Open Access Article
Eilidh L. Quinn†
a,
Hugh Lohan†ab,
Elita Tmavaa,
Shiling Dong
a,
Aron Walsh
b and
Robert L. Z. Hoye
*a
aInorganic Chemistry Laboratory, University of Oxford, South Parks Road, Oxford, OX1 3QR, UK. E-mail: robert.hoye@chem.ox.ac.uk
bDepartment of Materials and Centre for Processable Electronics, Imperial College London, Exhibition Road, London, SW7 2AZ, UK
First published on 28th October 2025
Efforts to develop lead-free and stable alternatives to halide perovskites have thus far mostly yielded materials with power conversion efficiencies (PCEs) well below 10% in solar cells. Recently, photovoltaics based on CsBiSCl2 were reported to achieve 10.38% PCE. Still, the crystal structure is unknown, and it is unclear whether the reported thin film synthesis method could realize thin films with the desired phase and stoichiometry. Herein, we use ab initio Random Structure Searching (AiRSS) with a bespoke machine learned interatomic potential to explore the potential energy surface of CsBiSCl2, finding the previously-proposed cubic perovskite structure to be implausible. The lowest-energy structure we find is a four formula unit orthorhombic structure (Pnma space group) that lies 2.4 meV per atom above the convex hull. There is strong competition in the Cs–Bi–S–Cl family, which can lead to phase impurities. By examining the reported solution synthesis method, we find that it is challenging to obtain the dimethylammonium bismuth sulfide intermediate product, and that Bi2S3 with dimethylammonium iodide on the surface likely forms instead. The significant I-containing residues in this intermediate results in Cs3Bi2I9 being preferentially formed in thin films instead of CsBiSCl2. Solid state synthesis without I present leads to phase impurities, consistent with the lowest-energy CsBiSCl2 phase being metastable. Taking these experimental and computational results together, it is unlikely that >10%-efficient CsBiSCl2 solar cells have been achieved.
Broader contextSolar absorbers that are simultaneously efficient and able to be manufactured cost-effectively have long been sought after. Lead-halide perovskites (LHPs) were a surprising discovery that fulfils these requirements, and are now entering into commercial production. However, the toxicity and limited ambient stability of LHPs have prompted a search for alternatives that could mimic their exceptional optoelectronic properties, whilst overcoming their limitations. Most ‘perovskite-inspired’ materials thus far have yielded power conversion efficiencies (PCEs) <10% in solar cells. It is therefore remarkable that CsBiSCl2 was recently reported to achieve 10.38% PCE under 1-sun illumination, and maintain 97% of its original PCE after 150 days of storage in air without encapsulation. CsBiSCl2 is also free from any toxic elements, and is reported to have a bandgap of 2.0 eV, which is ideal for applications as the top-cell in tandem photovoltaics, for indoor photovoltaics, or for photoelectrochemical cells. However, the structure and synthesis of the material are poorly evidenced and understood. Herein, we use computational investigations to rigorously determine the lowest-energy structure for CsBiSCl2, and we show that the original synthesis method reported does not yield CsBiSCl2 thin films. We find the claimed 10.38% PCE to be spurious and recommend that the field avoids investing research efforts towards this material. In particular, we find that the lowest-energy structure of CsBiSCl2 is above the convex hull, making it challenging to avoid phase impurities from the Cs–Bi–S–Cl family of materials. |
Remarkably, Huang et al. reported a PCE of 10.4% from solar cells based on CsBiSCl2.17 Surprisingly, this high PCE was achieved in the first report of this material in solar cells. A PCE of 10% is considered a threshold for a new photovoltaic material to hold promise for commercialisation. Indeed, looking back at the development of LHPs: although the first publication of this material in solar cells was reported by Miyasaka and co-workers in 2009,18 it was not until Snaith and co-workers reported a PCE exceeding 10% in 201219 that the field began to gravitate towards this material.20–22 It is therefore significant that CsBiSCl2 was reported to immediately achieve >10% PCE in the first publication on this material for solar cells. Furthermore, this material was claimed to be substantially more environmentally stable than LHPs, with the PCE maintaining 97% of the original value after storage for 150 days without encapsulation in ambient air. CsBiSCl2 is also free from any toxic elements regulated by the Restriction of Hazardous Substances (RoHS) directive.23 The bandgap of 2.0 eV for this material is close to the ideal value for indoor photovoltaics,24,25 as well as X-ray detectors,26 and could also be well suited for solar water splitting.27
However, the crystal structure of CsBiSCl2 is not known, and the original report of CsBiSCl2 solar cells did not use any refinement methods to fit the measured X-ray diffraction pattern with a proposed structure.17 Understanding the structure of a material is critical. The crystal structure defines essential properties, including the stability, as well as the electronic structure, which in turn affects the bandgap and charge-carrier mobility (through the effective mass). Determining the structure is mandatory for performing atomistic simulations, including defect calculations, and determining the upper limit in efficiency of the material. Therefore, it is paramount to first understand the crystal structure of CsBiSCl2 and verify the stoichiometry of the compound formed.
