronger polarization and broader spectral response. Collectively, these dielectric characteristics reveal that Ga2CeCl6 and Tl2CeBr6 are particularly promising for optoelectronic applications requiring high dielectric screening and effective visible-light absorption.
![]() | (15) |
The absorption coefficient, α(ω), indicates how efficiently a material absorbs light at a particular angular frequency ω. It is commonly measured in cm−1 and is directly related to the extinction coefficient k(ω), which represents the imaginary part of the complex refractive index.71,72 Fig. 9(b) shows that all six D2CeX6 compounds exhibit significant absorption coefficients, starting at photon energies of 0.8 to 1.7 eV. As the photon energy increases toward 3.0 eV, the absorption coefficients rise sharply, reaching values of 18
524 cm−1 (Ga2CeCl6), 24
394 cm−1 (Ga2CeBr6), 16
784 cm−1 (In2CeCl6), 27
560 cm−1 (In2CeBr6), 17
765 cm−1 (Tl2CeCl6), and 25
221 cm−1 (Tl2CeBr6). Such high absorption coefficients (α > 104 cm−1)73,74 in the visible range are highly desirable for thin-film solar cells, as they allow efficient photon harvesting with minimal material thickness, indicating that all studied materials are promising candidates for photovoltaic applications. The absorption behavior is also influenced by the halide substitution (Cl to Br), which shifts the absorption edge toward lower energies due to the narrower band gap of Br-based compounds, thereby enhancing visible-light absorption. Additionally, the choice of D-site cation (Ga, In, Tl) affects the magnitude and position of absorption peaks, with In- and Tl-based compounds generally exhibiting higher absorption coefficients than Ga-based ones. These trends highlight the tunability of optical properties in D2CeX6 double perovskites through compositional engineering, which is crucial for optimizing performance in both solar energy conversion and optoelectronic applications.
![]() | (16) |
Fig. 9(c) shows the real and imaginary parts of optical conductivity (σ) as a function of photon energy (0 to 4 eV) for the double perovskites Ga2CeCl6, Ga2CeBr6, In2CeCl6, In2CeBr6, Tl2CeCl6, and Tl2CeBr6. Among all, Ga2CeCl6 exhibits the highest real conductivity, rising sharply beyond 2.5 eV, indicating superior photon-induced carrier excitation and possible metallic-like behavior at high photon energy. Ga2CeBr6 also shows strong conductivity but slightly lower than Ga2CeCl6, suggesting that replacing Cl with Br slightly reduces the optical response, likely due to the larger ionic radius and weaker bonding of Br. The In-based compounds (In2CeCl6 and In2CeBr6) show moderate conductivity values, implying less efficient charge carrier transition compared to Ga-based systems; however, In2CeBr6 displays a smoother rise, indicating better optical absorption at higher energies than In2CeCl6. In contrast, Tl-based compounds (Tl2CeCl6 and Tl2CeBr6) possess the lowest real conductivity across the entire photon energy range, reflecting their relatively poor electronic transition probability and higher resistive character. As shown in Fig. 6(c), the imaginary part of the optical conductivity initially starts with negative values, indicating a reactive polarization response. With increasing photon energy, the conductivity decreases further and reaches minimum values of −2.0 at 4.00 eV for Ga2CeCl6, −1.95 at 3.40 eV for Ga2CeBr6, −2.184 at 4.00 eV for In2CeCl6, −3.43 at 3.42 eV for In2CeBr6, −1.57 at 4.00 eV for Tl2CeCl6, and −3.41 at 3.56 eV for Tl2CeBr6. This transient negative behavior reflects strong reactive polarization, which is advantageous for applications such as optical switching and memory devices. Notably, cation substitution (Ga → In → Tl) and halide replacement (Cl → Br) systematically influence the peak positions and depth of the negative conductivity response.
Heavier cations and larger halides generally shift the spectral features toward lower energies and enhance polarization effects, thereby tailoring the material's suitability for frequency-selective optoelectronic and photonic applications.
![]() | (17) |
Fig. 9(d) illustrates the variation of reflectivity (ρ) with photon energy (E) from 0 to 4 eV for six Ce-based double perovskite compounds: Ga2CeCl6, Ga2CeBr6, In2CeCl6, In2CeBr6, Tl2CeCl6, and Tl2CeBr6. Reflectivity represents the fraction of incident light that is reflected from a material's surface and serves as an important indicator of its optical response and surface electronic structure. Overall, all compounds exhibit relatively low reflectivity (<0.4), suggesting strong light absorption and significant potential for optoelectronic and photovoltaic applications. At 0.1 eV, the initial reflectivity values are 0.128, 0.1589, 0.131, 0.220, 0.1213, and 0.1213 for Ga2CeCl6, Ga2CeBr6, In2CeCl6, In2CeBr6, Tl2CeCl6, and Tl2CeBr6, respectively. All materials show notable optical activity in the visible range (1.5–4 eV). As the photon energy increases, the reflectivity rises to 0.189, 0.23, 0.185, 0.32, 0.135, and 0.27 for Ga2CeCl6, Ga2CeBr6, In2CeCl6, In2CeBr6, Tl2CeCl6, and Tl2CeBr6, respectively. Overall, these reflectivity characteristics highlight distinct, application-specific roles, underscoring the importance of reflectivity in optimizing these materials for optoelectronic technologies.
