Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

First-principles study on enhancing the photocatalytic hydrogen evolution performance in Cs3Bi2I9/MoS2 heterostructure with interfacial defect engineering

Kyong-Mi Kima, Yun-Sim Kima, Dok-Ho Hyonb, Chol-Hyok Riac and Chol-Jun Yu*a
aComputational Materials Design, Faculty of Materials Science, Kim Il Sung University, Taesong District, Pyongyang, Democratic People's Republic of Korea. E-mail: cj.yu@ryongnamsan.edu.kp
bInstitute of Nano Engineering, State Academy of Science, Rakrang District, Pyongyang, Democratic People's Republic of Korea
cFaculty of Physics, O Jung Hup Chongjin University of Education, Chongjin, North Hamgyong Province, Democratic People's Republic of Korea

Received 22nd July 2025 , Accepted 5th September 2025

First published on 2nd October 2025


Abstract

Hydrogen has been attracting continuously growing interest as a highly efficient and clean energy source for replacing fossil fuels in the future, and thus developing highly efficient photocatalytic materials for hydrogen evolution is much desirable. In this work, we study the structural, electronic and optical properties of heterostructures composed of the bismuth-based vacancy-ordered iodide double perovskite Cs3Bi2I9 and a two-dimensional dichalcogenide 2H-MoS2 monolayer without and with a vacancy defect using first-principles calculations. Our calculations demonstrate that the Cs3Bi2I9/MoS2 heterostructures are energetically stable and induce an interfacial dipole moment, which is beneficial for the prevention of charge carrier recombination. Due to the proper band-edge alignment and the smallest Gibbs free energy difference for hydrogen adsorption, the defective interface with a Cs-vacancy (VCs) is found to be the most promising for photocatalytic hydrogen evolution. Moreover, we find that the interfacial VCs defect can be formed favourably under the I-rich/Cs-poor condition, where VI and VS formations are suppressed. This work provides a way to develop high-performance photocatalysts based on heterostructures composed of the Bi-based halide perovskites and transition metal dichalcogenides for hydrogen evolution from solar-driven water splitting.


1 Introduction

Superseding natural resources with clean sustainable energy sources has become an extremely urgent issue to meet the increasing demand for energy and achieve net-zero carbon emissions in light of the rapid exhaustion of fossil fuels and acute environmental problems. In this context, hydrogen has been attracting continuously growing interest as a highly efficient and clean energy source for the future.1 Among various methods for producing hydrogen, solar-driven photocatalytic hydrogen evolution is the most attractive approach in terms of energy balance and clean production due to solar energy utilization.2–4 Photocatalysts for hydrogen evolution reactions (HERs) reported so far include carbon nitrides,5–7 metal oxides,8,9 metal sulfides,10–12 organic materials,13,14 and heterojunctions.15,16 When using those photocatalysts, however, the solar-to-hydrogen (STH) conversion efficiencies are still far below the target threshold of ∼10% required for commercialization.17 To increase the STH efficiency towards 10%, many strategies have been developed, including the search for new materials, heteroatom doping, interface engineering and phase engineering.18

Since their first use as light absorbers in solar cells in 2009,19 halide perovskites (HPs) have been attracting significant attention in the field of photovoltaics, with high power conversion efficiencies exceeding 27% being reported.20–23 This is due to the superior optoelectronic properties of HPs, including strong light absorption,24 tunable bandgap,25 long carrier lifetime,26 long carrier diffusion lengths27,28 and high carrier mobility.29 The best-known photocatalysts for HERs are lead-containing HPs,30 which are usually used in the form of heterojunctions with other photocatalysts to prevent the serious recombination of photogenerated charge carriers. As reported by Zhao et al.,31 the organic–inorganic hybrid HP MAPbI3 (MA = CH3NH3, methylammonium), when integrated with a MoS2 monolayer, demonstrated a STH efficiency of 1.09% and prominent hydrogen generation activity (HGA) of 13.6 mmol g−1 h−1. By integrating a polyfluorene co-catalyst onto the surface of MAPbI3, Pal et al.32 reported a HGA of up to 6.2 mmol g−1 h−1 in HI solution, almost 200-fold higher than that of pristine MAPbI3. In spite of such outstanding photovoltaic and photocatalytic properties, the organic–inorganic hybrid HPs have suffered from poor stability due to the existence of the organic moiety.33 Therefore, the organic species (e.g., MA) were replaced with inorganic ones; for instance, CsPb(Br1−xIx)3 with a noble metal co-catalyst of Pt demonstrated a HGA value of 1.12 mmol g−1 h−1 in the HI/HBr mixed solution.34

In the Pb-based HPs, toxicity of lead is another problem, causing environmental and health care concerns. To address these problems, several metals, such as Sn, Ge and Bi, can be used instead of Pb in the all-inorganic HPs.35–37 Typically, Bi-based HPs have been utilized in photocatalytic and photovoltaic applications due to their optoelectronic properties being comparable with Pb-based ones, including good intrinsic stability and eco-friendliness.38,39 For instance, Cs3Bi2Br9 nanocrystals coupled with graphitic carbon nitride (g-C3N4) nanosheets were found to exhibit a good HGA of 4.6 mmol g−1 h−1.40 Moreover, Pancielejko et al.41 have demonstrated a 24-fold improvement in HGA by combining Bi-based HPs of Cs3Bi2X9 (X = I, Br, Cl) with different types of TiO2. Tang et al.42 also reported an improved photocatalytic performance for a Cs3Bi2I9/Ti3C2 composite due to the promoted charge transfer and separation. A stable composite of Cs3Bi2I9 and MoS2 quantum dots has also been reported to show great enhancement in photocatalytic performance,43 achieving a HGA of 6.09 mmol g−1 h−1, which notably surpasses that of pristine Cs3Bi2I9 and Pt/Cs3Bi2I9 composite by 8.8 and 2.5 times, respectively. Such an impressive improvement, leading to a new record for Bi-based HP photocatalysts under visible light, was attributed to the type-II heterojunction between Cs3Bi2I9 and MoS2.

The experimental findings mentioned above highlight the importance of interface engineering in enhancing the photocatalytic performance of Bi-based HP photocatalysts. To fully understand the underlying mechanism of such an enhancement through interface formation, a first-principles study based on density functional theory (DFT) is necessary.16 Since the photocatalytic performance encompasses both the intrinsic properties and the efficiency of a photocatalyst, one should clarify the electronic and optical properties such as band structure, band-edge alignment against the water redox potential, photoabsorption coefficients and the Gibbs free-energy difference upon hydrogen adsorption on the catalysis surface to estimate the photocatalytic performance, which further determines the STH conversion efficiency or HGA. To the best of our knowledge, however, no theoretical studies on the Cs3Bi2I9/MoS2 heterostructure have been reported to date, meaning that the effects of interface and defect engineering on its photocatalytic performance remain unclear. In this work, we perform first-principles calculations of the Cs3Bi2I9/MoS2 heterostructure to gain atomistic insights into the interfacial and defect effects on the electronic and optical properties. In particular, we consider typical vacancy defects at the interface, evaluating their formation energies, various interfacial properties and HER performance.

