The stability of CsGeX3 (X = I, Br, Cl) selectively tuned by the crystal structures and halide ions as inferred from the calculated phonon spectrum†
Received
28th July 2024
, Accepted 20th November 2024
First published on 21st November 2024
Abstract
In spite of the considerable advancements achieved in enhancing the power conversion efficiency (PCE) of lead-based all-inorganic perovskite solar cells, there persists a need for materials that are both more stable and environmentally friendly. This investigation systematically explores the structural and thermodynamic stability, and electronic properties of Ge-based all-inorganic perovskite CsGeX3 (X = Cl, Br, I) in two space groups, Pm
m and R3m, utilizing first-principles calculations. Introducing the novel concept of the “imaginary frequency coefficient” alongside the tolerance factor and stabilizing the chemical potential window, we collectively characterize the stability of CsGeX3 based on the phonon spectrum and phonon density of states calculations. The findings reveal that the stable phase of the Ge-based perovskite differs from that of lead-based systems, with the R3m structure of CsGeX3 being the most stable in the rhombohedral phase. Moreover, the stability of R3m-CsGeX3 can be manipulated by adjusting the halide composition with a gradual increase in stability observed as halogen atoms shift from I to Cl. This comprehensive approach, integrating the phonon spectrum, innovative measurement indicators, and tolerance factor, presents an effective strategy for designing materials that are both non-toxic and stable.
1. Introduction
Organic–inorganic halide perovskite solar cells (PVSCs), denoted by the chemical formula ABX3, represent a highly promising class of renewable energy materials, showcasing a remarkable power conversion efficiency (PCE) of up to 26.41%.1–5 However, their thermal instability and light instability have always impeded the widespread commercialization of perovskite solar cells. A pivotal advancement in addressing this challenge emerged in 2015 when Eperon et al.6 pioneered the development of all-inorganic perovskite solar cells based on CsPbI3. Subsequent investigations revealed that the incorporation of metal cations in lieu of organic cations at the A-site significantly enhances thermal and phase stability, propelling the rapid advancement of all-inorganic perovskite technology.7–10 Notably, the most recent PCE milestone has surged to 21.15%,11 positioning all-inorganic perovskite materials in structures like ABX3 (A = Cs; B = Ge, Sn, Pb; X = Cl, Br, I) at a focal point of contemporary research. This achievement is particularly significant as it approaches the current highest photoelectric conversion rate of organic–inorganic hybrid perovskite solar cells. Moreover, in response to concerns regarding the toxicity of Pb and the susceptibility of Sn to oxidation,12,13 researchers have turned their attention to Ge-based perovskites. This strategic shift has not only resulted in groundbreaking power conversion efficiency (PCE) surpassing 10% but has also demonstrated remarkable stability, enduring beyond 500 hours, which offers a promising avenue to overcome environmental and stability challenges associated with traditional perovskite materials.14–16
Compared to the Pb-based perovskite systems, CsGeX3 (X = Cl, Br, I) materials exhibit distinctive structural characteristics. It has been reported in the literature that the Pb-based inorganic perovskite CsPbI3 manifests a tetragonal (β) phase under specific room temperature conditions and transforms into a cubic (α) phase at elevated temperatures.17,18 The substitution of Pb with Ge introduces structural modifications to CsGeX3, yet debates persist regarding the most stable configuration of Ge-based perovskites. In 1982, Guen et al.19 synthesized CsGeI3 crystals with the Pmmm space group for the first time. Following this, Thiele et al.20 synthesized CsGeX3 (X = Cl, Br, I) crystals and observed their phase transition behavior. At low temperatures, they exhibited a rhombohedral phase structure with the R3m space group, which transitioned to a cubic phase structure with the Pm
m space group as the temperature increased.20 In 2015, Sherburne et al.21 successfully synthesized CsGeI3 crystals with a stable R3m configuration and confirmed the absence of phase transitions within the operational temperature range of the device. In 2018, Wu et al.22 synthesized R3m-phase CsGeI3 perovskite nanocrystals and investigated their degradation mechanism in air, observing their eventual transformation into polycrystalline CsI. In 2021, Luo et al.23 computationally confirmed the R3m phase as the stable structure for CsGeI3, with the Pmmm phase being metastable. In the same year, Dias et al.'s24 theoretical calculations indicated that for CsGeX3 (X = Cl, Br, I), the Pm
m phase is not the lowest energy configuration, while the R3m phase demonstrates superior stability.