Huang et al. suggested that CsBiSCl2 adopts an n = 1 Dion–Jacobson perovskite structure, comprised of layers of BiS2Cl4 corner-sharing units, with Cs+ cations in between these layers (Fig. 1a, left).17 However, there is no evidence for this structure. The general formula for the Dion–Jacobson homologous series is A′An−1PbnX3n+1 (A′ = 1+ or 2+ cation, A = 1+ cation, X = 1− anion).28 Adapting this formula to the Cs–Bi–S–Cl quaternary system depicted in Fig. 1a left, where n = 1, gives the stoichiometry CsBi(S,Cl)4, not CsBiSCl2. This CsBi(S,Cl)4 stoichiometry cannot be achieved with the standard oxidation states for the cations (I for Cs, III for Bi) and anions (II for S, I for Cl) without violating the rules of charge neutrality (details in Table S1, SI). If we take the structure as drawn in Fig. 1a left, then S
:
Cl should be 1
:
3, but this requires Bi to be in the IV oxidation state, which is unlikely given the stability of the 6s2 electron pair. Therefore, the suggestion that the polyhedra in CsBiSCl2 form layers with a corner-sharing Dion–Jacobson structure is unfounded.
![]() | ||
| Fig. 1 Structure of CsBiSCl2 (a) Proposed crystal structure of CsBiSCl2 by Huang et al.17 and Quarta et al.33 Illustration of structure proposed by Huang et al. reprinted with permission from J. Huang et al., J. Phys. Chem. Lett., 2024, 15, 3383–3389. Cubic perovskite structure prepared from the information provided in ref. 33. (b) Energy above the Cs–Bi–S–Cl convex hull of CsBiSCl2 structures selected for DFT refinalization. Shown are the lowest energy structure (yellow star), tetragonal perovskite structure (blue star), and remaining CsBiSCl2 structures within 70 meV per atom of the hull (orange circles). (c) Phonon dispersion curve of the lowest-energy CsBiSCl2 structure found from AiRSS. Lack of imaginary modes (i.e., no negative frequencies) demonstrates the dynamic stability of this structure. (d) Comparison of the simulated powder X-ray diffraction (PXRD) pattern for the lowest energy Pnma structure and the CsBiSCl2 PXRD pattern reported by Huang et al.17 Comparison with tetragonal perovskite reference pattern shown in Fig. S6, SI. | ||
Alternative structures for CsBiSCl2 may be more plausible. There have been studies into other I–III–VI–VII2 materials, with a particular focus on lead-free analogues of CH3NH3PbI3. In these materials, the Pb2+ cation is replaced with trivalent Sb3+ or Bi3+ and one I− anion is replaced with a chalcogenide anion (S2− or Se2−) to maintain charge neutrality, forming a split-anion perovskite.29 Films of CH3NH3SbSI2 have been reported with the suggestion that the phase adopts a near-ideal perovskite structure based on a near-ideal calculated Goldschmidt tolerance factor of t = 0.99.30 The Bi analogue, CH3NH3BiSI2, has also been reported, although it was proposed that this material preferentially adopts an orthorhombic or rhombohedral lattice as t < 0.9.31 However, further experimental and density functional theory (DFT) studies found no evidence for the formation of I–III–VI–VII2 perovskites, instead showing that the formation of the ternary A3M2X9 phase, along with binary chalcogenides and halides, is more energetically favourable.32
Quarta et al. attempted to prepare nanocrystals of CsBiSCl2.33 They proposed a cubic perovskite structure for this material (Fig. 1a, right) by using CrystaLLM, which is based on an autoregressive large language model, trained on millions of crystallographic information files.34 However, their synthesis attempts were unsuccessful, and CsBiSCl2 with a perovskite phase was not obtained experimentally, instead forming a mixture of Cs3BiCl6, Cs3Bi2Cl9 and Bi2S3. Another possibility is the UFeS3 structure type. This has been reported for both AgBiSCl2 nanocrystals and polycrystalline powders of CuBiSCl2.35,36 This is a layered structure with alternating slabs of [AS2Cl4]7− distorted octahedra, and [BiS2Cl6]7− bi-augmented triangular prisms. However, this structure has not yet been investigated as a possibility for the CsBiSCl2 phase.