![]() | (18) |
Fig. 9(e) illustrates the variation of the refractive index (n), both real and imaginary components, with photon energy (0 to 4 eV) for six Ce-based double perovskite compounds: Ga2CeCl6, Ga2CeBr6, In2CeCl6, In2CeBr6, Tl2CeCl6, and Tl2CeBr6. The solid lines denote the real part of the refractive index, which signifies the phase velocity of light within the material, while the dashed lines indicate the imaginary part, which represents the extinction coefficient or absorption losses. Overall, the real part of n remains between approximately 2.0 and 4.5, showing moderate to high optical density typical of semiconducting materials. Ga2CeCl6 exhibits the highest real refractive index across most of the energy range, peaking near 4.5, implying stronger light–matter interaction and higher optical polarization. Ga2CeBr6 also shows a comparatively high refractive index but slightly lower than its Cl counterpart, suggesting that Br substitution reduces the electronic polarizability due to its larger ionic radius and weaker bond strength. The In-based compounds (In2CeCl6 and In2CeBr6) have intermediate refractive indices around 2.5–3.0, indicating moderate optical response with more uniform energy dependence, suitable for balanced optical transmission and absorption. In contrast, Tl2CeCl6 and Tl2CeBr6 display the lowest refractive indices (∼2.0 to 2.5), reflecting weaker electronic polarization and higher optical transparency. The imaginary components increase gradually beyond 2 eV for all compounds, highlighting the onset of interband electronic transitions that contribute to absorption. Among these, Ga-based compounds again exhibit the strongest imaginary peaks, confirming their intense optical activity, while Tl-based compounds show the weakest, consistent with their low absorption tendency. Overall, both halide and cation substitution significantly influence refractive behavior, with the comparative trend for real refractive index following Ga2CeCl6 > Ga2CeBr6 > In2CeBr6 > In2CeCl6 > Tl2CeCl6 > Tl2CeBr6. This indicates that Ga-based systems, particularly Ga2CeCl6, possess superior optical confinement and are more suitable for light-modulating or refractive applications, whereas Tl-based perovskites are better for transparency-driven optoelectronic devices.
![]() | (19) |
As shown in Fig. 9(f), the relationship between the Loss Function, L(ω), on the y-axis and photon energy, E (in electron volts, eV), on the x-axis for six different materials: Ga2CeCl6, Ga2CeBr6, In2CeCl6, In2CeBr6, Tl2CeCl6, and Tl2CeBr6. Each material is represented by a distinct colored line as indicated in the legend: green for Ga2CeCl6, red for Ga2CeBr6, brown for In2CeCl6, cyan for In2CeBr6, magenta for Tl2CeCl6, and orange for Tl2CeBr6. The general trend for all materials shows that the loss function starts near zero at low photon energies (0 eV) and increases as the photon energy increases up to 4 eV. However, the rate and magnitude of increase differ for each compound. Ga2CeBr6 exhibits the highest loss function values, surpassing 0.10 at around 4 eV, indicating the strongest loss effect at high photon energies among the six. Tl2CeCl6 follows closely, peaking slightly below Ga2CeBr6 but still above 0.08. Ga2CeCl6 and In2CeCl6 show moderate increases, with their loss functions rising to approximately 0.08 and 0.07, respectively, at 4 eV. The In2CeBr6 and Tl2CeBr6 compounds display the lowest loss functions overall, with Tl2CeBr6 being the lowest, peaking near 0.06. Comparatively, the bromide variants (Ga2CeBr6, In2CeBr6, Tl2CeBr6) generally display higher loss functions than their chloride counterparts at higher photon energies, except for In2CeBr6, which is lower than In2CeCl6. Among the cations, Ga-based compounds tend to show higher loss functions than In-based compounds, while Tl-based compounds are intermediate but closer to Ga-based, especially in the chloride form. This suggests that both the halide type and the central metal ion influence the loss function behavior, with bromides and Ga ions generally enhancing the loss function, except in specific cases like In2CeBr6.
| σij = Cijklεkl | (20) |
| C11 > 0, 4C44 > 0, C11 − C12 > 0 and C11 + 2C12 > 0 | (21) |
All the conditions of the Born–Huang stability criteria are satisfied and yield positive values, indicating that the D2CeX6 double perovskites are mechanically stable, as shown in Fig. 10. This confirms their ability to withstand external stress without structural failure, making them suitable for practical applications in devices where mechanical integrity is essential.
The mechanical properties of D2CeX6 double perovskites are critical for ensuring structural integrity, thermal stability, and resistance to external stress. The bulk modulus, shear modulus, Young's modulus, Poisson's ratio, and Pugh's ratio serve as indicators of ductility or brittleness, guiding decisions about the materials' suitability for device integration.48,61 Mechanical stability plays a direct role in the long-term reliability and lifespan of perovskite-based solar cells. Therefore, a thorough understanding and optimization of these properties are essential for developing mechanically robust, efficient, and commercially viable perovskite solar absorbers for future energy applications.
Fig. 11 illustrates the mechanical parameters of D2CeX6 double perovskites, including bulk modulus (B), shear modulus (G), Young's modulus (Y), Poisson's ratio (ν), Pugh's ratio (B/G), and anisotropy factor (AU), which are represented by equations as follows.86
![]() | (22) |
![]() | (23) |
![]() | (24) |
![]() | ||
| Fig. 11 Mechanical parameters of (a) Bulk modulus, (b) Shear modulus, (c) Young's modulus, (d) Poisson's ratio, (e) Pugh's modulus, and (f) anisotropy of D2CeX6 double perovskite materials. | ||
Bulk modulus (B) measures a material's resistance to uniform compression and indicates mechanical stiffness. Higher B values reflect strong interatomic bonding and limited compressibility, while lower values correspond to flexible or soft materials. Typically, materials with B > 40 GPa are classified as hard, whereas B < 40 GPa denotes soft or flexible behavior.48 As shown in Fig. 11(a), In2CeBr6 (50.30 GPa) and Tl2CeBr6 (52.72 GPa) are hard materials, while Ga2CeCl6, Ga2CeBr6, In2CeCl6, and Tl2CeCl6 are soft, capable of accommodating volume changes more easily. Shear modulus (G) quantifies resistance to shape (shear) deformation at constant volume. Higher G values indicate strong internal bonding and shear rigidity, while lower values correspond to easier shear deformation. As depicted in Fig. 11(b), Tl2CeBr6 (13.16 GPa) and In2CeBr6 (11.77 GPa) exhibit the highest shear moduli, making them suitable for load-bearing and wear-resistant applications.