Analyzing phonon vibrations and phase diagrams serves as a valuable methodology for assessing thermal stability and phase transitions. Within the phase diagram, the polyhedron region delineated by the exclusion of competing phases signifies the establishment of a stable state.25 Concurrently, the presence of soft phonon modes in the phonon spectrum, characterized by imaginary frequencies, serves as a key indicator of structural instability.26 For instance, Huang et al. conducted an extensive investigation into the relationship between soft phonon modes and phase transitions in CsSnX3 (X = Cl, Br, I). They performed comprehensive calculations of the full phonon spectrum and the phonon density of states (PDOS) across various phases. Their findings revealed the exclusive manifestation of soft phonon modes in cubic and square phases, while their absence was noted in orthorhombic or monoclinic phases.27 Similarly, Zhu et al.28 demonstrated in 2019 that phonon calculations could unveil lattice instability resulting from dynamic disorder and anharmonicity of phonons in hybrid lead halide perovskites. Building upon these prior investigations, our study aims to explore the stable structure of all-inorganic Ge-based perovskites by analyzing the phase diagram and phonon dispersion of CsGeX3 (X = Cl, Br, I) in different phases and scrutinizing trends related to soft phonon modes.
In this work, we systematically studied the structure, phonons, thermodynamic stability, and electronic properties of Ge-based all-inorganic perovskite CsGeX3 (X = Cl, Br, I) in two space groups, Pm
m and R3m, using density functional theory (DFT) calculations. Firstly, we investigated the geometric structure and energy of both space groups and found that CsGeX3 in the R3m space group had lower energy due to the displacement of Ge ions. Secondly, we computed the phonon spectrum and phonon density of states (PDOS) of both space groups using the density functional perturbation theory (DFPT) method. For the first time, we introduced the concept of the “imaginary frequency coefficient” to quantify the stability of CsGeX3 in the two space groups, in conjunction with the tolerance factor value. The results indicated that CsGeX3 in the R3m space group exhibited a smaller imaginary frequency coefficient, suggesting a more stable structure. The stability could be adjusted by changing the halogen composition. Additionally, we systematically investigated the stability chemical potential window and found that the halogen composition could significantly regulate the size of the chemical potential window, effectively tuning the thermodynamic stability of CsGeX3. Lastly, we computed the electronic properties of CsGeX3 in both space groups and observed similar electronic band structures, although the R3m space group had a wider bandgap. This study provided a novel perspective for characterizing the stability of Ge-based all-inorganic perovskite materials by combining the tolerance factor with other parameters, contributing to the research and design of stable materials.
2. Computational methods
All first-principles calculations were performed using the projector augmented wave (PAW)29 method as implemented in the Vienna ab initio simulation (VASP)30 package, based on density functional theory (DFT). We chose the Perdew–Burke–Ernzerhof (PBE) generalized gradient approximation (GGA) as the exchange–correlation functional.31
For the calculation of the energy band structure, in the first Brillouin zone integration, we set the cut-off energy of the plane waves to 500 eV. We used a Γ-centered mesh with k-points set to 6 × 6 × 6.32 Highly symmetric k-points were introduced for the band structure calculation, including Γ(0,0,0), X(0,1/2,0), M(1/2,1/2,0), R/T(1/2,1/2,1/2), L(1/2,0,0), and F(1/2,0,1/2). All structures were relaxed using a conjugate-gradient algorithm until the energy on the ions was less than 1.0 × 10−5 eV and all forces were minimized to 0.02 eV Å−1. The initial wave function was randomly generated, using the superposition of atomic charge densities. Finally, the projected band structure of the material was plotted.
In order to obtain accurate phonon dispersion curves, it is necessary to first perform high-precision optimization of the crystal structure. Set the convergence energy of the system to 10−6 eV and progressively refine the accuracy from low to high until the atomic force convergence criterion reaches 10−6 eV Å−1. For the calculation of the phonon spectrum, we utilized the density functional perturbation theory (DFPT) method implemented in the PHONOPY code.33,34 The supercell was set at 2 × 2 × 2, and VASP was used to calculate the forces on atoms in the reconstructed supercell generated by PHONOPY to obtain the phonon vibrational properties and the phonon density of states (PDOS). To ensure convergence, we reduced the k-point sampling for force calculation based on the increase in supercell volume, ultimately setting the Γ-centered mesh k-point at 3 × 3 × 3. Finally, the corresponding phonon spectra and phonon density of states images of the materials were plotted. The crystal structures were constructed using the VESTA35 program.