Already, several review papers have cited CsBiSCl2 as an important prospect in lead-free photovoltaics due to the reportedly high performance and excellent cell stability.9,37–40 However, such claims are, as-yet, premature without understanding what the structure adopted is, and whether the reported thin film synthesis route17 is a valid approach to obtain this structure with the intended stoichiometry. It is critical to address these questions before the photovoltaics field dedicates more effort into this material.19–23
In this work, we first adopt a systematic global structure search to determine the ground state structure of CsBiSCl2. The approach is ab initio Random Structure Searching (AiRSS), coupled with a bespoke Ephemeral Data Derived Potential (EDDP) to determine whether (meta-)stable CsBiSCl2 structures exist and if previously-proposed perovskite structures are energetically accessible. Next, we repeat the synthesis method reported by Huang et al. for thin films. We examine the composition, and phases present in the reaction intermediate and final thin film product through energy-dispersive X-ray spectroscopy (EDX), as well as Pawley refinement of the measured diffraction patterns with possible structures for CsBiSCl2 and phase impurities. We use Fourier transform infrared spectroscopy (FTIR) to identify molecular species present, along with UV-visible (UV-vis) spectrophotometry to identify the absorption profiles of the thin film products formed. Given that the reported thin film synthesis route involves the use of HI, we compare with a solid-state melt synthesis approach that only uses the elements present in the target material, i.e., using CsCl, BiCl3 and Bi2S3 precursors. This work provides a thorough evaluation of the potential of CsBiSCl2 as a solar absorber, and whether the wider photovoltaics field should dedicate further effort into this novel lead-free material.
For relatively simple materials (unaries, binaries) with small unit cells, it is tractable to carry out all geometry relaxations using DFT. However, for compositionally complex materials where the search space is large, and/or those where a large unit cell is expected, traditional structure search soon becomes prohibitively expensive. Accelerating the structure search could be achieved using machine learning interatomic potentials (MLIPs). MLIPs can be classified as foundation or bespoke models, the former being trained on databases of diverse training data (e.g., Materials Project's MPtrj41–43) with the goal of being applicable to arbitrary chemical systems, while the latter are trained on much smaller and narrower datasets, usually prepared by a single researcher, with a specific system in mind (e.g. Cs–Bi–S–Cl). Bespoke models are not transferable to arbitrary systems.
We initially attempted to use the foundation model MatterSim44 to drive structural relaxations, but found it to have limited output structural diversity within the Cs–Bi–S–Cl system, as detailed in SI Note S1. We therefore trained a bespoke EDDP tailored to predicting the energy of CsBiSCl2 structures. EDDPs are formulated specifically for random structure searching and, by design, avoid common pitfalls associated with modelling far-from-equilibrium structures using MLIPs.45 To train the EDDP, the energies of 51
212 structures in the Cs–Bi–S–Cl space were calculated using DFT (PBEsol-D4). While the majority of structures were of the CsBiSCl2 stoichiometry, the dataset also included all low-energy Materials Project structures within the Cs–Bi–S–Cl space, optimised perovskite CsBiSCl2 structures, as well as a set of randomly generated unary, binary and ternary structures to improve the model fit. Crucially, this dataset included structures spanning the entire search space, from high-energy to near-equilibrium ones. The as-trained EDDP had final training root mean squared error (RMSE) and mean absolute error (MAE) of 48.16 meV and 30.98 meV, while the testing RMSE and MAE were 235.94 meV and 42.40 meV.