In contrast, Ga2CeCl6 (1.94 GPa) and Ga2CeBr6 (3.57 GPa) are more flexible, advantageous for damping, vibration isolation, or low-shear optical layers. Young's modulus (E) reflects stiffness and resistance to elongation under uniaxial stress. Fig. 11(c) shows that In2CeCl6 and Tl2CeBr6 have the highest E values, consistent with their B and G moduli, making them ideal for rigid supports, optical mounts, and stiff structural components. Conversely, Ga2CeCl6 (5.34 GPa) exhibits enhanced elasticity, suitable for flexible electronics, wearable devices, and strain-tolerant applications. Overall, the variation of B, G, and E across cation–anion substitutions demonstrate the tunable mechanical behavior of D2CeX6 perovskites, enabling design for either rigid, high-strength applications (In- and Tl–Br based) or flexible, damping-oriented applications (Ga-based). The ductile or brittle nature of a material can be assessed using Poisson's ratio (ν) and Pugh's ratio (B/G). In general, materials with ν exceeding 0.26 and B/G greater than 1.75 are classified as ductile.29
![]() | (25) |
Poisson's ratio (ν), illustrated in Fig. 11(d), describes the ratio of lateral to axial strain under applied stress, reflecting both bonding characteristics and ductility. For the D2CeX6 perovskites studied, the calculated ν values are 0.372, 0.297, 0.318, 0.391, 0.302, and 0.384 for Ga2CeCl6, Ga2CeBr6, In2CeCl6, In2CeBr6, Tl2CeCl6, and Tl2CeBr6, respectively. Since all values exceed the critical threshold of 0.26, these compounds exhibit ductile behavior, indicating their ability to undergo plastic deformation and their suitability for flexible applications such as strain-tolerant devices, vibration-damping layers, and ductile electrodes. Likewise, Pugh's ratio (B/G), shown in Fig. 11(e), provides a complementary measure of ductility, where values above 1.75 confirm ductile behavior. All investigated D2CeX6 perovskites surpass this criterion, reinforcing the conclusion from Poisson's ratio that these materials are inherently ductile.
The Universal Anisotropy Index (AU) in Fig. 11(f) quantifies the directional dependence of elastic properties, where zero denotes perfect isotropy.87
![]() | (26) |
In2CeBr6 (AU = 0.0016) and Tl2CeBr6 (AU = 0.005) exhibit near-isotropic behavior, suitable for applications where mechanical stability is critical, for example, in sensors, microelectronic substrates, and integrated circuits. On the other hand, Ga2CeCl6 displays significant anisotropy (AU = 1.946), implying strongly direction-dependent mechanical responses, which could be leveraged in anisotropic optical or piezoelectric devices where orientation-dependent properties are desired.
From the graphical assessments, Fig. 12–14, Ga2CeCl6 displays the highest mechanical anisotropy with Young's modulus ranging between ∼2.5 and 9.0 GPa, indicating its stiffness varies significantly with direction. In contrast, Tl2CeBr6 shows the highest and most isotropic stiffness (∼35 to 37 GPa), followed by In2CeBr6 (∼32 to 33 GPa), highlighting their superior mechanical robustness and uniformity. Similarly, in shear modulus, Tl2CeBr6 peaks (∼13.5 GPa), while Ga2CeCl6 registers the lowest (∼1.17 to 1.24 GPa).
![]() | ||
| Fig. 12 Graphical assessment of Ga2CeCl6 and Ga2CeBr6 double perovskites anisotropy showing 2D and 3D figures. | ||
![]() | ||
| Fig. 13 Graphical assessment of In2CeCl6 and In2CeBr6 double perovskites anisotropy showing 2D and 3D figures. | ||
![]() | ||
| Fig. 14 Graphical assessment of Tl2CeCl6 and Tl2CeBr6 double perovskites anisotropy showing 2D and 3D figures. | ||
For Poisson's ratio, Tl2CeBr6 and In2CeBr6 exhibit broader, more spherical distributions (∼0.2 to 1.0), suggesting low anisotropy, whereas Ga2CeCl6 again demonstrates high directionality with values dipping to ∼0.05.
Comparing the compounds, bromide-based materials are generally stiffer and more isotropic than their chloride counterparts, and mechanical performance improves from Ga → In → Tl due to increasing atomic size and enhanced bonding.88,89 These insights are vital for tailoring materials for specific directional mechanical responses or ensuring mechanical stability in multifunctional devices. Accurate anisotropy assessment helps optimize materials for real-world applications, guiding experimental synthesis and structural design.
![]() | ||
| Fig. 15 Time evolution of potential, kinetic, and total energy for Ga2CeCl6 double perovskite at (a) 300 K, (b) 600 K, (c) 900 K, and (d) temperature variation over time. | ||
The potential energy remains comparatively lower and negative in all cases, consistent with the system's bound state. Fig. 15(d) presents the temperature variation over time, showing steady temperature maintenance around the target values (300 K, 600 K, and 900 K), with moderate oscillations attributed to thermal equilibrium fluctuations.
Fig. 16 depicts the variation of kinetic energy (K.E.), potential energy (P.E.), total energy (T.E.), and temperature (T) of Ga2CeBr6 as a function of time at different temperatures (300 K, 600 K, and 900 K) during molecular dynamics (MD) simulations. Fig. 16(a)–(c) shows that as temperature increases, the amplitude and average values of K.E. and T.E. rise, reflecting enhanced atomic vibrations and thermal activity. At 300 K, the system exhibits relatively small oscillations and stable energy values, indicating a well-equilibrated and structurally stable phase. At 600 K, both K.E. and T.E increase moderately, with more frequent fluctuations due to higher thermal motion.
![]() | ||
| Fig. 16 Time evolution of potential, kinetic, and total energy for Ga2CeBr6 double perovskite at (a) 300 K, (b) 600 K, (c) 900 K, and (d) temperature variation over time. | ||
At 900 K, the oscillations become much larger, and the total energy reaches its maximum level, signifying strong lattice vibrations and possibly anharmonic effects. P.E remains negative across all temperatures but shows a gradual upward shift with increasing temperature, suggesting partial weakening of interatomic interactions. Fig. 16(d) compares the temperature variation over time, showing that each system stabilizes near its target temperature (300 K, 600 K, and 900 K) with expected thermal fluctuations.
Fig. 17 illustrates the variation of kinetic energy (K.E.), potential energy (P.E.), total energy (T.E.), and temperature (T) of In2CeCl6 during molecular dynamics (MD) simulations at 300 K, 600 K, and 900 K over a 50 ps timescale. Fig. 17(a)–(c), the energy profiles show that as temperature increases, both K.E. and T.E rise significantly, while P.E remains negative but slightly increases, indicating stronger atomic vibrations and enhanced atomic displacement at elevated temperatures.