3. Results and discussion
We investigated two crystal structures of CsGeX3 (X = Cl, Br, I): cubic phase (Pm
m space group, no. 221) and rhombohedral phase (R3m space group, no. 160). They are referred to as Pm
m-CsGeX3 and R3m-CsGeX3 in this paper. Fig. 1(a) and (b) illustrate CsGeCl3 as an example for comparing and explaining the differences between the two crystal structures of CsGeX3 materials. The detailed crystal structure parameters and schematic diagrams for all materials are respectively presented in Table 1 and Fig. S1 (ESI†). In the Pm
m-CsGeX3, the Ge ions occupy the geometric center of the GeX6 octahedron (0.5, 0.5, 0.5). The Ge ions and X ions form the GeX6 regular octahedral structure through shared angles, where the Ge–X bond lengths are equal (all six Ge–X bond lengths in CsGeCl3 are 2.67 Å). However, in the R3m-CsGeX3, the Ge ions undergo displacement in the direction of the body diagonal (0.516, 0.516, 0.516), resulting in distortion of the GeX6 octahedra. Three Ge–X bonds are compressed, while the other three Ge–X bonds are elongated (e.g., the longer and shorter Ge–X bond lengths in CsGeCl3 are 3.118 Å and 2.415 Å, respectively). Furthermore, in the Pm
m-CsGeX3, the lattice parameters a, b, and c are equal, and the Bravais lattice angles α = β = γ = 90°. In contrast, the lattice parameters and cell volume in the R3m-CsGeX3 are larger, and the Bravais lattice angles α = β = γ < 90°. We manipulated the positions of the Ge ions in the Pm
m-CsGeX3 crystal along the 〈111〉 crystal plane to achieve initial crystal structures with varying Ge ion displacement distances, the optimization of atomic positions and unit cell shape was subsequently conducted. Based on the changes in the crystal structure and energy depicted in Fig. 1(c), we can infer that the displacement of Ge ions along the 〈111〉 crystal plane may be the primary cause of the phase transition from Pm
m-CsGeX3 to R3m-CsGeX3. The stereochemically active lone pair (4s2) of Ge(II) cations can play a central role in promoting the Ge off-centering displacement.36 The parameters of these initial crystal structures are listed in Table S1 (ESI†). We further applied the static DFT method to calculate the relationship between the displacement of Ge along the 〈111〉 crystal plane from the geometric center of the octahedron and the relative energy ΔE, as shown in Fig. 1(d). The displacement of Ge resulted in a decrease in the total energy of the structure, with no energy barrier observed. This further supports the results in Fig. 1(c), indicating that the structural fluctuations caused by the displacement of Ge atoms lead to a transition of the Pm
m phase structure to a more stable configuration.
 |
| Fig. 1 The crystal structure of (a) Pm m-CsGeCl3 and (b) R3m-CsGeCl3, (c) the relationship between the displacement of Ge ions along the 〈111〉 crystal plane direction and the energy. The left inset in the figure illustrates the crystal structure when the Ge ion displacement is equal to 0 Å and 0.02 Å. The right inset in the figure depicts the relationship between the displacement of Ge ions and the variation in cell volume. (d) The relationship between the displacement of Ge along the 〈111〉 crystal plane from the geometric center of the octahedra and the relative energy ΔE. Here, ΔE is defined as the energy difference between the structure after the displacement of Ge atoms and the initial state. This result was obtained through static DFT calculations. | |
Table 1 Tolerance factor and the lattice parameters, Bravais lattice angles, cell volume, Ge–X bond lengths, and energy of Pm
m-CsGeX3 and R3m-CsGeX3. The bold values in the table represent experimentally measured data20
Space group |
Material |
a = b = c (Å) |
α = β = γ (deg) |
Volume (Å^3) |
D
Ge-X
s(l) Å |
Energy (eV) |
t
G
|
Pm m (cubic phase) |
CsGeCl3 |
5.339 |
90 |
152.204 |
2.670 |
−17.56 |
0.969 |
|
5.47
|
|
163.7
|
2.73
|
|
|
CsGeBr3 |
5.610 |
90 |
176.600 |
2.805 |
−15.90 |
0.954 |
|
5.69
|
|
184.2
|
2.84
|
|
|
CsGeI3 |
6.001 |
90 |
216.108 |
3.001 |
−14.10 |
0.934 |
|
6.05
|
|
221.4
|
3.02
|
|
|
R3m (rhombohedral phase) |
CsGeCl3 |
5.522 |
89.085 |
168.328 |
2.415(3.118) |
−17.76 |
0.969 |
|
5.434
|
89.72
|
160.44
|
2.348(3.092)
|
|
|
CsGeBr3 |
5.758 |
88.430 |
190.691 |
2.580(3.200) |
−15.98 |
0.954 |
|
5.635
|
88.74
|
178.65
|
2.534(3.116)
|
|
|
CsGeI3 |
6.115 |
88.362 |
228.435 |
2.792(3.359) |
−14.17 |
0.934 |
|
5.983
|
88.61
|
213.84
|
2.745(3.263)
|
|
|
The crystal structure differences between the Pm
m-CsGeX3 with the R3m-CsGeX3 resulted in distinct variations in structural stability. The tolerability factor (tG) is an important parameter for evaluating the structural stability of perovskite materials:
, where r is the radius of the corresponding ion.37,38 By varying the ion radii at the A, B, and X positions, the stability of perovskite materials can be adjusted. The commonly encountered ion radii for inorganic halide perovskites are Cs+ = 1.67 Å, Ge2+ = 0.73 Å, Cl− = 1.81 Å, Br− = 1.96 Å, I− = 2.20 Å.39 According to this formula, the tG values for the six CsGeX3 systems all fall within the range of 0.9–1.0, indicating that their crystal structures can stably exist in the perovskite structure. As the radius of the halide anions at the X position decreases (from I to Br to Cl), the tG value increases and gradually approaches 1, indicating that as the halide ion radius at the X position decreases, the CsGeX3 perovskite structure tends towards the ideal cubic perovskite structure. According to the results shown in Table 1, the R3m-CsGeX3 exhibits lower energy compared to the Pm
m-CsGeX3. This indicates that the R3m-CsGeX3 is more structurally stable. Further observations reveal that R3m-CsGeX3, as the radius of the halide anions at the X position decreases, the cell volume also decreases. The Bravais lattice angles gradually approach 90°, indicating a lattice contraction and a tendency towards a transformation from a lower symmetry to an ideal cubic cell structure. This crystal structure transformation aligns with the trend described by the tolerability factor regarding structural stability and contributes to the enhancement of stability. This strengthening may be linked to the strengthening of chemical bonding between the Ge ion and the X ion. As the electronegativity of the halide anions at the X position increases from I to Cl, the interaction between Ge and X becomes stronger, further enhancing the stability of the crystal structure. Therefore, halide ions play a significant role in regulating the stability of CsGeX3 materials.