2 Methods

2.1 Computational methods

In the early steps, the DFT calculations were performed by using the pseudopotential plane-wave method as implemented in the Quantum ESPRESSO (QE, version 7.2) package.44 For describing the ion–electron interactions, we used the ultrasoft pseudopotentials provided in the GBRV library,45 which were generated with scalar relativistic calculations by using the valence electron configurations of Cs-5s25p66s1, Bi-6s26p35d10, I-5s25p5, Mo-4s24p65s24d4 and S-3s23p4. The Perdew–Burke–Ernzerhof (PBE) formalism46 within the generalized gradient approximation (GGA) was used for the exchange–correlation interaction. Since the interface systems under study include layered materials of MoS2 and Cs3Bi2I9, the van der Waals (vdW) interactions between the layers were taken into account by using the vdW-DF-ob86 functional.47 As the major computational parameters, the kinetic cutoff energies were set to 40 and 400 Ry for wave functions and electron density, respectively, and the special k-point meshes for the Brillouin zone sampling were set to (4 × 4 × 2). All the atomic positions were relaxed until the residual force on the atom converged to 5 × 10−4 Ry per Bohr, while the crystalline lattice was optimized until the stress was less than 0.05 kbar.

For the supercell calculations to consider the interfacial defects, the pseudo-atomic orbital (PAO) basis sets method as implemented in the SIESTA (version 4.1) package48 was used. We selected the double-ζ plus polarization (DZP) scheme for the PAO basis sets, which were generated using an energy shift of 300 meV for the orbital-confining cutoff radii and a split norm of 0.30 for the split-valence. The plane-wave cutoff energy was set to 100 Ry and the special k-points mesh was set to (4 × 4 × 1) for atomic relaxations and (6 × 6 × 1) for energetic and optical calculations. The atomic relaxations were performed until the atomic forces converged to 0.02 eV Å−1. We note that such calculations using the combination of QE and SIESTA packages were found to give reasonable results for different interface systems, with an affordable computational cost.16,49–51

The formation energy of a vacancy defect i with charge q was calculated as follows:52,53

 
Ef(i,q) = Etot(i,q) − Eperf + μi + q(EF + EVBM), (1)
where Eperf and Etot(i,q) are the DFT total energies of the perfect supercell and the defective supercell containing one vacancy defect i with a charge state q, EF is the Fermi energy referenced to the energy level of the valence band maximum (VBM) EVBM, and μi is the chemical potential of the i-th species removed from the supercell. The chemical potential μi can be written as μi = Ei + Δμi, where Ei is the total energy per atom of the corresponding simple substance. Under the thermodynamic equilibrium condition, the existence of bulk Cs3Bi2I9 and MoS2 puts constraints on the chemical potentials as follows:51,54,55
 
ΔECs3Bi2I9 = 3ΔμCs + 2ΔμBi + 9ΔμI, (2)
 
ΔEMoS2 = ΔμMo + 2ΔμS, (3)
where image file: d5ra05294g-t1.tif and image file: d5ra05294g-t2.tif are the formation energies of the Cs3Bi2I9 and MoS2 compounds, respectively. To prevent the formation of elemental bulk i, the condition Δμi < 0 should be satisfied. To ensure the formation of Cs3Bi2I9 from the binary compounds CsI and BiI3, the following constraints must be met:
 
ΔμCs + ΔμI < ΔECsI (4)
 
ΔμBi + 3ΔμI < ΔEBiI3 (5)

For other binary compounds, such as CsI3, CsI4, Bi9I2, Cs3Bi, CsBi and CsBi2, similar constraints are satisfied. Moreover, chemical dissociation of Cs3Bi2I9 into the binary constituents (CsI and BiI3) should be prohibited and thus the following condition should be satisfied:

 
ΔECs3Bi2I9 − 3ΔECsI < 2ΔμBi + 6ΔμI < 2ΔEBiI3 (6)

The hydrogen adsorption energy ΔEH is determined as56

 
image file: d5ra05294g-t3.tif(7)
where Esub+H and Esub are the total energies of the substrate systems with and without an adsorbed H atom on the surface, respectively, and EH2 is the total energy of an isolated H2 molecule placed in a supercell of the same size as the substrate system, which is large enough to neglect the artificial interactions between the molecule and its periodic images. Then, the Gibbs free energy of hydrogen adsorption ΔGH* can be evaluated as follows:
 
ΔGH* = ΔEH + ΔEZPETΔSH, (8)
where ΔEZPE is the difference between the zero-point energy (ZPE) of H2 in the adsorbed state and that in the gas-phase state, and ΔSH is the difference in entropy of H2 between the gas phase and adsorbed state. Here, ΔSH can be written as image file: d5ra05294g-t4.tif, where image file: d5ra05294g-t5.tif is the entropy of H2 under standard conditions. We utilized 0.24 eV for ΔEZPETΔSH, as reported in the previous work for hydrogen adsorption.56

The imaginary and real parts of the frequency-dependent complex dielectric function, ε(ω) = ε1(ω) + 2(ω), were calculated using the density functional perturbation theory (DFPT) as implemented in the SIESTA package.48,57 Then, the optical absorption coefficients α(ω) were obtained as follows:58–60

 
image file: d5ra05294g-t6.tif(9)
where c is the speed of light.

2.2 Interface modelling

As a layered vacancy-ordered double perovskite, Cs3Bi2I9 is crystallized in a hexagonal lattice with a space group of P63/mmc under ambient conditions,61 as shown in Fig. 1(a). The structure of Cs3Bi2I9 is characterized by two neighboring [Bi2I9]3− octahedra, which are face-shared along the c-axis to form a [Bi2I9]3− bioctahedral cluster, and the [Bi2I9]3− clusters are separated by hexagonal channels filled with Cs+ cations, leading to the formation of a zero-dimensional molecular salt crystal structure. Within the [Bi2I9]3− bioctahedral cluster, there exist three long bridging Bi–I bonds, which involve I atoms on the shared face, and six terminal short Bi–I bonds oriented away from the shared face. For MoS2 bulk, we selected the 2H phase among its polymorphs, which is also in a hexagonal lattice with the space group of P63/mmc. With the optimized bulk structures, we built the interface slab models of Cs3Bi2I9 (001) surface and MoS2 (001) monolayer (ML). According to the experimental report for the structural characteristics,62 the Cs3Bi2I9 (001) surface was suggested to have CsI3 termination. To minimize the lattice mismatch between the Cs3Bi2I9 (001) and MoS2 (001) surfaces, we applied the coincidence lattice method.63 As a result, we obtained the surface supercells along the lines of [310] and [230] for 2H-MoS2 (Fig. 1(c)) and [100] and [110] for Cs3Bi2I9 (Fig. 1(d)), with a significantly small lattice mismatch of 0.1% evaluated as follows:64
 
image file: d5ra05294g-t7.tif(10)
where a1 and a2 are the lattice constants of the Cs3Bi2I9 and 2H-MoS2 lattices. The resultant supercells correspond to (1 × 1) and (7 × 7) for Cs3Bi2I9 (001) and 2H-MoS2 (001) surfaces, respectively. The slab models of the interface contained ten atomic layers of Cs3Bi2I9 and an S–Mo–S layer of MoS2 (49 atoms) with a vacuum layer of 15 Å thickness, as shown in Fig. 1(e). We confirm that the 10 atomic layers for the Cs3Bi2I9 is thick enough to guarantee convergence of the surface energy and the 15 Å vacuum thickness is sufficiently large to inhibit the artificial interactions between the top and bottom atoms of the neighbouring slabs. While fixing the lattice constants of the slab supercell and the atomic positions of the 6 atomic layers in the Cs3Bi2I9 side, all other atomic positions were relaxed using the QE code. Here, we considered two different possible sliding structures between Cs3Bi2I9 and MoS2, denoted as Conf1 and Conf2 (Fig. S2, SI), and performed atomic relaxations for these sliding configurations. From the calculation, the Conf1 sliding structure, where one of the S atoms in the MoS2 layer is placed on top of the Cs atom of the Cs3Bi2I9, was found to have a lower total energy with a shorter interlayer binding distance than the Conf2 structure. Therefore, we only considered the Conf1 sliding structure for further calculations of the electronic and photocatalytic properties.65

image file: d5ra05294g-f1.tif
Fig. 1 (a) Polyhedral view of the unit cell for the Cs3Bi2I9 crystal optimized with QE. (b) Electronic band structure of the bulk Cs3Bi2I9 calculated in this study. Top view of the lattice-matched structure of (c) the MoS2 (001) surface (7 × 7) supercell and (d) the Cs3Bi2I9 (001) surface unit cell. (e) Side view of the Cs3Bi2I9/MoS2 ML interface supercell model relaxed with SIESTA.