![]() | ||
| Fig. 17 Time evolution of potential, kinetic, and total energy for In2CeCl6 double perovskite at (a) 300 K, (b) 600 K, (c) 900 K, and (d) temperature variation over time. | ||
At 300 K, the energy fluctuations are relatively small, implying structural stability with limited atomic motion. At 600 K, oscillations in K.E. and T.E become more pronounced, showing moderate thermal agitation. At 900 K, energy oscillations are much larger, signifying intense atomic vibration and possible lattice distortion. Fig. 17(d) compares the temperature profiles, showing stable fluctuations around the target temperatures (300 K, 600 K, and 900 K), confirming proper thermal equilibration during the MD simulations.
Fig. 18 presents the variation of kinetic energy (K.E.), potential energy (P.E.), total energy (T.E), and temperature (T) of In2CeBr6 over 50 ps at three different temperatures, 300 K, 600 K, and 900 K, obtained from molecular dynamics simulations. Fig. 18(a)–(c), the system exhibits increasing energy magnitudes with rising temperature, indicating stronger atomic vibrations and enhanced lattice dynamics. At 300 K, K.E. and T.E. remain relatively stable with small oscillations, reflecting a thermally equilibrated and structurally stable configuration. As the temperature increases to 600 K, both K.E. and T.E rise moderately, accompanied by more pronounced oscillations due to intensified thermal motion. At 900 K, large fluctuations in K.E. and T.E are observed, implying vigorous atomic vibrations and possible anharmonic effects, while P.E remains negative but shifts slightly upward, suggesting reduced interatomic binding strength. Panel (d) compares temperature variations over time, showing consistent stabilization near the target values of 300 K, 600 K, and 900 K with periodic fluctuations characteristic of dynamic equilibrium.
![]() | ||
| Fig. 18 Time evolution of potential, kinetic, and total energy for In2CeBr6 double perovskite at (a) 300 K, (b) 600 K, (c) 900 K, and (d) temperature variation over time. | ||
Fig. 19 presents the variation of energy and temperature with simulation time for Tl2CeCl6 at three different temperatures: 300 K, 600 K, and 900 K. Fig. 19(a to c) show the evolution of kinetic energy (K.E), potential energy (P.E), and total energy (T.E), while Fig. 19(d) depicts the temperature fluctuations over time. At 300 K, Fig. 19(a), the system displays stable energy oscillations with minor fluctuations, indicating good thermal equilibrium and structural stability. As the temperature rises to 600 K, Fig. 19(b), both kinetic and total energies increase noticeably, accompanied by slightly larger oscillations, suggesting enhanced atomic motion. At 900 K, Fig. 19(c), the total and kinetic energies reach their highest average values, and the oscillations become more pronounced, reflecting greater thermal agitation and the possible onset of lattice instability. The potential energy becomes less negative with increasing temperature, implying partial weakening of interatomic interactions. The temperature profile in Fig. 19(d) confirms consistent thermal behavior, where higher set temperatures produce broader fluctuations but maintain equilibrium around their respective mean values.
![]() | ||
| Fig. 19 Time evolution of potential, kinetic, and total energy for Tl2CeCl6 double perovskite at (a) 300 K, (b) 600 K, (c) 900 K, and (d) temperature variation over time. | ||
Fig. 20 illustrates the variation of total energy (T.E), kinetic energy (K.E), potential energy (P.E), and temperature (T) with time for Tl2CeBr6 at different temperatures, 300 K, 600 K, and 900 K, over a simulation period of 50 ps. Fig. 20(a)–(c) shows the energy profiles, while panel Fig. 17(d) presents the temperature fluctuations. At 300 K, Fig. 20(a), the system exhibits relatively stable energy oscillations with moderate kinetic and total energy values and a potential energy of around −40 kcal mol−1, indicating structural stability. As the temperature increases to 600 K, Fig. 20(b), both K.E. and T.E rise, showing higher fluctuation amplitudes, whereas P.E slightly shifts toward less negative values, reflecting increased atomic vibrations.
![]() | ||
| Fig. 20 Time evolution of potential, kinetic, and total energy for In2CeBr6 double perovskite at (a) 300 K, (b) 600 K, (c) 900 K, and (d) temperature variation over time. | ||
At 900 K, Fig. 17(c), kinetic and total energies reach their highest values, and fluctuations become more pronounced, suggesting enhanced atomic motion and thermal agitation. The temperature vs. time plot Fig. 17(d) confirms that Tl2CeBr6 maintains near-equilibrium average temperatures around 300 K, 600 K, and 900 K, respectively, though oscillations widen with rising temperature.
The phonon spectra indicate whether a crystal structure is stable: the absence of imaginary (negative) frequencies confirms dynamical stability, while the presence of such frequencies may signal lattice instabilities or soft phonon modes.94 Fig. 21 presents the phonon dispersion curves for six Ce-based halide double perovskites plotted along the high-symmetry K-points (X–R–M–Γ–R) with frequency (THz) as a function of wave vector: (a) Ga2CeCl6, (b) Ga2CeBr6, (c) In2CeCl6, (d) In2CeBr6, (e) Tl2CeCl6, and (f) Tl2CeBr6 double perovskite materials. From the phonon spectra, Ga2CeCl6, In2CeCl6, and Tl2CeCl6 exhibit no negative frequencies across the entire Brillouin zone, confirming their excellent dynamical stability and structurally robust behavior under equilibrium conditions. In contrast, Ga2CeBr6, In2CeBr6, and Tl2CeBr6 display very small negative frequencies near the Γ-point, suggesting the presence of slight lattice instabilities or soft phonon modes. These modes may become stabilized under finite temperature or pressure, or through minor structural relaxation. Comparing the compounds, chloride-based perovskites possess a more rigid lattice and stronger bonding interactions due to the lighter Cl atoms, resulting in higher phonon frequencies and enhanced dynamic stability. Bromide-based compounds, on the other hand, display softer vibrational modes due to the heavier Br atoms, leading to slightly reduced structural robustness.