To investigate the dynamical stability of CsGeX3 (X = Cl, Br, I), we calculated the phonon spectrum and phonon density of states for Pm
m-CsGeX3 and R3m-CsGeX3. The results are depicted in Fig. 2. Both Pm
m-CsGeX3 and R3m-CsGeX3 exhibit imaginary frequencies in their respective phonon spectrum. However, the difference lies in the distribution of the imaginary frequencies. In the Pm
m-CsGeX3, imaginary frequencies are observed in the optical branches and are distributed within the high-symmetry points of the Brillouin zone, specifically within the range of Γ–X–M–Γ. In contrast, in the R3m-CsGeX3, imaginary frequencies are only present in the acoustic branches and are located in the vicinity of two high-symmetry points, T and F. This demonstrates that both lattice structures in the two space groups show factors that lead to lattice instability. By analyzing the phonon density of states and the schematic representation of atomic vibration modes in Fig. S2 (ESI†), we observe that the lattice instability primarily originates from the GeX6 octahedra. Specifically, the instability of the Pm
m-CsGeX6 structure arises from the vibrations occurring in opposite directions between the Ge and halide atoms, while the instability of the R3m-CsGeX3 structure is attributed to the collective “swinging” vibrations of the Ge and halide atoms as an integrated unit. On the other hand, in the Pm
m-CsGeX3, as halide ions change from Cl to I, the absolute value of the imaginary frequency decreases. This factor explains the narrowing of the frequency distribution range in the phonon spectrum from CsGeCl3 to CsGeI3. For instance, the frequency distribution range for Pm
m-CsGeCl3 materials is −150 to 225 cm−1, while for Pm
m-CsGeI3 materials, it is −75 to 125 cm−1. Moreover, this indicates that Cs, as the heaviest atom, contributes solely to the phonon spectrum at lower frequencies.
 |
| Fig. 2 The phonon spectrums and phonon density of states (PDOS) for (a)–(c) Pm m-CsGeX3 (X = Cl, Br, I) and (d)–(f) R3m-CsGeX3. The acoustic branches are highlighted in red, with the PDOS plot vertically aligned to correspond to the phonon dispersion curve frequencies on the left. The gray area indicates the total phonon density of states, and the colored lines represent the contributions of individual ions. | |
In order to compare the differences in dynamic stability between Pm
m-CsGeX3 and R3m-CsGeX3 and obtain quantitative results from the phonon spectrum and phonon density of states, we adopted the concept of the “imaginary frequency coefficient” to represent the imaginary frequency information in the phonon density of states. The imaginary frequency coefficient (referred to as I) was defined by integrating the portion corresponding to imaginary frequencies with the total phonon density of states curve.
fDos represents the phonon density of states as a function of frequency, while “
a” represents the lowest imaginary frequency at which the density of states is non-zero. This definition of the imaginary frequency coefficient primarily takes into account two points. Firstly, the magnitude of the imaginary frequency is an important parameter for dynamic stability. Stronger imaginary frequencies indicate poorer stability. Secondly, considering that the density of states is also a crucial parameter for dynamic stability, a higher number of states occupied by the imaginary frequency implies greater instability of the material. Hence, a larger imaginary frequency coefficient indicates a worse structural stability of the material. We calculated the imaginary frequency coefficients of
Pm
m-CsGeX
3 and
R3
m-CsGeX
3 combined with the tolerance factors, as shown in
Fig. 3. The imaginary frequency coefficient of
Pm
m-CsGeX
3 is significantly higher than that of
R3
m-CsGeX
3, indicating that
R3
m-CsGeX
3 exhibits stronger dynamic stability. It is noteworthy that in
R3
m-CsGeX
3, there is an increasing trend in the frequency coefficient as the halide ions change from Cl to I, leading to a decrease in the materials’ stability. This trend in variation is consistent with the trend observed when evaluating the material stability using the tolerance factor. Therefore, by considering both the frequency coefficient and the
tG, we can conclude that halide ions play a crucial role in regulating the stability of the material.