In order to consider the vacancy defects, we built 2 × 2 × 1 supercells (196 atoms) with a very thick vacuum layer (70 Å) using the relaxed interface slabs with the QE code. The atomic relaxations were also performed using the SIESTA code, while fixing the lattice constants and the 6 atomic layers of the Cs3Bi2I9 side. We confirm that the supercell sizes are large enough for defect calculations as used in the previous works.16,66 Then, each of the three types of interfacial vacancy defect, such as caesium (VCs), iodine (VI) and sulfur (VS), was created at different positions around the interface, and the lowest energy configuration for each defect was identified by performing the atomic relaxation (see Fig. S3 and Table S1, SI). We considered different charge states for each vacancy defect in this work.

The formation energy of the interface is defined as follows:67

 
image file: d5ra05294g-t8.tif(11)
where image file: d5ra05294g-t9.tif, image file: d5ra05294g-t10.tif and image file: d5ra05294g-t11.tif are the DFT total energies of the Cs3Bi2I9/MoS2 interface, the Cs3Bi2I9 (001) surface and the MoS2 (001) surface, respectively, and A is the interface area. The total energies for the surfaces were calculated by using the fully relaxed isolated surface systems with the same cell size with the interface. The interlayer binding energy is defined as follows:
 
image file: d5ra05294g-t12.tif(12)
where Eint(deq) and Eint(d) are the total energies of the interface systems with the equilibrium interlayer distance and the infinity distance, respectively, and N is the number of atoms in the MoS2 layer. The spatial charge density difference was evaluated as follows:
 
Δρ = ρintρCs3Bi2I9ρMoS2, (13)
where ρint, ρCs3Bi2I9 and ρMoS2 are the charge densities of the interface, isolated Cs3Bi2I9 and MoS2 surface sides, respectively. Then, the planar-averaged charge density difference integrated on the xy plane was calculated as follows:
 
image file: d5ra05294g-t13.tif(14)
where A is the area of the xy plane.16,51

3 Results and discussion

3.1 Structural stability of the interface

Firstly, we scrutinized the geometrical and electronic properties of bulk Cs3Bi2I9 and 2H-MoS2. By performing the structural optimizations of the unit cells, we determined their lattice constants to be a = b = 8.371 Å, c = 21.132 Å for Cs3Bi2I9 and a = b = 3.167 Å, c = 12.374 Å for 2H-MoS2, which agree well with the previous experimental results.61,68,69 Using the QE code with the PBE functional, we then calculated their electronic band structures. Fig. 1(b) shows the calculated band structure of Cs3Bi2I9, demonstrating an indirect bandgap of 1.97 eV in good agreement with the previous results.61,70 Meanwhile, the band structure of the 2H-MoS2 monolayer shows a direct bandgap of 1.84 eV, also in good agreement with the previous DFT calculations69 (see Fig. S1, SI). Such good agreements indicate that the computational settings employed in this work are reasonably acceptable. We note that the GGA-PBE functional was found to give bandgap values that agreed well with the experimental ones for lead halide perovskites, due to the fortuitous error compensation between the underestimation of PBE and overestimation by ignoring the spin–orbit coupling effect.71,72

With these calculated lattice constants of the bulk materials, the cell dimensions of the Cs3Bi2I9/MoS2 interface were found to be 8.375 × 8.375 Å and those of the 2 × 2 supercells was 16.75 × 16.75 Å. To estimate the stability of the Cs3Bi2I9/MoS2 interface system, the formation energy Ef was calculated. The formation energy was found to be negative, indicating that the formation of the Cs3Bi2I9/MoS2 interface is exothermic and that binding between the constituent Cs3Bi2I9 surface and MoS2 monolayer is energetically favorable. In fact, the formation energy of the perfect heterostructure without any defects was calculated to be −0.70 J m−2, being lower than that of the CsPbI3/MoS2 interface obtained in our previous work16 (−0.22 and −0.37 J m−2 for CsI and PbI2 terminations, respectively). This indicates the stronger binding between the constituent layers in the Cs3Bi2I9/MoS2 heterostructure than in the CsPbI3/MoS2 heterostructure.

To further assess the binding strength between the constituent layers of the interface, the interlayer binding energy was also calculated. Fig. 2 shows the resultant Eb as a function of interlayer distance dint. Table 1 lists the calculated values for the interface formation energy, interlayer binding energies and interlayer equilibrium distances for the Cs3Bi2I9/MoS2 heterostructures considered in this work. Irrelevant to the existence of vacancy defects, the binding energy Eb values for the interface systems were found to be negative, indicating the attractive binding in the Cs3Bi2I9/MoS2 interface systems. The binding energy for the perfect system was estimated to be −127.81 meV per atom, which is comparable to that of the CsPbI3/MoS2 interface.16,73 Among the different defective systems, the VS interface system was found to have the lowest binding energy of −132.54 meV per atom and accordingly the shortest interlayer distance of 3.08 Å. For the cases of VI and VCs, the Eb values were found to slightly increase to −125.81 and −117.95 meV per atom, while the dint values decrease to 3.18 and 3.22 Å, respectively.


image file: d5ra05294g-f2.tif
Fig. 2 The interlayer binding energy per atom Eb for the Cs3Bi2I9/MoS2 heterojunction as a function of interlayer binding distance dint. The inset represents the magnified view around the equilibrium distance deq.
Table 1 The interface formation energy Ef, interlayer binding energy Eb and interlayer equilibrium distances dint in the Cs3Bi2I9/MoS2 heterostructures without and with vacancy defects
System Ef (J m−2) Eb (meV per atom) dint (Å)
Perfect −0.70 −127.82 3.13
VCs   −117.95 3.22
VI   −125.81 3.18
VS   −132.54 3.08


3.2 Formation of interfacial vacancy defect

Defects play a decisive role in the photovoltaic and photocatalytic applications. To check the formation possibility of the interfacial vacancy defects under the synthesis conditions, we estimated their formation energies from the thermodynamic point of view. According to eqn (1), the defect formation energy varies with the chemical potentials of the elements, which reflect the synthesis conditions in the experiments. Therefore, we first determined the chemical potential ranges of all the species included in the interface system for ensuring the stable existence of bulk Cs3Bi2I9 and MoS2 by using the thermodynamic constraints given by eqn (2)–(6) and the formation energy values of the available compounds (see Table S2 and Fig. S4, SI). Note that none of the ternary compounds are available in the Cs-Bi-I system, except for Cs3Bi2I9.