![]() | ||
| Fig. 21 Phonon dispersion curves of Ce-based (a) Ga2CeCl6, (b) Ga2CeBr6, (c) In2CeCl6, (d) In2CeBr6, (e) Tl2CeCl6, and (f) Tl2CeBr6 double perovskites materials. | ||
| Compound | Change spilling | Species | Mulliken atomic populations | Mulliken change | Hirshfeld change | ||||
|---|---|---|---|---|---|---|---|---|---|
| s | p | d | f | Total | |||||
| Ga2CeCl6 | 0.14% | Ga | 1.93 | 0.32 | 10.00 | 0.00 | 12.25 | 0.75 | 0.48 |
| Ce | 2.30 | 6.34 | 1.63 | 1.10 | 11.38 | 0.62 | 0.20 | ||
| Cl | 1.95 | 5.40 | 0.00 | 0.00 | 7.35 | −0.35 | −0.19 | ||
| Ga2CeBr6 | 0.10% | Ga | 2.11 | 0.39 | 10.00 | 0.00 | 12.50 | 0.50 | 0.45 |
| Ce | 2.45 | 6.63 | 1.73 | 1.13 | 11.94 | 0.06 | 0.12 | ||
| Br | 1.85 | 5.33 | 0.00 | 0.00 | 7.18 | −0.18 | −0.17 | ||
| In2CeCl6 | 0.14% | In | 1.93 | 0.32 | 10.00 | 0.00 | 12.25 | 0.75 | 0.45 |
| Ce | 2.30 | 6.34 | 1.63 | 1.09 | 11.37 | 0.63 | 0.21 | ||
| Cl | 1.95 | 5.41 | 0.00 | 0.00 | 7.36 | −0.36 | −0.19 | ||
| In2CeBr6 | 0.16% | In | 2.23 | 0.59 | 10.00 | 0.00 | 12.82 | 0.18 | 0.29 |
| Ce | 2.43 | 6.47 | 1.89 | 1.15 | 11.94 | 0.06 | 0.07 | ||
| Br | 1.81 | 5.26 | 0.00 | 0.00 | 7.07 | −0.07 | −0.11 | ||
| Tl2CeCl6 | 0.11% | Tl | 3.93 | 6.31 | 10.00 | 0.00 | 20.24 | 0.76 | 0.46 |
| Ce | 2.30 | 6.34 | 1.63 | 1.09 | 11.36 | 0.64 | 0.21 | ||
| Cl | 1.95 | 5.41 | 0.00 | 0.00 | 7.36 | −0.36 | −0.19 | ||
| Tl2CeBr6 | 0.13% | Tl | 4.16 | 6.59 | 10.00 | 0.00 | 20.75 | 0.25 | 0.30 |
| Ce | 2.44 | 6.47 | 1.89 | 1.15 | 11.95 | 0.05 | 0.07 | ||
| Br | 1.82 | 5.27 | 0.00 | 0.00 | 7.09 | −0.09 | −0.11 | ||
In the present study, population analysis was carried out for a series of D2CeX6 compounds. The Mulliken atomic populations, categorized into s, p, d, and f orbitals, reveal substantial trends in Table 5. Notably, Tl atoms exhibit the highest total Mulliken population at 20.24 (in Tl2CeCl6), owing to their large atomic size and heavy electron contribution from the 6 s and 6p orbitals, whereas halide anions (Cl and Br) consistently show the lowest, with Br in Tl2CeBr6 at just 7.09. Among the cations, Ga atoms in Ga2CeCl6 possess the lowest total Mulliken population of 12.25, reflecting the lighter, less electron-rich nature of Ga compared to In and Tl. Cerium (Ce) atoms display significant f-orbital occupancy (1.10 to 1.15), underlining the role of 4f electrons in determining the material's optical and electronic properties. Further, the Mulliken charge change values indicate the extent of electron gain or loss relative to neutral atoms.
The highest positive charge transfer occurs for Tl in Tl2CeCl6 (0.76), while the largest electron acceptance is seen in Cl in the same compound (−0.36), reaffirming the ionic nature of halides as electron acceptors. Hirshfeld charge changes, though lower in magnitude, mirror these trends, with Tl showing a maximum of 0.46 and Cl a minimum of −0.19 in Tl2CeCl6. Comparing across halide substitutions, replacing Cl with Br generally results in reduced total atomic populations and charge transfers, consistent with Br's lower electronegativity. Similarly, moving from Ga to heavier In and Tl systematically increases total Mulliken populations and charge changes, reflecting the influence of cationic size and electronic configuration on charge distribution. The significance of such population analysis extends to practical applications, as the charge distribution in these double perovskites directly impacts their optical absorption behavior, defect tolerance, and charge carrier dynamics. These factors are critical in determining the performance of perovskite-based materials in applications such as photovoltaics, X-ray scintillators, and optoelectronic devices. Accurate knowledge of charge transfer tendencies aids in tuning material properties, enabling the rational design of compounds with optimized band gaps, enhanced stability, and superior optoelectronic performance. This comprehensive population analysis thus not only elucidates the fundamental electronic structure of D2CeX6 double perovskites but also guides the engineering of advanced materials for future energy and photonic technologies.
Table 6 lists the material and structural parameters employed to configure the SCAPS-1D simulation architecture for the studied D2CeX6 (D = Ga, In, Tl; X = Cl, Br) double perovskite solar cells. The layer thickness is set at 50 nm for the transparent conducting oxide (FTO) and electron transport layer (WS2), while the absorber layer is 800 nm, providing adequate optical absorption while minimizing recombination losses. The band gap (Eg) values, ranging from 0.859 eV (In2CeBr6) to 3.6 eV (FTO), define the spectral absorption range, with the perovskites covering a broad portion of the visible-near-infrared region for effective photon harvesting. Relative dielectric permittivity (εr) values between 3.36 and 13.60 influence the electrostatic screening of charge carriers, directly impacting exciton dissociation and defect tolerance. The electron affinity (χ), varying from 3.950 eV (WS2) to 7.660 eV (In2CeBr6), determines band alignment at interfaces, which is crucial for efficient carrier transport and reduced energy barriers.