 |
| Fig. 3 The imaginary frequency coefficients and tolerance factors of CsGeX3 (X = Cl, Br, I) in the Pm m and R3m-CsGeX3. | |
To further investigate the effective regulation of halide ions on the stability of CsGeX3, computational phase diagram measurements were used to visualize the chemical potential stability regions for Pm
m-CsGeX3 and R3m-CsGeX3.7,40,41 The physical implications of eqn (1)–(4) are as follows: the existence of a region that satisfies the specified inequality constraints indicates that, under certain conditions, it is feasible to synthesize single-phase crystals free of impurities, thereby suggesting that they are thermodynamically stable in essence. Conversely, the absence of such a region implies that the synthesis of crystals is challenging, and their phase stability is questionable.42 Taking R3m-CsGeCl3 as an example, under thermodynamic equilibrium growth conditions, the existence of CsGeCl3 should satisfy eqn (1). To exclude the possible secondary phases CsCl and GeCl4, the following constraints must also be satisfied (2)–(4). The chemical potential of Cs and Ge satisfying eqn (1)–(4) are shown as the green region in Fig. 4(a). This chemical range indicates the growth conditions for synthesizing R3m-CsGeCl3 under equilibrium conditions.
| μCs + μGe + 3μCl = ΔHCsGeCl3 = −17.557 | (1) |
| μCs + μCl < ΔHCsCl = −6.639 | (2) |
| μCs + 4μGe < ΔHGeCl4 = −16.424 | (3) |
| μi < 0 (i = Cs, Ge, Cl) | (4) |
where
μCs,
μGe, and
μCl represent the chemical potentials of each constituent element relative to its most stable phase, Δ
H represents the formation enthalpy of the elemental material. Fig. S3 (ESI
†) shows the constraints for
Pm
m-CsGeX
3 and
R3
m-CsGeX
3. As illustrated in
Fig. 4,
R3
m-CsGeCl
3 exhibits the largest stable chemical potential window, while the respective windows gradually decrease for the materials with halide ion variation from Cl to Br and then to I, indicating a weakening of thermodynamic stability. This trend is also observed for
Pm
m-CsGeX
3. The chemical potential window of the same compound in the
Pm
m space group is smaller than that in the
R3
m space group, suggesting that
Pm
m-CsGeX
3 has lower thermodynamic stability than
R3
m-CsGeX
3.
Pm
m-CsGeI
3 exhibits no chemical potential window, indicating the inability to form a stable
Pm
m-CsGeI
3 material under these constraints.
 |
| Fig. 4 The thermodynamic stability diagrams of (a)–(c) CsGeX3 (X = Cl, Br, I) in the R3m-CsGeX3 as functions of the chemical potentials of μCs and μGe. | |
Finally, we calculated the electronic band structures and density of states (DOS) for CsGeX3 (X = Cl, Br, I) to understand their electronic properties. Fig. 5 illustrates the projected band structures and DOS. Firstly, we observe that the band structures of Pm
m-CsGeX3 and R3m-CsGeX3 are highly similar, with the main distinction being the splitting of the threefold degenerate CBM in the R3m-CsGeX3 due to crystal field splitting or reduced symmetry,43 and the degree of splitting becomes more pronounced as the halide ion at the X-site changes from Cl to I. Both Pm
m-CsGeX3 and R3m-CsGeX3 are direct band gap semiconductors, with the valence band maximum (VBM) and conduction band minimum (CBM) located at the T point (0.5, 0.5, 0.5). The VBM is mainly contributed by Ge-s and X-p orbitals, while the CBM is largely influenced by Ge-p orbitals. The band gap of Pm
m-CsGeX3 is smaller than that of R3m-CsGeX3, and their band gaps decrease with increasing halide ion radius.
 |
| Fig. 5 Electronic projected band structures and density of states (DOS) of (a)–(c) Pm m-CsGeX3 (X = Cl, Br, I) and (d)–(f) R3m-CsGeX3. The Ge-p, Ge-s, and X-p orbitals were represented by the colors blue, red, and green, respectively. The value of Eg indicated the magnitude of the band gap in the material. | |
4. Conclusions
In this comprehensive study, we systematically explored the structural, phononic, thermodynamic, and electronic properties of Ge-based all-inorganic perovskites CsGeX3 (X = Cl, Br, I) in the Pm
m and R3m space groups through rigorous first-principles calculations. The stability assessment of CsGeX3 was conducted utilizing a multifaceted approach, incorporating the tolerance factor coefficient (tG), imaginary frequency coefficient (I), and a stable chemical potential window. Our theoretical analysis discerned the stability disparity in CsGeX3 in the two space groups, revealing that R3m-CsGeX3 possesses heightened stability characterized by lower energy, a reduced imaginary frequency coefficient, and a broader stable chemical potential window. Additionally, we observed that the stability of R3m-CsGeX3 can be modulated by varying the halide composition, exhibiting a gradual increase in stability as the halogen atom transitions from I to Cl. This innovative approach, amalgamating novel measurement indicators such as phonon spectra with tolerance factors, serves as an effective strategy for designing materials with enhanced stability and reduced toxicity.