Fig. 3(a) presents the chemical potential diagram obtained in this work concerning the chemical potentials of Cs (ΔμCs) and Bi (ΔμBi), which are referenced to those of the corresponding elemental bulks. The thermodynamically stable region for the Cs3Bi2I9 compound is marked as the grey-colored region shown as the pentagon ABCDE (see Table S3 for the chemical potential values at the five vertexes of the pentagon, ESI). Among the vertex points of the pentagon, point B has the maximum value of ΔμCs and the minimum value of ΔμIμCs = −2.70 eV, ΔμBi = 0.02 eV, ΔμI = −0.99 eV), while point E has the maximum value of ΔμI and the minimum value of ΔμCsμCs = −3.74 eV, ΔμBi = −2.13 eV, ΔμI = −0.17 eV). Therefore, point B is said to represent the Cs-rich/I-poor condition, while point E represents the I-rich/Cs-poor condition. The possible chemical potential ranges for Mo and S species were also evaluated (see Table S3, SI). Note that outside of the pentagon region, the undesirable binary compounds such as CsI, BiI3, CsI3 and CsI4 are more stable than Cs3Bi2I9. For instance, BiI3 will form in the left-hand side of the A–E line, while the CsI compound is stable in the right-hand side of the B–C line. It is worth noting that the stable region has a rather wide shape in contrast to the very narrow shapes for the Pb-based HPs, such as MAPbI3 and CsPbI3.16,51,73 This indicates that the Bi-based HP Cs3Bi2I9 is more stable than the Pb-based HPs, for which instability is the main obstacle to their photovoltaic applications.


image file: d5ra05294g-f3.tif
Fig. 3 (a) Chemical potential ranges of Cs and Bi for the stable Cs3Bi2I9 compound, marked as grey-colored pentagon ABCDE. Outside the pentagon, the other binary compounds such as CsI, BiI3, CsI3, CsI4, Bi9I2, CsBi, Cs3Bi and CsBi2 will form instead of the Cs3Bi2I9 perovskite. Insets depict the magnified views around the vertexes of the pentagon. (b) Calculated formation energies of VCs and VI defects at theCs3Bi2I9 part of the interface as a function of the Fermi energy (EF) at points A, B, C, D and E. (c) The formation energy of vacancy defect VS at the MoS2 part of the interface under S-poor and S-rich conditions.

We then calculated the band structures of the Cs3Bi2I9 and MoS2 surfaces having the same planar size as the interface to determine the Fermi levels EF, bandgaps image file: d5ra05294g-t14.tif and VBM levels (EVBM) (see Fig. S5, SI). To determine the EVBM levels for the interface, we adopted the line-up-averaged electrostatic potential method74,75 for the Cs3Bi2I9/MoS2 interface, and Cs3Bi2I9 and MoS2 surfaces (see Fig. S6, SI). From the obtained band structures and the work functions (calculated by ϕ = EvEF, where Ev is the vacuum level), the VBM levels were corrected to be image file: d5ra05294g-t15.tif and image file: d5ra05294g-t16.tif for the Cs3Bi2I9 and MoS2 sides, respectively.

Fig. 3(b) shows the calculated formation energies of the vacancy defects VCs and VI as functions of EF at the points A, B, C, D and E. The formation energies of all the vacancy defects were found to be positive, indicating that their formation is endothermic. In most cases (points A, B, C, and D), V0I has the lowest formation energy, suggesting that V0I is easier to form than other vacancies and thus the dominant defect at the Cs3Bi2I9/MoS2 interface. In particular, this argument is more evident under the I-poor/Cs-rich condition (points A, B and C). Note that in the Pb-based HPs, the VI defect was also found to be the major defect with the lowest formation energy.54,55 When going to the I-rich/Cs-poor condition (points D and E), it is natural to observe that the formation energy of VCs decreases while that of VI increases, and finally VCs has a lower formation energy than VI at point E. On the other hand, VS was also found to have a positive formation energy, as shown in Fig. 3(c), implying that its formation at the interface is not spontaneous but endothermic in accordance with the case of the MoS2 bulk.76 Moreover, the formation energy of the V0S state was found to be lower than those of the VS1+ and VS2+ states in the whole range of EF. Interestingly, none of the thermodynamic transition levels between the differently charged defects, where the electrons could be donated or accepted, were observed, suggesting that these vacancy defects do not cause the undesirable charge recombination at the Cs3Bi2I9/MoS2 interface.

3.3 Interfacial charge redistribution and DOS analysis

We evaluated the charge density difference upon interface formation to gain intuitive insights into the binding strength and charge carrier transfer across the interface. Fig. 4(a) depicts the calculated planar-averaged charge density differences upon the formation of Cs3Bi2I9/MoS2 interfaces without and with the vacancy defects. At a glance, the charge redistribution occurs locally around the interface region, while being barely influenced by the defect formation. The charge accumulation was observed mostly in the interstitial region biased to the MoS2 side, whereas the charge depletion was found mostly around the CsI3 layer of the Cs3Bi2I9 side. This observation is further clarified by the isosurface plot of the spatial electron density difference, as can be seen in Fig. 4(b), where the charge depletion (blue colour) is distributed around the I atoms of Cs3Bi2I9, while charge accumulation (red colour) is seen around the S atoms of the MoS2 layer (see Fig. S7 for the defective systems, SI). The observed charge redistribution indicates that the significant amount of electrons are transferred from the MoS2 layer to the Cs3Bi2I9 side, leading to the strong binding between them at the interface. Moreover, the interface dipole originating from this charge redistribution is expected to help the transfer of photogenerated electrons and holes as will be discussed later.
image file: d5ra05294g-f4.tif
Fig. 4 (a) Planar-averaged charge density difference (Δ[small rho, Greek, macron]) integrated over the xy plane along the z-axis upon the formation of Cs3Bi2I9/MoS2 interfaces without and with the vacancy defects under study. (b) Isosurface plot of the spatial charge density difference upon the formation of the perfect interface projected on the (100) planes passing the I atoms (top panel) and Cs atoms (bottom panel) of the Cs3Bi2I9 surface, ranging from 0.001 to 0.001 |e| Å−3. Partial density of states (PDOS) and isosurface and isoline plots of local density of states (LDOS) for (c) the perfect interface, (d) VCs-, (e) VI- and (f) VS-containing interfaces.

For a quantitative insight into the charge redistribution, we integrated the planar-averaged electron density difference to obtain the total transferred charge as Δq = ∫Δ[small rho, Greek, macron](z)dz, which was further divided into the contributions from the Cs3Bi2I9qCs3Bi2I9) and MoS2qMoS2) sides. Table 2 lists the calculated transferred charge contributions from two constituent layers in the perfect and defective interfaces. For all the cases, the Cs3Bi2I9 part was found to have negative values whereas the MoS2 side had positive ones. This indicates that charge is depleted at the terminal surface of Cs3Bi2I9, while it is accumulated at the MoS2 monolayer, resulting in the creation of an interfacial dipole moment oriented from the Cs3Bi2I9 side to the MoS2 side, which prevents the charge carriers from transferring across the interface. Among the different systems without and with the vacancy defects, the largest charge transfer was observed in the VI-containing system, whereas the smallest charge transfer occurred in the VCs-containing interface.