The testing RMSE was raised by a single outlier with an error of >15 eV. While the global RMSE/MAE is quite high, this is mostly increased by non-stoichiometric and far-from-hull structures included in the dataset. The testing RMSE/MAE for CsBiSCl2 structures within 1 eV of the hull are significantly improved at 22.7 meV per atom and 13.8 meV per atom respectively (Fig. S1, SI). Full details may be found in the Experimental section.
Using the as-trained EDDP, 92
770 CuBiSCl2 structures with 2–10 formula units were generated and relaxed to their local minima. Of these, 3732 structures within 70 meV per atom of the convex hull were selected, and similar structures unified using a similarity threshold of 0.01, resulting in 118 unique low energy structures, after removal of one unphysical ‘low-energy’ structure (Fig. S2, SI). The lowest energy structure was rediscovered eleven times, and is a four formula unit Pnma structure (Fig. 1b). This structure was significantly lower in energy (by 28.2 meV per atom) than the next lowest-energy structure.
In an attempt to catch any low energy structures missed by standard AiRSS, 25 ps molecular dynamics (MD) simulations near the melting point of the Pnma structure were carried out for the top 10
000 structures (2778 after unification) using the bespoke MLIP and supercells of at least 40 atoms (so called ‘hot-AiRSS’ method).46 While this reduced the number of unique structures after symmetrization and unification by 14% and resulted in the rediscovery of the Pnma structure another eight times; no lower energy structures were found. This supports the conclusion that the low-energy Pnma structure is dynamically stable. The structural density of states was flattened and broadened by this process, suggesting melting occurred in the far-from-hull regions (Fig. S3, SI).
The 118 unique low-energy structures, Materials Project (MP) near-the-hull structures, and the previously-proposed perovskite structures were selected for relaxation using precise DFT settings and more accurate Hamiltonian (r2SCAN-D3(BJ)+SOC). This high level of theory was chosen as the distance from the hull was within the margin of error set by our previous GGA/high-throughput DFT dataset. We found the lowest-energy Pnma structure predicted using EDDP-AiRSS remains the lowest energy CsBiSCl2 structure in the refinalized DFT dataset and sits just 2.4 meV per atom above the Cs–Bi–S–Cl convex hull (Table S2, SI). This implies it is an energetically accessible structure, but no stable CsBiSCl2 structures (i.e., directly on the convex hull of the Cs–Bi–S–Cl system) were discovered.
To confirm the dynamic stability of the low-energy Pnma structure, a tight relaxation and subsequent harmonic phonon dispersion curve was calculated (Fig. 1c). We found that, despite the presence of low-lying optical modes near the Z point, the structure is dynamically stable with no imaginary modes present. Due to its proximity to the convex hull and its dynamic stability, the low-energy Pnma structure is a plausible metastable synthetic target. We estimate this structure has a bandgap of >1.9 eV (Fig. S4, SI). Whilst this bandgap is suitable for top-cells in tandem photovoltaics, and for indoor photovoltaics, we note that using a meta-GGA functional would typically underestimate the bandgap, and there is a risk that the actual bandgap of the low-energy Pnma compound is substantially above 2 eV, making it not well suited for these applications. Furthermore, visual comparison of the simulated powder X-ray diffraction (PXRD) pattern of the low-energy Pnma structure with the PXRD pattern reported by Huang et al. for CsBiSCl2 showed very little similarity (Fig. 1d).
Interestingly, we found the previously-proposed cubic perovskite structure to be energetically implausible. Upon relaxation, the cubic perovskite structure undergoes a second-order Jahn–Teller distortion, resulting in a tetragonal structure where bismuth atoms are off-centred in their S2−/Cl− octahedra. This tetragonal perovskite has a large energy above hull of 123 meV per atom (Table S2, SI). In addition, visual comparison between the pattern reported by Huang et al. and the simulated pattern of the tetragonal structures shows little similarity (Fig. S6, SI). As such, we do not expect perovskite-structured CsBiSCl2 to be experimentally accessible using standard synthetic approaches.