Author contributions
The supervision of this research project was undertaken by Tingting Shi, Weiguang Xie, and Pengyi Liu, who provided essential computational resources to support the project. The conceptualization of the project theme and the determination of the research approach were collaboratively accomplished by Tingting Shi, Tengcheng Huang, and Zheyu Zhang. Tengcheng Huang and Zheyu Zhang also undertook the literature review and organization, as well as the computation, graphical representation, and analysis of the research findings, and participated in the drafting and revision of the initial manuscript. Tingting Shi offered professional guidance in the writing process and played a significant role in the construction of the paper's structure and theoretical guidance throughout the research process. Baoyun Liang and Ang Li contributed to the literature review and organization, as well as the graphical representation and analysis of the experimental results. Xin Xu and Yujia Gao engaged in the analysis and discussion of the phonon spectrum computational results. Zhuxia Wu, Haiyan Li, and Fei Yuan participated in the analysis and discussion of the structural characteristics and electronic structure computational results of the materials. All authors have reviewed and approved the final manuscript.
Data availability
The data supporting this article have been included as part of the ESI.†
Conflicts of interest
The authors declare no competing financial interest.
Acknowledgements
Tengcheng Huang and Zheyu Zhang contributed equally to this work. This work was supported by the National Natural Science Foundation of China (Grants No. 61674070, 62174072, 11804117 and 21973034), the Guangdong Basic and Applied Basic Research Foundation (Grant No. 2019B151502049 and 2022A1515010276), and the Uniqueness and Innovation Projects for the Universities in Guangdong Province (Grant No. 2022KTSCX010). Computer time at the National Supercomputer Center in Guangzhou (NSCCGZ) and the High-Performance Computing Platform of Jinan University is gratefully acknowledged. T. S. and P. L. also acknowledge the Guangzhou Key Laboratory of Vacuum Coating Technologies and New Energy Materials (No. 201605030008).
References
- A. Kojima, K. Teshima, Y. Shirai and T. Miyasaka, Organometal halide perovskites as visible-light sensitizers for photovoltaic cells, J. Am. Chem. Soc., 2009, 131, 6050–6051 CrossRef CAS PubMed.
- F. Lin, D. Nordlund, Y. Y. Li, M. K. Quan, L. Cheng, T. Weng, Y. J. Liu, H. L. L. Xin and M. M. Doeff, Metal segregation in hierarchically structured cathode materials for high-energy lithium batteries, Nat. Energy, 2016, 1, 1–8 Search PubMed.
- W. S. Yang, J. H. Noh, N. J. Jeon, Y. C. Kim, S. Ryu, J. Seo and S. I. Seok, High-performance photovoltaic perovskite layers fabricated through intramolecular exchange, Science, 2015, 348, 1234–1237 CrossRef CAS.
- M. H. Kumar, N. Yantara, S. Dharani, M. Graetzel, S. Mhaisalkar, P. P. Boix and N. Mathews, Flexible, low-temperature, solution processed ZnO-based perovskite solid state solar cells, Chem. Commun., 2013, 49, 11089–11091 RSC.
- J. J. Zhou, L. G. Tan, Y. Liu, H. Li, X. P. Liu, M. H. Li, S. Y. Wang, Y. Zhang, C. F. Jiang, R. M. Hua, W. Tress, S. Meloni and C. Y. Yi, Highly efficient and stable perovskite solar cells via a multifunctional hole transporting material, Joule, 2024, 8, 1–16 CrossRef.
- G. E. Eperon, G. M. Paternò, R. J. Sutton, A. Zampetti, A. A. Haghighirad, F. Cacialli and H. J. Snaith, Inorganic caesium lead iodide perovskite solar cells, J. Mater. Chem. A, 2015, 3, 19688–19695 RSC.
- Y. X. Chen, T. T. Shi, P. Y. Liu, W. G. Xie, K. Chen, X. Xu, L. L. Shui, C. Q. Shang, Z. H. Chen, H. L. Yip, G. F. Zhou and X. Wang, The distinctive phase stability and defect physics in CsPbI2Br perovskite, J. Mater. Chem. A, 2019, 7, 20201–20207 RSC.
- J. Huang, H. Wang, C. Y. Chen, Y. Z. Tang, H. S. Yang, S. Z. Liu and D. Zhang, Adjustable Skeleton of Bilateral Lewis Base Passivator for CsPbI3 Perovskite Solar Cells with PCE over 20% and Superior Stability, Chem. Eng. J., 2023, 466, 143120 CrossRef CAS.