Table 2 Transferred charge upon the formation of the Cs3Bi2I9/MoS2 heterostructures under study, which is divided into contributions from each of the subsystems (ΔqCs3Bi2I9, ΔqMoS2), and the interfacial dipole moment Δμ
System ΔqCs3Bi2I9 (|e|) ΔqMoS2 (|e|) Δμ (Debye)
Perfect −0.306 0.308 4.23
VCs −0.202 0.204 2.41
VI −0.318 0.321 4.79
VS −0.250 0.254 4.29


Charge transfer often induces interface polarization, which plays an important role in the properties of a heterostructure and is quantified by the interfacial dipole moment. As mentioned above, the interfacial dipole moment, orienting from the Cs3Bi2I9 side to the MoS2 side, helps the photogenerated electrons and holes move across the interface in opposite directions, thereby being favourable for preventing the recombination of photogenerated charge carriers and enhancing the photocatalytic performance in the Cs3Bi2I9/MoS2 heterostructure. The interface polarization induces a step in the electrostatic potential across the interface, which results in changes in the band-edge alignment. The interfacial dipole moment also originates the downward and upward band bending with the formation of a charge accumulation layer at the interface. The interfacial dipole moment, induced at the interface from the charge redistribution, was evaluated by Δμ = −∫zΔ[small rho, Greek, macron]dz. The calculated Δμ values are listed in Table 2. The VI-containing interface was also found to have the largest Δμ value of 4.79 D among the interface systems under study. Meanwhile, the VCs defect causes a reduction in Δμ to 2.41 D. Such tendencies are consistent with the previous calculations for CsPbI3/2H-MoS2 (ref. 16) and MAPbI3/2H-MoS2 interfaces,77 where VI defect systems have obviously larger Δμ values and VCs systems have smaller Δμ values. When compared with those of the Pb-based HP/MoS2 heterostructures, however, Δμ values are relatively larger, suggesting further improvement of photocatalytic performance in the Cs3Bi2I9/MoS2 interface.

In order to understand the generation and transport processes of charge carriers induced by photon incidence, we calculated the electronic density of states (DOS) of the interface systems. Fig. 4(c)–(f) display the calculated atom-projected partial density of states (PDOS) of the Cs3Bi2I9 and MoS2 sides and the local density of states (LDOS) along the z-axis for all the interface systems. Our calculations revealed that the lower conduction bands (CBs) responsible for the transport of photo-induced electrons are contributed from Bi and I atoms on the Cs3Bi2I9 side and both Mo and S atoms on the MoS2 side. From the LDOS plots, the CB edge states of MoS2 were found to be lower in energy than those of Cs3Bi2I9, being beneficial to the transfer of electrons photo-generated in the Cs3Bi2I9 side to the MoS2 side. It was found that the VCs defect causes barely any change in the electronic properties of the perfect interface system, whereas the VI and VS defects induce shallow trap states near the CBM level. In particular, the VS-containing interface exhibits the trap state relatively far below the CBM level. Since trap states inside the bandgap would behave as recombination centers, they can be detrimental to the photocatalytic performance. However, it can be said that the VS defect is difficult to be formed due to its relatively high formation energy.

3.4 Photocatalytic performance

In understanding the photocatalytic performance for HER, the band edge alignment is an important criterion. For a high HER performance, it is required that the CBM level should be higher than the hydrogen reduction potential of H+/H2, while the VBM level should be lower than the water oxidation potential of O2/H2O. Selecting the vacuum level as the reference, the hydrogen reduction potential and the water oxidation potential are calculated at different pH values as follows:56,73
 
image file: d5ra05294g-t17.tif(15)
 
image file: d5ra05294g-t18.tif(16)

Once the Cs3Bi2I9 surface is combined with the MoS2 monolayer to build the heterostructure, some electrons are transferred to ensure the same Fermi level at the interface.64 By using the calculated work function ϕ, we determined the Fermi levels of the Cs3Bi2I9/MoS2 interfaces, and then obtained the EVBM and ECBM values for the Cs3Bi2I9 and MoS2 parts, which are referenced to the vacuum level Ev (see Fig. S8 for the band structures, SI).

For the perfect interface system, the VBM and CBM levels were calculated to be −4.66 and −3.10 eV for the Cs3Bi2I9 side and −5.18 and −3.40 eV for the MoS2 side, respectively. Therefore, the interface was characterized as a type-II heterojunction, being significantly beneficial to the separation of electrons and holes in opposite directions, i.e., preventing the recombination of charge carriers. The interfacial dipole moment induced from the charge redistribution originates the downward and upward band bending with the formation of a charge accumulation layer at the interface.75 Fig. 5(a) shows a scheme that illustrates the generation and transfer processes of charge carriers at the heterojunction. When photons are absorbed in the Cs3Bi2I9/MoS2 heterojunction, electrons are excited from the VBM to the CBM on each side of the interface, due to their proper bandgaps for light absorption. Then, the electrons placed on the CBM of Cs3Bi2I9 are attracted toward the MoS2 side by the interface dipole, while the holes left in the VBM of MoS2 are drawn to the Cs3Bi2I9 side.


image file: d5ra05294g-f5.tif
Fig. 5 (a) Schematic diagram of band alignment in the Cs3Bi2I9/MoS2 interface with a type-II heterojunction. Here, Ev and EF are the vacuum and Fermi energy levels, and EVB and ECB are the VBM and CBM levels of the Cs3Bi2I9 and MoS2 sides. (b) Band-edge alignments of Cs3Bi2I9/MoS2 interface systems, referenced to the vacuum level. Turquoise and magenta bars indicate VBs and CBs of the Cs3Bi2I9 and MoS2 parts, respectively. The Fermi level is depicted as a horizontal black line for each interface system. Blue and red horizontal long lines represent the hydrogen reduction and water oxidation potentials, respectively, at different pH values.

Fig. 5(b) illustrates the calculated band-edge alignments of the interface systems under study (see Table S4 for the corresponding numerical values, SI). In all cases, the CBM is higher than the H+/H2 reduction potential, indicating that the Cs3Bi2I9/MoS2 heterostructure has the potential to reduce H+ to H2. In the case of the VI-containing interface, however, it was revealed that the band-edge alignment is not appropriate for water splitting, since the VBM is even higher than the H+/H2 potential. For the perfect interface and VS-defective systems, the CBMs are lower than the O2/H2O potential only at pH values higher than 8, i.e., in the basic environment. Therefore, they cannot produce O2 by oxidizing H2O under acidic conditions. With the proper band-edge alignment, the VCs-containing interface is likely to be the most promising photocatalyst for water redox reactions, especially in an acidic environment.

For further insights into the photocatalytic performance of the interface, we evaluated the Gibbs free energy for hydrogen adsorption (ΔGH*).56 From the viewpoint of HER activity, the absolute value of ΔGH* should be as small as possible. We calculated ΔGH* for the Cs3BI2I9/MoS2 heterojunctions, together with the isolated Cs3BI2I9 surface and the MoS2 monolayer. Considering the previous result,16 we proposed that the H atom could be adsorbed on the S atom in the top layer of the MoS2 monolayer (see Fig. S9, SI). For the case of the Cs3Bi2I9 surface, the adsorption site of the H atom was identified using energetic calculations. As shown in Fig. 6, the Cs3Bi2I9 surface was found to show the highest ΔGH* value of 2.20 eV, while the perfect heterostructure and MoS2 monolayer have nearly the same value of 2.01 eV, which is consistent with the ΔGH* value obtained for the graphene/MoS2 heterostructure.78 This indicates that the formation of the interface between Cs3Bi2I9 and MoS2 without any defects hardly affects on the Gibbs free energy for hydrogen adsorption on the MoS2 layer. The vacancy defect formation at the interface was found to have a positive effect on HER activity due to the lower ΔGH*, except for the VI defect which was found to slightly increase the ΔGH* value. For VS- and VCs-containing systems, ΔGH* was reduced to 1.48 and 0.50 eV, respectively, implying some improvement of HER performance.


image file: d5ra05294g-f6.tif
Fig. 6 Gibbs free energy difference for hydrogen adsorption (ΔGH*) on the Cs3Bi2I9/MoS2 interface systems without and with the vacancy defects such as VCs, VI and VS, together with those on the pristine Cs3Bi2I9 surface and MoS2 monolayer.