| Parameters | FTO | WS2 | Ga2CeCl6 | Ga2CeBr6 | In2CeCl6 | In2CeBr6 | Tl2CeCl6 | Tl2CeBr6 |
|---|---|---|---|---|---|---|---|---|
| Thickness (nm) | 50 | 50 | 800 | 800 | 800 | 800 | 800 | 800 |
| Band gap, Eg (eV) | 3.6 | 2.2 | 1.733 | 1.273 | 1.550 | 0.859 | 1.755 | 1.364 |
| Dielectric permittivity, ∈r | 10 | 13.60 | 3.520 | 3.750 | 3.650 | 3.360 | 3.540 | 3.36 |
| Electron affinity, χ (eV) | 4.5 | 3.950 | 4.472 | 5.410 | 4.560 | 7.660 | 4.280 | 7.160 |
| VB effective density of states, NV (cm−3) | 1.8 ×1019 | 2.2 × 1018 | 1.450× 1019 | 1.492 × 1019 | 1.489 × 1019 | 1.440 × 1019 | 1.478 × 1019 | 1.429× 1019 |
| CB effective density of states, NC (cm−3) | 2 ×1018 | 1.8 × 1019 | 2.460 × 1019 | 2.476 × 1019 | 2.980 × 1019 | 2.930 × 1019 | 2.504 × 1019 | 2.483 × 1019 |
| Donor density, ND (cm−3) | 1 × 1018 | 1 × 1018 | 0 | 0 | 0 | 0 | 0 | 0 |
| Acceptor density, NA (cm−3) | 0 | 0 | 1 ×1017 | 1 ×1017 | 1 ×1017 | 1 ×1017 | 1 ×1017 | 1 ×1017 |
| Hole mobility, µh (cm2 V−1 s −1) | 20 | 100 | 65 | 60 | 67 | 65 | 69 | 69 |
| Electron mobility, µn (cm2 V−1 s −1) | 100 | 100 | 40 | 35 | 37 | 42 | 46 | 46 |
| Total defect density, Nt (cm−3) | 1 ×1014 | 1 × 1014 | 1 × 1014 | 1 × 1014 | 1 × 1014 | 1 × 1014 | 1 × 1014 | 1 × 1014 |
The effective density of states (NV, NC) for the valence and conduction bands, in the range of 1.429 × 1019–1.492 × 1019 cm−3 (NV) and 2.460 × 1019–2.980 × 1019 cm−3 (NC) for perovskites, influences carrier concentration and recombination dynamics. Doping concentrations are assigned such that the FTO and WS2 layers have a donor density (ND) of 1 × 1018 cm−3 to ensure high conductivity, while the absorber layers are lightly p-type doped (NA = 1 × 1017 cm−3) to facilitate charge separation. Hole mobility (µh) values for the absorbers range from 60 to 69 cm2 V−1 s−1, and electron mobility (µn) from 35 to 46 cm2 V−1 s−1, supporting balanced charge transport and minimizing space-charge buildup. The defect density (Nt), fixed at 1 × 1014 cm−3 for all layers, represents a moderate trap concentration to reflect realistic fabrication conditions while avoiding excessive recombination losses. Collectively, these parameters ensure that the simulated device reflects realistic material properties, enabling accurate prediction of optoelectronic performance and providing insight into how halide (Cl, Br) and A-site cation (Ga, In, Tl) substitutions influence solar cell efficiency.
This interface defect analysis in Table 7 is significant because the WS2/absorber junction plays a critical role in determining overall solar cell efficiency, particularly in double perovskite-based devices like D2CeX6. Interface defects can act as recombination centers, trapping photogenerated carriers before they contribute to the photocurrent, which reduces the open-circuit voltage (VOC), short-circuit current density (JSC), and ultimately the power conversion efficiency (PCE).97 By quantifying parameters such as capture cross-section, defect type, and defect density, the simulation can realistically model the carrier recombination behavior at the heterojunction.
| Interfaces | Capture cross section: Electrons/holes (cm2) | Defect type | Total interface defect density (cm−2) |
|---|---|---|---|
| WS2/Ga2CeCl6 | 1 × 10−19 | Neutral | 1 × 1011 |
| WS2/Ga2CeBr6 | 1 × 10−19 | Neutral | 1 × 1011 |
| WS2/In2CeCl6 | 1 × 10−19 | Neutral | 1 × 1011 |
| WS2/In2CeBr6 | 1 × 10−19 | Neutral | 1 × 1011 |
| WS2/Tl2CeCl6 | 1 × 10−19 | Neutral | 1 × 1011 |
| WS2/Tl2CeBr6 | 1 × 10−19 | Neutral | 1 × 1011 |
Table 7 summarizes the interface defect parameters used in the SCAPS-1D simulations for the heterojunctions between the electron transport layer (WS2) and the absorber layers Ga2CeCl6, Ga2CeBr6, In2CeCl6, In2CeBr6, Tl2CeCl6, and Tl2CeBr6. For all interfaces, the capture cross-section for both electrons and holes is set at 1 × 10−19 cm2, representing a relatively low probability of carrier trapping, which helps to minimize interface recombination losses. The defect type is assigned as neutral, indicating that these traps do not possess a fixed positive or negative charge state, thereby reducing the likelihood of strong coulombic scattering. The total defect density is maintained at 1 × 1011 cm−2 for all combinations, corresponding to a moderate but realistic level of interface states that could arise from lattice mismatch or imperfections at the WS2/absorber junction. Such parameter uniformity across the Ga, In, and Tl-based D2CeX6 double perovskite absorbers ensures a controlled comparison of their intrinsic optoelectronic performance, allowing any variations in simulated device efficiency to be attributed primarily to the absorber material properties rather than interface quality differences.
![]() | ||
| Fig. 22 Energy band diagrams of (a) Ga2CeCl6, (b) Ga2CeBr6, (c) In2CeCl6, (d) In2CeBr6, (e) Tl2CeCl6, and (f) Tl2CeBr6 double perovskite absorber layers in the simulated solar cell structures. | ||
While the transport layers (FTO/WS2) maintain consistent band offsets across all structures, the internal energy level positioning within the absorbers varies considerably, which directly affects charge separation, extraction, and recombination dynamics. Overall, Ga2CeCl6 and Tl2CeCl6 are promising candidates for high-VOC applications due to their wider band gaps and reduced thermalization losses, whereas In2CeBr6, with the narrowest band gap, offers strong potential for near-infrared light harvesting but requires careful optimization to suppress recombination and enhance device performance.