The synthetic route to this ‘DMABiS2’ intermediate involves introduction of the dimethylammonium cation (DMA+) to Bi2S3 through in situ acid hydrolysis of the DMF amide bond, catalyzed by HI (Fig. 2a). The generation of DMA+ cations through DMF hydrolysis has been previously applied to the synthesis of Cs1−xDMAxPbI3 hybrid perovskites.47 There was no suggested structure for the previously unreported DMABiS2 intermediate. Bi2S3 has a 2D layered (lamellar) structure with layers connected by weak van der Waals interactions, with an interlayer spacing of 0.23 nm.48,49 Insertion of cations between the layers has been used to prepare intercalation compounds of Bi2S3 that have potential applications as battery materials.50,51 It may be the case that DMABiS2 is an intercalation product, with DMA+ cations between the layers of Bi2S3. If this were the case, it is to be expected that PXRD peaks assigned to Bi2S3 should shift to lower 2θ values.
To synthesize the DMABiS2 intermediate, we mixed 1.8 g Bi2S3 powder (3.5 mmol) with 3 mL DMF solvent in an N2-filled glovebox, according to the method from Huang et al. We found that full dissolution of Bi2S3 did not occur, nor during the addition of 2.5 mL HI dropwise to this mixture. After mixing for 4 h at 500 rpm, we used Büchner filtration of the mixture inside a fume hood, obtaining a dark brown intermediate product, which is consistent with the report from Huang et al.17 But we also obtained an orange precipitate from the filtrate after washing with ethanol. This was not reported in the original synthesis. Fig. 2b shows photographs of both the dark brown product powder and the orange precipitate recovered from the filtrate.
We measured the PXRD pattern of the dark brown product and orange powder from the filtrate, and analyzed through Pawley fitting (Fig. 2c). This refinement method fits the peak positions according to the structure files input, but numerically fits the peak intensities. Pawley fitting is therefore suitable for obtaining lattice parameters and identifying whether the measured pattern is accounted for by the proposed structures, but does not solve the atomic coordinates. This is an appropriate step for determining the phase purity of the powder before performing Rietveld refinement to obtain atomic coordinates, ideally from a phase-pure sample that does not have preferred orientation.
Given that the structure of DMABiS2 has not been solved, we first extracted the reported XRD peaks of this intermediate from ref. 17, and visually compared against the PXRD pattern for our dark brown product. As shown in Fig. 2c, there is a good match to all of the reference DMABiS2 peaks, especially the strong peaks at 12° 2θ, which were claimed by Huang et al. to “confirm” the formation of DMABiS2.17 There are, however, some peaks present in our dark brown product not present in the previously-reported pattern for DMABiS2, suggesting that there are other phases present, or that we do not have a DMABiS2 intercalation product.
We tested the hypothesis that the dark brown product is Bi2S3, performing Pawley fitting (Fig. S7a, SI) using the structure file from the Inorganic Crystal Structure Database (ICSD), collection code: 30775. As shown in Fig. 2a, there is a good match, with a goodness of fit (GoF) of 3.13. However, residuals are present, notably at 12°, 20°, and 30° 2θ. We therefore tested the hypothesis that a second phase was present: dimethylammonium iodide (DMAI). This improved the refinement (Fig. S7b, SI), with a goodness of fit of 2.99, although the key residuals at 12°, 20°, and 30° 2θ are still present. Therefore, we propose that the dark brown product we prepared is likely to contain two separate phases: Bi2S3 and DMAI, and possibly with a third unidentified phase. None of the peaks assigned to the Bi2S3 phase show a shift in 2θ position compared to the reference pattern. We therefore conclude that it is unlikely that there is an intercalation product.
SEM and EDX measurements were performed to aid the characterization of the prepared material. We observed two distinct phases: larger particles with a layered surface, and smaller particles dispersed across the surface of the larger particles. EDX analysis (Fig. S8 and Table S3, SI) suggested that the layered material was Bi2S3, with I impurities. The smaller particles contained a significant proportion of I and higher levels of N, suggesting that these smaller particles are likely to be primarily DMAI. This provides further evidence that no uniform DMABiS2 phase was synthesized, but that the reported synthetic route prepares Bi2S3 with DMAI surface contamination, along with I residues throughout the sample. We considered the possibility that BiSI may be present, however, Pawley refinement to BiSI structure file (ICSD collection code: 133376) revealed a poor match to this phase with a GoF of 4.79 (Fig. S9, SI). This allowed us to discount the presence of a BiSI phase.