For the final step, we calculated the photoabsorption spectra of the Cs3Bi2I9/MoS2 heterostructures without and with the interfacial vacancy defect, bulk Cs3Bi2I9 and the MoS2 monolayer, as shown in Fig. 7. From these spectra, we could estimate the photoabsorption edge and intensity, which are crucial factors to achieve high photocatalytic activity (see Fig. S10 for dielectric functions and other optical properties, SI). For bulk Cs3Bi2I9 and the MoS2 monolayer, the absorption edges were found to be 1.54 and 1.84 eV, respectively. Meanwhile, those of the heterostructures were found to be significantly shifted to the lower-energy region varying from 1.18 to 1.48 eV. Moreover, the Cs3Bi2I9/MoS2 heterostructures exhibited higher absorption intensities than their constituents in the visible-light and near-ultraviolet regions, representing their superior photoabsorption capability. The perfect interface has greater absorption intensity than the defective interfaces in the visible-light region, while in the ultraviolet region the defective heterostructures have higher absorption coefficients than the perfect interface. It is worth noting that among the different vacancy defects, the VCs defect induces the greatest intensity in the visible-light region. This is mainly due to the reduced bandgaps and the built-in electric field originating from the charge redistribution at the interface.79,80


image file: d5ra05294g-f7.tif
Fig. 7 Photoabsorption spectra of bulk Cs3Bi2I9, the MoS2 monolayer and the Cs3Bi2I9/MoS2 interface systems without and with a vacancy defect. The vertical dashed lines indicate the visible-light region.

4 Conclusions

We have systematically investigated the structural, electronic and optical properties of Cs3Bi2I9/MoS2 heterostructures using first-principles calculations to give a comprehensive theoretical understanding on the enhancement of photocatalytic water splitting activity for hydrogen evolution. Using the (1 × 1) unit cell of the Cs3Bi2I9 (001) surface and a (7 × 7) cell of the 2H-MoS2 monolayer, we built a slab model of the Cs3Bi2I9/MoS2 interface with a very small lattice mismatch of 0.1% without and with a vacancy defect, such as VCs, VI and VS. The formation and binding energies of the interface systems were calculated, confirming that these heterostructures are energetically stable. We computed the defect formation energies in the defective Cs3Bi2I9/MoS2 interfaces under the different growth conditions through chemical potentials of the species, finding that the VI defect could be generated with the lowest formation energy under the I-poor/Cs-rich condition, while the VCs defect would be dominant under the Cs-poor/I-rich condition. Through the analysis of charge redistribution upon interface formation, we demonstrated that some electrons are transferred from the MoS2 side to the Cs3Bi2I9 side, inducing the interfacial dipole moment beneficial to preventing the recombination of photogenerated charge carriers. Our calculations of DOS and photoabsorption spectra revealed that the perfect interface system and the VCs-containing interface could be favourable for photocatalytic activity compared with other defective systems, since they do not induce shallow trap states inside the bandgap range and exhibit higher absorption intensities than others. Finally, we found that the VCs-containing heterostructure is particularly promising for improving the photocatalytic activity for hydrogen generation due to the significant reduction of the Gibbs free energy for hydrogen adsorption to 0.5 eV and the most suitable band alignment for photo-induced overall water splitting, whose band edge positions straddle the water redox potentials.

Author contributions

Yun-Sim Kim and Chol-Jun Yu developed the original project. Kyong-Mi Kim and Yun-Sim Kim performed the calculations and drafted the first manuscript. Dok-Ho Hyon and Chol-Hyok Ri contributed to useful discussions. Chol-Jun Yu supervised the work. All authors reviewed the manuscript.

Conflicts of interest

There are no conflicts to declare.

Data availability

The data are available in the supplementary information (SI) and further from the authors upon reasonable request. Supplementary information: Tables for defect position search, formation energies of the compounds, chemical potential values and band alignment, and figures for the band structure of the MoS2 monolayer, two different sliding configurations, vacancy sites at the interfaces, the Cs-Bi-I ternary system, band structures and electrostatic potential of the interface systems, spatial charge density difference of the heterostructures with vacancy defects, adsorption geometries for hydrogen adsorption on different systems, real and imaginary parts of the dielectric function and other optical properties. See DOI: https://doi.org/10.1039/d5ra05294g.

Acknowledgements

This work is supported as part of the fundamental research project “Design of New Energy Materials” (no. 2021-12) funded by the State Commission of Science and Technology, DPR Korea. Computations in this work have been done on the HP Blade System C7000 (HP BL460c) that is owned and managed by Faculty of Materials Science, Kim Il Sung University.