![]() | ||
| Fig. 23 Optimization of absorber layer thickness on PCE, FF, JSC, and VOC for (a) D2CeCl6 and (b) D2CeBr6 double perovskite materials. | ||
The VOC remains nearly constant across thickness variations, emphasizing that it is primarily dictated by the intrinsic bandgap and defect-related recombination of the material rather than absorber thickness. The simulated VOC values are 1.259 V (Ga2CeCl6), 1.163 V (In2CeCl6), 1.297 V (Tl2CeCl6), 0.913 V (Ga2CeBr6), 0.497 V (In2CeBr6), and 0.969 V (Tl2CeBr6). These values confirm that Cl-based absorbers favor higher VOC, whereas Br-based compounds favor higher JSC, consistent with bandgap-dependent photovoltaic theory. Regarding the FF, increasing thickness from 0.2 µm to 0.8 µm causes slight decreases for most materials, reflecting the interplay between increased series resistance and enhanced absorption. FF decreases from 82.67% to 81.75% (Ga2CeCl6), 82.39% to 82.55% (In2CeCl6), 83.26% to 82.44% (Tl2CeCl6), 80.13% to 79.03% (Ga2CeBr6), 71.35% to 64.56% (In2CeBr6), and 77.03% to 76.91% (Tl2CeBr6). When the thickness is further increased to 2.3 µm, FF stabilizes for most absorbers, except In2CeBr6, which decreases from 64.56% to 58.82% due to increased bulk recombination and resistive losses. Overall, these results demonstrate that both cation choice (Ga, In, Tl) and anion substitution (Cl vs. Br) significantly influence thickness-dependent performance: Br-based perovskites achieve higher JSC owing to narrower band gaps, while Cl-based perovskites retain higher VOC, and the optimal thickness for balancing light absorption, carrier transport, and recombination generally lies at 0.8 µm for all materials.
![]() | ||
| Fig. 24 Optimization of total defect density on PCE, FF, JSC, and VOC for (a) D2CeCl6 and (b) D2CeBr6 double perovskite materials. | ||
However, when the defect density exceeds 1014 cm−3, a noticeable performance degradation occurs. At a higher defect density of around 1018 cm−3, this decline becomes severe, with significant reductions observed in all key photovoltaic parameters. The PCE, FF, JSC, and VOC values drop sharply for each material, indicating increased non-radiative recombination and hindered charge transport. These steep declines stem from the fact that higher defect concentrations introduce deep-level trap states, which increase carrier recombination rates, lower carrier diffusion lengths, and reduce quasi-Fermi level splitting, thereby strongly suppressing both VOC and JSC. Overall, the analysis highlights that cation (Ga → In → Tl) and anion (Cl → Br) substitutions significantly affect defect tolerance, with Cl-based compounds generally showing better stability against defect-induced performance loss compared to their Br counterparts. This insight provides critical guidance for material selection in designing defect-tolerant halide double perovskite solar cells.
As illustrated in Fig. 25(b), increasing the NA from 1013 to 1017 cm−3 results in nearly constant JSC values for most of the investigated D2CeX6 solar cells, except for In2CeBr6. The JSC remains approximately 20.05 mA cm−2 (Ga2CeCl6), 25.36 mA cm−2 (In2CeCl6), 19.33 mA cm−2 (Tl2CeCl6), 35.07 mA cm−2 (Ga2CeBr6), and 32.37 mA cm−2 (Tl2CeBr6), while In2CeBr6 shows a slight increase from 50.35 to 51.25 mA cm−2 over the same NA range. With further doping up to 1020 cm−3, JSC values largely remain constant, except for slight reductions observed in Ga2CeCl6 (18.70 mA cm−2) and Tl2CeCl6 (16.53 mA cm−2). This indicates that JSC in these perovskite systems is less sensitive to shallow acceptor density variations compared to VOC and FF, suggesting that carrier collection efficiency remains largely unaffected by excess doping, except in specific cases where recombination effects begin to dominate.
![]() | ||
| Fig. 25 Optimization of shallow acceptor density on (a) VOC, (b) JSC, (c) FF, and (d) PCE for D2CeX6 double perovskite materials. | ||
As illustrated in Fig. 25(d), the variation of PCE exhibits a similar trend to that of VOC and FF. When the acceptor concentration (NA) increases to around 1017 cm−3, the PCE values show a notable improvement for both chloride- and bromide-based devices. Upon further increasing NA to approximately 1020 cm−3, the efficiencies rise even higher across all configurations, with Ga2CeBr6 and Tl2CeBr6 achieving the best performance. This overall enhancement with increasing NA suggests improved charge transport and reduced recombination losses, leading to more efficient device operation. These results indicate that moderate NA of 1017 cm−3 significantly enhances device performance by improving carrier concentration and transport, while excessive doping continues to raise PCE but may approach saturation due to increased recombination losses. The optimized NA for VOC, JSC, FF, and PCE was found to be 1017 cm−3 for all the studied D2CeX6 materials. At this NA value, the devices achieve a balanced enhancement of carrier concentration and transport, resulting in maximum overall photovoltaic performance before recombination losses from excessive doping become significant.
![]() | ||
| Fig. 26 Temperature-dependent photovoltaic parameters of (a) VOC, (b) JSC, (c) FF, and (d) PCE of D2CeX6-based solar cells. | ||
Finally, Fig. 26(d) reveals that PCE declines with rising temperature for all materials, with Tl2CeCl6 and Tl2CeBr6 retaining superior efficiencies compared to Ga- and In-based counterparts, suggesting that Tl-containing perovskites offer a better balance between voltage, current, and charge transport. This analysis is crucial because it reveals how intrinsic material properties and halide composition govern the thermal stability and operational performance of perovskite solar cells, aiding in identifying the most thermally robust candidates for high-temperature photovoltaic applications.