Infrared (IR) spectroscopy data confirms the presence of an ammonium species in the dark brown product powder. Fig. 2d shows the IR spectrum of the material alongside the IR spectrum of reference DMACl material. The peak at 3146 cm−1 is assigned to the N–H stretch of the DMA+ cation. This is at a higher wavenumber than the corresponding N–H stretch peak in the DMACl reference (2980 cm−1). We attribute this observation to the reduction in electronegativity of the environment surrounding the DMA+ cation, since replacing Cl with I in DMAX would increase the electron density in the N–H bonds, making them stronger and oscillate faster.
We also analyzed the orange precipitate obtained from the filtrate. Its PXRD pattern is broadly similar to the pattern of the dark brown product, but significantly less crystalline, especially at higher 2θ values. We therefore concluded that this orange precipitate was not the desired DMABiS2 intermediate material. IR measurements of the orange precipitate (Fig. S10, SI) showed no distinct functional group features, providing further support that this precipitate was not the desired ammonium bismuth sulfide intermediate.
A significant limitation of this method for preparing the intermediate product is the high I contamination from DMAI. In an attempt to reduce this contamination, we increased the volumes of diethyl ether (DEE) and ethanol (EtOH) used to wash the powders from 25 mL to 50 mL, 75 mL, and 100 mL. Fig. 2e shows that despite increased washing volumes, iodine contamination was still significant, and no reduction in I wt% was observed from EDX of the dark brown powders. This may be because DMAI is strongly bound to the surface of Bi2S3, and cannot be removed through washing in polar solvents.
![]() | ||
| Fig. 3 Challenges with CsBiSCl2 synthesis (a) PXRD pattern of thin films synthesized according to Huang et al. (solid circles), with the corresponding Pawley refinement to Cs3Bi2I9 and Bi2S3 (red line); Rwp = 9.19 GoF = 1.21. Inset: photograph of the as-prepared thin film on 1.4 cm2 glass. (b) UV-vis absorption spectrum of prepared thin films and Cs3Bi2I9 reference spectrum,52 with the characteristic Cs3Bi2I9 excitonic peaks labelled. (c) PXRD pattern of solid residue from precursor solution from the attempted synthesis of CsBiSCl2 (solid circles), with the corresponding Pawley refinement to Bi2S3 and CsCl (red line); Rwp = 13.00 GoF = 2.14, unidentified peaks are labelled by *. Inset: photograph of the solid residue in a glass vial. (d) PXRD pattern of material from attempted solid-state synthesis of CsBiSCl2 (solid circles), with the corresponding Pawley refinement to CsCl, Bi2S3, and CsBi3S5 (red line); Rwp = 10.97, GoF = 1.74. | ||
Another problem with this solution synthesis approach is that the solid reagents did not completely dissolve in the DMF/DMSO mixture. The residual solid was collected, dried in a vacuum oven at 60 °C overnight, and analyzed by PXRD. Pawley refinement of this solid confirms that this powder is a mixture of Bi2S3 and CsCl (Fig. 3c).54–56 This is unsurprising given the poor solubility of CsCl in both DMF and DMSO due to its large lattice energy,57 and provides further support for the dark brown intermediate product being primarily Bi2S3 with DMAI attached.
No single-phase powders were obtained from this solid-state synthesis. Rietveld refinement (Fig. 3d) of the powders revealed a mixture of binary (CsCl, Bi2S3) and ternary (Bi3CsS5) phases (goodness of fit of 1.74). As shown in Table S2, SI, all of these phases are on the convex hull of the Cs–Bi–S–Cl system. No peaks corresponding to a CsBiSCl2 perovskite phase were observed in the powders. A lower temperature synthesis was also run at 175 °C, but this contained binaries and ternaries, as well as some unidentified products. This shows that for low-temperature solid-state synthesis, binary and ternary phases form preferentially over the desired quaternary phase. Thus, although the lowest-energy CsBiSCl2 phase is 2.4 meV per atom above the DFT convex hull, it still proves challenging to access this phase.