References

  1. N. S. Lewis, Science, 2016, 351, 353 CrossRef CAS PubMed.
  2. P. V. Kamat and K. Sivula, ACS Energy Lett., 2022, 7, 3149–3150 CrossRef CAS.
  3. M. Sohail, S. Rauf, M. Irfan, A. Hayat, M. M. Alghamdi, A. A. El-Zahhar, D. Ghernaout, Y. Al-Hadeethihij and W. Lv, Nanoscale Adv., 2024, 6, 1286–1330 RSC.
  4. Y. Wang, H. Suzuki, J. Xie, O. Tomita, D. J. Martin, M. Higashi, D. Kong, R. Abe and J. Tang, Chem. Rev., 2018, 118, 5201–5241 CrossRef CAS PubMed.
  5. Q. Hao, Y. Zhu, Y. Li, Z. Li, H. Yuan and S. Ouyang, Ind. Chem. Mater., 2025, 3, 203 RSC.
  6. D. Bhanderi, P. Lakhani and C. K. Modi, RSC Sustain., 2024, 2, 265–287 RSC.
  7. V. Jeyalakshmi, S. Wu, S. Qin, X. Zhou, B. B. Sarma, D. E. Doronkin, J. Kolařík, M. Šoóš and P. Schmuki, Chem. Sci., 2025, 16, 4788 RSC.
  8. A. M. Huerta-Flores, F. Ruiz-Zepeda, C. Eyovge, J. P. Winczewski, M. Vandichel, M. Gaberšček, N. D. Boscher, H. J. Gardeniers, L. M. Torres-Martínez and A. Susarrey-Arce, ACS Appl. Mater. Interfaces, 2022, 14, 31767–31781 CrossRef CAS PubMed.
  9. S. D. Tilley, ACS Energy Lett., 2023, 8, 2338–2344 CrossRef CAS.
  10. W. Hou, S. A. J. Hillman, S. Gonzalez-Carrero, S. Yao, H. Zhuc and J. R. Durrant, Sustain. Energy Fuels, 2025, 9, 576 RSC.
  11. R. Peng, X. Ma, Z. D. Hood, A. Boulesbaa, A. A. Puretzky, J. Tong and Z. Wu, J. Mater. Chem. A, 2023, 11, 16714–16723 RSC.
  12. S. M. Younan, Z. Li, X. Yan, D. He, W. Hu, N. Demetrashvili, G. Trulson, A. Washington, X. Xiao, X. Pan, J. Huang and J. Gu, ACS Nano, 2023, 17, 1414–1426 CrossRef CAS PubMed.
  13. Q. Cai, I. M. McWhinnie, N. W. Dow, A. Y. Chan and D. W. C. MacMillan, J. Am. Chem. Soc., 2024, 146, 12300–12309 CrossRef PubMed.
  14. I. M. A. Mekhemer, M. M. Elsenety, A. M. Elewa, K. D. G. Huynh, M. M. Samy, M. G. Mohamed, D. M. Dorrah, D. C. K. Hoang, A. F. Musa, S.-W. Kuo and H.-H. Chou, J. Mater. Chem. A, 2024, 12, 10790–10798 RSC.
  15. C. Yuan, H. Yin, H. Lv, Y. Zhang, J. Li, D. Xiao, X. Yang, Y. Zhang and P. Zhang, J. Am. Chem. Soc. Au, 2023, 3, 3127–3140 CAS.
  16. C.-H. Ri, H.-U. Han, Y.-S. Kim, U.-G. Jong, Y.-H. Kye and C.-J. Yu, J. Phys. Chem. Lett., 2022, 13, 4007–4014 CrossRef CAS PubMed.
  17. F. Xu and B. Weng, J. Mater. Chem. A, 2023, 11, 4473–4486 RSC.
  18. P. Wu, J. Lu, F. Xi, X. Li, W. Ma, F. Kang, S. Li, Z. Tong and Q. Zhang, Chem. Sci., 2025, 16, 4127 RSC.
  19. A. Kojima, K. Teshima, Y. Shirai and T. Miyasaka, J. Am. Chem. Soc., 2009, 131, 6050–6051 CrossRef CAS PubMed.
  20. C. Liu, Y. Yang, H. Chen, J. Xu, A. Liu, A. S. R. Bati, H. Zhu, L. Grater, S. S. Hadke, C. Huang, V. K. Sangwan, T. Cai, D. Shin, L. X. Chen, M. C. Hersam, C. A. Mirkin, B. Chen, M. G. Kanatzidis and E. H. Sargent, Science, 2023, 382, 810–815 CrossRef CAS PubMed.
  21. M. Saliba, Science, 2018, 359, 388–389 CrossRef CAS PubMed.
  22. Z. Wei, Q. Zhou, X. Niu, S. Liu, Z. Dong, H. Liang, J. Chen, Z. Shi, X. Wang, Z. Jia, X. Guo, R. Guo, X. Meng, Y.-D. Wang, N. Li, Z. Xu, Z. Li, A. G. Aberle, X. Yin and Y. Hou, Energy Environ. Sci., 2025, 18, 1847–1855 RSC.
  23. A. Mohandes and M. Moradi, Mater. Adv., 2025, 6, 1520–1539 RSC.
  24. Y. Yang, M. Yang, D. T. Moore, Y. Yan, E. M. Miller, K. Zhu and M. C. Beard, Nat. Energy, 2017, 2, 16207 CrossRef CAS.
  25. T. M. Brenner, D. A. Egger, L. Kronik, G. Hodes and D. Cahen, Nat. Rev. Mater, 2016, 1, 15007 CrossRef CAS.
  26. Y. Yang, D. P. Ostrowski, R. M. France, K. Zhu, J. van de Lagemaat, J. M. Luther and M. C. Beard, Nat. Photonics, 2016, 10, 53–59 CrossRef CAS.
  27. L. M. Herz, Annu. Rev. Phys. Chem., 2016, 67, 65–89 CrossRef CAS PubMed.
  28. Q. Dong, Y. Fang, Y. Shao, P. Mulligan, J. Qiu, L. Cao and J. Huang, Science, 2015, 347, 967–970 CrossRef CAS PubMed.
  29. C. Wehrenfennig, G. E. Eperon, M. B. Johnston, H. J. Snaith and L. M. Herz, Adv. Mater., 2014, 26, 1584–1589 CrossRef CAS PubMed.
  30. H. Huang, B. Pradhan, J. Hofkens, M. B. J. Roeffaers and J. A. Steele, ACS Energy Lett., 2020, 5, 1107–1123 CrossRef CAS.
  31. X. Zhao, S. Chen, H. Yin, S. Jiang, K. Zhao, J. Kang, P. F. Liu, L. Jiang, Z. Zhu, D. Cui, P. Liu, X. Han, H. G. Yang and H. Zhao, Matter, 2020, 3, 935–949 CrossRef.
  32. T. Pal, S. Bhowmik, A. S. Tanwar, A. Suhail, N. R. Peela, C. V. Sastri and P. K. Iyer, Energy Adv., 2024, 3, 2965 RSC.
  33. H. Singh, N. Minhas, G. Mustafa, G. Singh, A. Kaura and J. K. Goswamy, RSC Adv., 2025, 15, 4779 RSC.
  34. Z. Guan, Y. Wu, P. Wang, Q. Zhang, Z. Wang, Z. Zheng, Y. Liu, Y. Dai, M.-H. Whangbo and B. Huang, Appl. Catal., B, 2019, 245, 522–527 CrossRef CAS.
  35. C.-H. Kuan, Y.-A. Ko and E. W.-G. Diau, ACS Energy Lett., 2023, 8, 2423–2425 CrossRef CAS.
  36. U.-G. Jong, C.-J. Yu, Y.-H. Kye, S.-N. Hong and H.-G. Kim, Phys. Rev. Mater., 2020, 4, 075403 CrossRef CAS.
  37. A. Das, S. D. Peu, M. A. M. Akanda, M. M. Salah, M. S. Hossain and B. K. Das, Energies, 2023, 16, 2328 CrossRef CAS.
  38. M. Miodyńska, T. Klimczuk, W. Lisowski and A. Zaleska-Medynska, Catal. Commun., 2023, 177, 106656 CrossRef.
  39. C. Tedesco and L. Malavasi, Molecules, 2023, 28, 339 CrossRef CAS PubMed.
  40. M. Medina-Llamas, A. Speltini, A. Profumo, F. Panzarea, A. Milell, F. Fracassi, A. Listorti and L. Malavasi, Nanomaterials, 2023, 13, 263 CrossRef CAS PubMed.