- M. Kulbak, S. Gupta, N. Kedem, I. Levine, T. Bendikov, G. Hodes and D. Cahen, Cesium enhances long-term stability of lead bromide perovskite-based solar cells, J. Phys. Chem. Lett., 2016, 7, 167–172 CrossRef CAS.
- C. Liu, W. Z. Li, C. L. Zhang, Y. P. Ma, J. D. Fan and Y. H. Mai, All-inorganic CsPbI2Br perovskite solar cells with high efficiency exceeding 13%, J. Am. Chem. Soc., 2018, 140, 3825–3828 CrossRef CAS.
- J. G. Wang, Y. H. Che, Y. W. Duan, Z. K. Liu, S. M. Yang, D. F. Xu, Z. M. Fang, X. R. Lei, Y. Li and S. Z. Liu, 21.15%-Efficiency and Stable γ-CsPbI3 Perovskite Solar Cells Enabled by an Acyloin Ligand, Adv. Mater., 2023, 35, 2210223 CrossRef CAS PubMed.
- P. Xu, S. Y. Chen, H. J. Xiang, X. G. Gong and S. H. Wei, Influence of defects and synthesis conditions on the photovoltaic performance of perovskite semiconductor CsSnI3, Chem. Mater., 2014, 26, 6068–6072 CrossRef CAS.
- Y. M. Lee, J. Park, B. D. Yu, S. Hong, M. Jung and M. Nakamura, Surface instability of Sn-based hybrid perovskite thin film, CH3NH3SnI3: the origin of its material instability, J. Phys. Chem. Lett., 2018, 9, 2293–2297 CrossRef CAS PubMed.
- M. Chen, M. G. Ju, H. F. Garces, A. D. Carl, L. K. Ono, Z. Hawash, Y. Zhang, T. Y. Shen, Y. B. Qi, R. L. Grimm, D. Pacifici, X. C. Zeng, Y. Y. Zhou and N. P. Padture, Highly stable and efficient all-inorganic lead-free perovskite solar cells with native-oxide passivation, Nat. Commun., 2019, 10, 1–8 CrossRef.
- M. Azadinia, M. Ameri, R. T. Ghahrizjani and M. Fathollahi, Maximizing the performance of single and multijunction MA and lead-free perovskite solar cell, Mater. Today Energy, 2021, 20, 100647 CrossRef CAS.
- D. Saikia, J. Bera, A. Betal and S. Sahu, Performance evaluation of an all inorganic CsGeI3 based perovskite solar cell by numerical simulation, Opt. Mater., 2022, 123, 111839 CrossRef CAS.
- C. C. Stoumpos and M. G. Kanatzidis, The renaissance of halide perovskites and their evolution as emerging semiconductors, Acc. Chem. Res., 2015, 48, 2791–2802 CrossRef CAS PubMed.
- J. A. Steele, H. D. Jin, I. Dovgaliuk, R. F. Berger, T. Braeckevelt, H. F. Yuan, C. Martin, E. Solano, K. Lejaeghere, S. M. J. Rogge, C. Notebaert, W. Vandezande, K. P. F. Janssen, B. Goderis, E. Debroye, Y. K. Wang, Y. T. Dong, D. X. Ma, M. Saidaminov, H. R. Tan, Z. H. Lu, V. Dyadkin, D. Chernyshov, V. V. Speybroeck, E. H. Sargent, J. Hofkens and M. B. J. Roeffaers, Thermal unequilibrium of strained black CsPbI3 thin films, Science, 2019, 365, 679–684 CrossRef CAS PubMed.
- L. Guen, P. Palvadeau, M. Spiesser and M. Tournoux, ChemInform abstract: the germanium(II) iodide-rubidium iodide and germanium(II) iodide-cesium iodide systems, Chem. Informationsdienst, 1982, 19, 1–10 CAS.
- G. Thiele, H. W. Rotter and K. D. Schmidt, Kristallstrukturen und phasentransformationen von caesiumtrihalogenogermanaten(II) CsGeX3 (X= Cl, Br, I), Z. Anorg. Allg. Chem., 1987, 545, 148–156 CrossRef CAS.
- M. Sherburne, S. Li, M. Asta, N. Mathews and S. G. Mhaisalkar, Lead-free germanium iodide perovskite materials for photovoltaic application, J. Mater. Chem. A, 2015, 3, 23829–23832 RSC.
- X. Wu, W. Song, Q. Li, X. Zhao, D. He and Z. Quan, Synthesis of lead-free CsGeI3 perovskite colloidal nanocrystals and electron beam-induced transformations, Chem. – Asian J., 2018, 13, 1654–1659 CrossRef CAS PubMed.