A typical AiRSS workflow involves four steps:
(1) Randomly generating a ‘sensible’ Cs–Bi–S–Cl structure using simple rules derived from chemical intuition, e.g., enforcing a minimum separation between atoms, fixed stoichiometry;
(2) Use a (DFT) energy/forces calculator to optimize this structure until forces on atoms are below a desired threshold;
(3) Add structure and to a ranked database of energies.
Steps 1–3 are repeated until some stopping criteria is reached, such as when the lowest energy structure has been rediscovered multiple times. At this point a subset of ‘interesting’ structures, usually those within 100 meV per atom of the convex hull, are refined using precise DFT settings.
An EDDP was iteratively trained using the ‘chain’ script and an initial dataset of 10
000 random CsBiSCl2 structures generated using AiRSS utility ‘buildcell’,61 as well as a ‘marker’ dataset of all <40 atom structures in the Cs–Bi–S–Cl phase space within 20 meV of the convex hull available on Materials Project.41,62,63 Marker structures were randomly shaken 50 times each with an amplitude of 0.04 Å. DFT single-point energies were calculated for each structure in the training set using CASTEP 24.164 and the high-throughput QC5 ultrasoft pseudopotential set with a cut-off energy of 425 eV, a minimum Monkhorst–Pack grid density of 0.05 Å−1, and a fine grid scale of 2. The PBEsol65 functional with D466 dispersion correction was used.
Approximately 256 DDPs were trained on this dataset using the ddp utility ‘forge’ trained and of these a subset were selected and cast into a weighted ensemble of DDPs (EDDP) by ‘flock’. DDPs were trained using up to 3-body terms and a cutoff of 8.0 Å, with a polynomial order of up to 16 for 2-body and up to 4 for 3-body terms, resulting in a feature vector with a length of 435. This was passed through a 5 node neural net and finally to the output layer.
Training, validation, and testing data was split in a ratio of 80
:
10
:
10. Using this EDDP and ‘repose’, 2000 random AiRSS structures were relaxed to their local minima. DFT total energies were calculated for these structures, added to the dataset, and used to train an improved EDDP. This iterative process was repeated a total of 5 times. To improve the performance of the MLIP, the training dataset was enriched with EDDP relaxed structures from the wider Cs–Bi–S–Cl space, as well as high and low density structures, resulting in a final dataset of 51
212 structures.
For the final MLIP, 213 DDPs were trained and 42 were selected and cast into an EDDP. To run the EDDP suite in a massively parallel way, the batch queuing utility ‘DDP-batch’ was utilised.67 Random structure search was carried out using the as trained EDDP and ‘repose’, a structural optimiser native to the EDDP suite. This trained MLIP, training dataset, and the ‘buildcell’/AiRSS input files used to generate it are available in the raw data repository. A total of 92
770 relaxations were carried out. Rediscovery rates were assessed by merging similar structures using the unite functionality of the ‘cryan’ tool in the ‘airss’ package, which compares unit cell volumes and interatomic distances within compared structures are within a specified threshold, e.g., 1%.
0.16 g of the dark brown intermediate product and 0.34 g CsCl were mixed in DMF/DMSO (9
:
1) and stirred at room temperature for 2 h. The suspension was allowed to settle for 24 h before the liquid phase was filtered with a 0.2 µm PTFE filter. A 1.2 cm × 1.2 cm substrate was cleaned by ultrasonication in acetone for 15 min, followed by isopropanol for 15 min. After cleaning, the substrates were blown dry with N2, followed by UV-ozone treatment (NanoBioAnalytics UV Ozone Cleaner UVC-1014) for 20 min. 40 µL of the filtered solution was spread onto the substrate followed by spinning at 1000 rpm for 10 s and 4000 rpm for 30 s. The deposited film was annealed at 220 °C for 1 h 20 min.
Supplementary information: details of random structure searching calculations, structure analysis and compositional analysis. See DOI: https://doi.org/10.1039/d5el00157a.
Footnote |
| † These authors contributed equally: Eilidh L. Quinn, Hugh Lohan. |
| This journal is © The Royal Society of Chemistry 2025 |