  41. A. Pancielejko, M. Miodyńska, H. Głowienke, A. Gołabiewska, E. Gontarek-Castro, T. Klimczuk, M. Krawczyk, G. Trykowski and A. Zaleska-Medynska, Catal. Today, 2024, 432, 114626 CrossRef CAS.
  42. T. Tang, X. Dou, H. Zhang, H. Wang, M. Li, G. Hu, J. Wen and L. Jiang, Molecules, 2024, 29, 5096 CrossRef CAS PubMed.
  43. Y. Fan, J. Hu, T. Li, S. Xu, S. Chen and H. Yin, Chem. Commun., 2024, 60, 1004 RSC.
  44. P. Giannozzi, S. Baroni, N. Bonini, M. Calandra, R. Car, C. Cavazzoni, D. Ceresoli, L. C. Guido, M. Cococcioni and I. Dabo, et al., J. Phys.: Condens. Matter, 2009, 21, 395502 CrossRef PubMed.
  45. K. F. Garrity, J. W. Bennet, K. M. Rabe and D. Vanderbilt, Comput. Mater. Sci., 2014, 81, 446–452 CrossRef CAS.
  46. J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865 CrossRef CAS PubMed.
  47. J. Klimeš, D. R. Bowler and A. Michaelides, Phys. Rev. B: Condens. Matter Mater. Phys., 2011, 83, 195131 CrossRef.
  48. J. M. Soler, E. Artacho, J. D. Gale, A. García, J. Junquera, P. Ordejón and D. Sánchez-Portal, J. Phys.: Condens. Matter, 2002, 14, 2745 CrossRef CAS.
  49. V. Roiati, E. Mosconi, A. Listorti, S. Colella, G. Gigli and F. de Angelis, Nano Lett., 2014, 14, 2168–2174 CrossRef CAS PubMed.
  50. E. Mosconi, E. Ronca and F. D. Angelis, J. Phys. Chem. Lett., 2014, 5, 2619–2625 CrossRef CAS PubMed.
  51. Y.-S. Kim, C.-H. Ri, U.-H. Ko, Y.-H. Kye, U.-G. Jong and C.-J. Yu, ACS Appl. Mater. Interfaces, 2021, 13, 14679–14687 CrossRef CAS PubMed.
  52. C. Freysoldt, B. Grabowski, T. Hickel, J. Neugebauer, G. Kresse, A. Janotti and C. G. Van de Walle, Rev. Mod. Phys., 2014, 86, 253–305 CrossRef.
  53. Y.-H. Kye, C.-J. Yu, U.-G. Jong, Y. Chen and A. Walsh, J. Phys. Chem. Lett., 2018, 9, 2196–2201 CrossRef CAS PubMed.
  54. W.-J. Yin, T. Shi and Y. Yan, Appl. Phys. Lett., 2014, 104, 063903 CrossRef.
  55. Y. Li, C. Zhang, X. Zhang, D. Huang, Q. Shen, Y. Cheng and W. Huang, Appl. Phys. Lett., 2017, 111, 162106 CrossRef.
  56. F. Opoku, O. Akoto, E. E. Kwaansa-Ansah, N. K. Asare-Donkor and A. A. Adimado, Mater. Adv., 2023, 4, 1949 RSC.
  57. M. Gajdoš, K. Hummer, G. Kresse, J. Furthmüller and F. Bechstedt, Phys. Rev. B: Condens. Matter Mater. Phys., 2006, 73, 045112 CrossRef.
  58. C.-J. Yu, U.-H. Ko, S.-G. Hwang, Y.-S. Kim, U.-G. Jong, Y.-H. Kye and C.-H. Ri, Phys. Rev. Mater., 2020, 4, 045402 CrossRef CAS.
  59. Y. Ying, X. Luo and H. Huang, J. Phys. Chem. C, 2018, 122, 17718–17725 CrossRef CAS.
  60. S. Saidi, S. Zriouel, L. B. Drissi and M. Maaroufi, Phys. E, 2018, 103, 171–179 CrossRef CAS.
  61. Z. Li, B. Jia, S. Fang, Q. Li, F. Tian, H. Li, R. Liu, Y. Liu, L. Zhang, S. F. Liu and B. Liu, Adv. Sci., 2023, 10, 2205837 CrossRef CAS PubMed.
  62. F. Bai, Y. Hu, Y. Hu, T. Qiu, X. Miao and S. Zhang, Sol. Energy Mater. Sol. Cells, 2018, 184, 15–21 CrossRef CAS.
  63. D. S. Koda, F. Bechstedt, M. Marques and L. K. Teles, J. Phys. Chem. C, 2016, 120, 10895–10908 CrossRef CAS.
  64. J. Zhou, D. Li, W. Zhao, B. Jing, Z. Ao and T. An, ACS Appl. Mater. Interfaces, 2021, 13, 23843–23852 CrossRef CAS PubMed.
  65. F. Opoku, A. Aniagyei, O. Akoto, E. E. Kwaansa-Ansah, N. K. Asare-Donkor and A. A. Adimado, Mater. Adv., 2022, 3, 4629 RSC.
  66. Y. Shi, O. V. Prezhdo, J. Zhao and W. A. Saidi, ACS Energy Lett., 2020, 5, 1346–1354 CrossRef CAS.
  67. C.-H. Ri, S.-H. Pak, S.-I. O., C.-S. Jang, Y.-S. Kim, J.-S. Kim and C.-J. Yu, Phys. Chem. Chem. Phys., 2025, 27, 5555 RSC.
  68. C. Ataca, H. Shahin, E. Aktruk and S. Ciraci, J. Phys. Chem. C, 2011, 115, 3934–3941 CrossRef CAS.
  69. Z. Li, D. Zhoub and J. Xiangke, Sens. Actuators, A, 2021, 322, 112637 CrossRef CAS.
  70. A. Lehner, D. Fabini, H. Evans, C. Hébert, S. Smock, J. Hu, H. Wang, J. Zwanziger, M. Chabinyc and R. Seshadri, Chem. Mater., 2015, 27, 7137–7148 CrossRef CAS.
  71. W. Geng, L. Zhang, Y.-N. Zhang, W.-M. Lau and L.-M. Liu, J. Phys. Chem. C, 2014, 118, 19565–19571 CrossRef CAS.
  72. U.-G. Jong, C.-J. Yu, Y.-S. Kim, Y.-H. Kye and C.-H. Kim, Phys. Rev. B, 2018, 98, 125116 CrossRef CAS.
  73. C.-H. Ri, S.-H. Pak, S.-I. O, C.-S. Jang, Y.-S. Kim, J.-S. Kim and C.-J. Yu, Phys. Chem. Chem. Phys., 2025, 27, 5555–5565 RSC.
  74. E. Cerrato, C. Gionco, M. C. Paganini, E. Giamello, E. Albanese and G. Pacchioni, ACS Appl. Energy Mater., 2018, 1, 4247–4260 CrossRef CAS.
  75. C.-J. Yu, Y.-H. Kye, U.-G. Jong, K.-C. Ri, S.-H. Choe, J.-S. Kim, S.-G. Ko, G.-I. Ryu and B. Kim, ACS Appl. Mater. Interfaces, 2020, 12, 1858–1866 CrossRef CAS PubMed.
  76. D. Cao, H. B. Shu, T. Q. Wu, Z. T. Jiang, Z. W. Jiao, M. Q. Cai and W. Y. Hu, Appl. Surf. Sci., 2016, 361, 199–205 CrossRef CAS.
  77. Y. Shi, O. V. Prezhdo, J. Zhao and W. A. Saidi, ACS Energy Lett., 2020, 5, 1346–1354 CrossRef CAS.
  78. R. K. Biroju, D. Das, R. Sharma, S. Pal, L. P. L. Mawlong, K. Bhorkar, P. K. Giri, A. K. Singh and T. N. Narayanan, ACS Energy Lett., 2017, 2, 1355–1361 CrossRef CAS.
  79. F. Ning, D. Wang, Y.-X. Feng, L.-M. Tang, Y. Zhang and K.-Q. Chen, J. Mater. Chem. C, 2017, 5, 9429–9438 RSC.
  80. C.-S. Liao, Q.-Q. Zhao, Y.-Q. Zhao, Z.-L. Yu, H. Zhou, P.-B. He, J.-L. Yang and M.-Q. Cai, J. Phys. Chem. Solids, 2019, 135, 109060 CrossRef CAS.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.