- T. Luo, Y. Fan Xia, J. W. Huang, X. Huang, Z. Y. Wu, Y. X. Chen, X. Xu, W. G. Xie, P. Y. Liu, C. Y. Hu, X. Lu and T. T. Shi, Different structural evolutions of inorganic perovskite CsGeI3, CrystEngComm, 2021, 23, 4917–4922 RSC.
- A. C. Dias, M. P. Lima and J. L. F. Da Silva, Role of structural phases and octahedra distortions in the optoelectronic and excitonic properties of CsGeX3 (X = Cl, Br, I) perovskites, J. Phys. Chem. C, 2021, 125, 19142–19155 CrossRef CAS.
- W. J. Yin, T. T. Shi and Y. F. Yan, Unusual defect physics in CH3NH3PbI3 perovskite solar cell absorber, Appl. Phys. Lett., 2014, 104, 063903 CrossRef.
- U. Jong, C. Yu, Y. Kye, S. Hong and H. Kim, Manifestation of the thermoelectric properties in Ge-based halide perovskites, Phys. Rev. Mater., 2020, 4, 075403 CrossRef CAS.
- L. Y. Huang and W. R. L. Lambrecht, Lattice dynamics in perovskite halides CsSnX3 with X = I, Br, Cl, Phys. Rev. B:Condens. Matter Mater. Phys., 2014, 90, 195201 CrossRef.
- T.
S. Zhu and E. Ertekin, Mixed phononic and non-phononic transport in hybrid lead halide perovskites: glass-crystal duality, dynamical disorder, and anharmonicity, Energy Environ. Sci., 2019, 12, 216–229 RSC.
- G. Kresse and D. Joubert, From ultrasoft pseudopotentials to the projector augmented-wave method, Phys. Rev. B:Condens. Matter Mater. Phys., 1999, 59, 1758–1775 CrossRef CAS.
- G. Kresse and J. Furthmüller, Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set, Phys. Rev. B:Condens. Matter Mater. Phys., 1996, 54, 11169–11186 CrossRef CAS PubMed.
- J. P. Perdew, K. Burke and M. Ernzerhof, Generalized gradient approximation made simple, Phys. Rev. Lett., 1996, 77, 3865–3868 CrossRef CAS PubMed.
- H. J. Monkhorst and J. D. Pack, Special points for Brillouin-zone integrations, Phys. Rev. B, 1976, 13, 5188 CrossRef.
- A. Togo and I. Tanaka, First principles phonon calculations in materials science, Scr. Mater., 2015, 108, 1–5 CrossRef CAS.
- A. Togo, F. Oba and I. Tanaka, First-principles calculations of the ferroelastic transition between rutile-type and CaCl2-type SiO2 at high pressures, Phys. Rev. B:Condens. Matter Mater. Phys., 2008, 78, 134106 CrossRef.
- M. Koichi and I. Fujio, VESTA: a three-dimensional visualization system for electronic and structural analysis, J. Appl. Crystallogr., 2008, 41, 653–658 CrossRef.
- D. H. Fabini, R. Seshadri and M. G. Kanatzidis, The underappreciated lone pair in halide perovskites underpins their unusual properties, MRS Bull., 2020, 45, 467–477 CrossRef.
- V. M. Goldschmidt, Die gesetze der krystallochemie, Naturwissenschaften, 1926, 14, 477–485 CrossRef CAS.
- A. Swarnkar, W. J. Mir and A. Nag, Can B-site doping or alloying improve thermal-and phase-stability of all-inorganic CsPbX3 (X = Cl, Br, I) perovskites?, ACS Energy Lett., 2018, 3, 286–289 CrossRef CAS.
- R. D. Shannon, Crystal physics, diffraction, theoretical and general crystallography, Acta Crystallogr., Sect. A, 1976, 32, 715–767 CrossRef.
- D. Han, T. Zhang, M. L. Huang, D. Y. Sun, M. H. Du and S. Y. Chen, Predicting the thermodynamic stability of double-perovskite halides from density functional theory, APL Mater., 2018, 6, 084902 CrossRef.
- Y. T. Qu, X. Xu, Y. J. Gao, A. Li, X. Huang, Y. T. Cai, T. C. Huang, Z. X. Wu, P. Yi Liu, K. Q. Chen, T. T. Shi and W. G. Xie, Thermodynamic stability and electronic properties of 2D all-inorganic halogen perovskites, Phys. B, 2023, 668, 415210 CrossRef CAS.
- J. Buckeridge, D. O. Scanlon, A. Walshc and C. R. A. Catlow, Automated procedure to determine the thermodynamic stability of a material and the range of chemical potentials necessary for its formation relative to competing phases and compounds, Comput. Phys. Commun., 2014, 185, 330–338 CrossRef CAS.
- L. Y. Huang and W. R. L. Lambrecht, Electronic band structure trends of perovskite halides: Beyond Pb and Sn to Ge and Si, Phys. Rev. B, 2016, 93, 195211 CrossRef.
|
This journal is © the Owner Societies 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.