Fig. 27(a) and (b) depict the J–V and QE characteristics of D2CeX6-based double perovskite solar cells. The optimized Al/FTO/WS2/D2CeX6/Au configuration was obtained using carefully tuned parameters: an absorber thickness of 800 nm, a total defect density of 1014 cm−3, and a shallow acceptor density of 1017 cm−3. As shown in Fig. 27(a), the simulated J–V curves yield JSC of 20.05, 25.35, 19.33, 35.07, 51.25, and 32.37 mA cm−2 for Ga2CeCl6, In2CeCl6, Tl2CeCl6, Ga2CeBr6, In2CeBr6, and Tl2CeBr6, respectively, under zero voltage. With increasing voltage, JSC initially remains nearly constant but then decreases sharply to zero at the corresponding VOC of 1.259, 1.163, 1.297, 0.913, 0.497, and 0.969 V. This reduction in photocurrent at higher voltage is primarily attributed to carrier saturation and recombination effects. As shown in Fig. 27(b), the simulated QE spectra exhibit an initial maximum value of nearly 100% for all compositions at short wavelengths, indicating excellent photon-to-carrier conversion efficiency under strong absorption conditions. With increasing wavelength, QE remains relatively constant over the visible region but decreases sharply to zero at the respective cutoff wavelengths of 720, 800, 710, 980, 1450, and 910 nm for Ga2CeCl6, In2CeCl6, Tl2CeCl6, Ga2CeBr6, In2CeBr6, and Tl2CeBr6, respectively. These cutoff positions are strongly correlated with the optical band gaps, where larger band gaps (e.g., Ga2CeCl6, Tl2CeCl6) correspond to shorter cutoff wavelengths, while smaller band gaps (e.g., In2CeBr6) extend absorption into the near-infrared region. The observed trends highlight the role of cationic substitution (Ga → In → Tl) and halide exchange (Cl → Br) in tailoring the optical response. Replacing Cl with Br reduces the band gap, red-shifting the absorption edge and enhancing long-wavelength QE, thereby improving light-harvesting capability. Similarly, the choice of the trivalent cation influences orbital hybridization and band dispersion, which governs carrier generation and transport. These theoretical insights confirm that compositional engineering in D2CeX6 double perovskites offers a viable pathway to optimize the spectral response and photovoltaic performance of the devices.
![]() | ||
| Fig. 28 Comparison of VOC, JSC, FF, and PCE for Ga2CeBr6, Ga2CeCl6, In2CeBr6, In2CeCl6, Tl2CeBr6, and Tl2CeCl6 absorber-based solar cells. | ||
A comparative evaluation highlights distinct trends associated with cationic and anionic substitutions. The chloride-based systems (Ga2CeCl6, In2CeCl6, Tl2CeCl6) generally provide higher VOC owing to their wider band gaps, although at the cost of reduced photocurrent. Among them, Ga2CeBr6 achieves the highest efficiency of 25.29%, reflecting a favorable balance between voltage and current density. The bromide-based systems (Ga2CeBr6, In2CeBr6, Tl2CeBr6), in contrast, show red-shifted absorption edges that enhance photocurrent but simultaneously lower the voltage due to narrower band gaps. In2CeCl6 demonstrates the best overall performance with an efficiency of 24.34%, arising from a well-balanced combination of VOC and JSC. In2CeBr6, despite delivering the highest photocurrent of 51.25 mA cm−2, suffers from severe voltage loss, restricting its efficiency to only 16.43%. Tl2CeBr6 exhibits intermediate behavior, with an efficiency of 24.13%, reflecting the role of cation substitution in fine-tuning electronic structure and carrier transport. These findings underscore that chloride-based compounds are favorable for higher voltage stability, while bromide substitution enhances photocurrent generation. Among the studied materials, Ga2CeBr6 emerges as the most promising candidate, demonstrating that halide exchange combined with careful cation selection can effectively optimize photovoltaic performance in D2CeX6 double perovskite solar cells.
Second, in the earlier version, the convergence criteria used for the phonon dispersion calculations were not sufficiently strict, and relatively small supercells were employed. As a result, the acoustic phonon frequencies at the Γ-point did not appear as zero, and slight imaginary modes were observed in some bromine-containing compounds. Although these issues have been significantly corrected in the revised calculations, very small negative or near-zero frequencies still appear near the Γ-point for Ga2CeBr6, In2CeBr6, and Tl2CeBr6. These features indicate possible soft phonon modes or minor lattice instabilities. In the future, the stability of these Br-based compounds may be enhanced through pressure, temperature effects, or doping. Finally, the phonon calculations were performed within the harmonic approximation at 0 K. Anharmonic effects, thermal stabilization, and pressure-dependent behavior were not considered and may influence the dynamical stability of the bromine-containing compounds.
m phase, with negative formation enthalpies and tolerance factors confirming thermodynamic and structural stability. Halide substitution (Cl → Br) enlarged lattice constants, narrowed band gaps, and improved visible-light absorption, while D-site cation variation (Ga → In → Tl) tuned bonding strength, charge distribution, and mechanical robustness. Electronic structure analysis revealed tunable direct and indirect band gaps (0.859 to 3.632 eV), with Ce-4f and halogen-p orbitals dominating near the band edges, supporting efficient charge-transfer transitions. Optical evaluations confirmed strong UV-visible absorption coefficients (α > 104 cm−1), high refractive indices, and large dielectric constants, especially in In2CeBr6 and Tl2CeBr6, making them promising for optoelectronics and photodetectors. Mechanical analysis showed ductile behavior across the series. Tl2CeBr6 had the highest bulk, shear, and Young's moduli, while In2CeBr6 exhibited excellent isotropy and flexibility suitable for stable thin-film or flexible solar cells. Phonon dispersion spectra confirm full dynamical stability for Ga2CeCl6, In2CeCl6, and Tl2CeCl6, while minor soft modes in bromides are expected to be stabilized at finite temperatures. Importantly, molecular dynamics simulations performed at 300 K, 600 K, and 900 K under NPT conditions demonstrate that all studied materials maintain thermal stability across a wide temperature range, highlighting their robustness for practical applications. Device simulations using SCAPS-1D demonstrated high photovoltaic potential. The device parameters were optimized with an absorber thickness of 800 nm, a total defect density of 1014 cm−3, and a shallow acceptor density of 1017 cm−3. In2CeCl6 reached an efficiency of 24.34%, while Ga2CeBr6 achieved the highest PCE of 25.29% due to a favorable trade-off between voltage and current. Bromide compounds enhanced photocurrent through band-gap narrowing, whereas chloride compounds delivered higher voltages. D2CeX6 family exhibits stability, tunable optoelectronic response, and strong photovoltaic performance, with In2CeCl6, Tl2CeBr6, and Ga2CeBr6 emerging as the most promising absorbers. These findings provide valuable guidance for experimental synthesis and compositional engineering of sustainable, lead-free perovskites for next-generation solar cells and optoelectronic technologies.
| This journal is © The Royal Society of Chemistry 2025 |