Open Access Article
Shizi Luoab,
Daxiong Liuc,
Xiang Deng*c,
Zhuoneng Bi
b,
Shuguang Caoj,
Tongjun Zhengab,
Liyao Xiongab,
Hao Lid,
Ning Li
d,
Lavrenty G. Gutsev
*ef,
Nikita A. Emelianov
e,
Victoria V. Ozerovae,
Nikita A. Slesarenkoe,
Alexander F. Shestakove,
Sergey M. Aldoshin
e,
Gennady L. Gutsev
g,
Pavel A. Troshin
*eh,
Bochuan Yang*i,
Zhibo Zhaoc and
Xueqing Xu
*ab
aSchool of Energy Science and Engineering, University of Science and Technology of China, Hefei, 230026, P. R. China
bKey Laboratory of Renewable Energy, Guangdong Provincial Key Laboratory of New and Renewable Energy Research and Development, Guangzhou Institute of Energy Conversion, Chinese Academy of Sciences, Guangzhou 510640, P. R. China. E-mail: xuxq@ms.giec.ac.cn
cGuangdong Mingyang Thin Film Technology Co., Ltd, Zhongshan, P. R. China
dInstitute of Polymer Optoelectronic Materials and Devices State Key Laboratory of Luminescent Materials and Devices, South China University of Technology, Guangzhou, 510640, P. R. China
eFederal Research Center for Problems of Chemical Physics and Medicinal Chemistry of RAS, Semenov Prospect 1, Chernogolovka, 142432, Russia
fInstitute for Micromanufacturing, Louisiana Tech University, Ruston, LA 71272, USA
gDepartment of Physics, Florida A&M 711 University, Tallahassee, Florida 32307, USA
hZhengzhou Research Institute of HIT, 26 Longyuan East 7th, Jinshui District, Zhengzhou, Henan Province 450000, P. R. China
iRisen Energy Co., Ltd, Ningbo, Zhejiang Province 315609, P. R. China
jSchool of Materials Science and Engineering, Henan Polytechnic University, Jiaozuo, Henan 454000, P. R. China
First published on 11th November 2025
An all-solution two-step method for preparing wide-bandgap perovskites has the advantages of low cost, good repeatability and scalability. However, achieving high-quality wide-bandgap perovskite films via an all-solution two-step method remains challenging due to uneven distribution of halogens and incomplete reactions between organic salts and inorganic salts. Herein, we introduced tris(pentafluorophenyl)borane (BCF) into an inorganic layer resulting in boron–halide bonding, which stabilized uniform halide distribution and regulated the porous structure of the lead halide films, facilitating the diffusion of organic salts. Additionally, the fluorine substituents formed hydrogen bonds with organic cations, making BCF a bifunctional additive that delayed the reaction between the organic ammonium salt and the inorganic precursor, which was conducive to the growth of large-grained perovskite crystals. During the perovskite crystallization process, BCF molecules migrated to grain boundaries and the film surface, achieving a highly positive regulation influence on the nanoscale morphology and structure of the perovskite absorber films, thus leading to a pinhole-free, stress-free and less defect perovskite films. Ultimately, the approach enabled single-junction 1.68 eV wide-bandgap perovskite solar cells with a champion efficiency of 23.49% (certified 22.73%) and a VOC of 1.291 V. Furthermore, the optimized perovskite films were pioneeringly and successfully integrated into monolithic perovskite/silicon tandem solar cells on textured silicon and achieved an efficiency of 31.12%, which is the highest value among tandem solar cells prepared by an all-solution two-step method, retaining >90% of initial performance after 500 hours of continuous operation.
Broader contextThe preparation of wide-bandgap perovskite solar cells via an all-solution two-step method offered potential pathways to regulate the crystallization rate of perovskite films. However, there are very few reports on the all-solution two-step sequential deposition of 1.68 eV wide-bandgap perovskite films. Therefore, it will become meaningful to investigate the crystallization kinetics of perovskite prepared by a two-step method. Herein, we regulated the crystallization kinetics of 1.68 eV wide-bandgap perovskite films by leveraging a simultaneous bonding of tris(pentafluorophenyl)borane (BCF) with halide anions and organic cations via the boron–halide and hydrogen bonding effects. These interactions stabilized halogen distribution, delayed and regulated the crystallization process of perovskite films, slowed down nonradiative recombination, and facilitated the rapid extraction of charges. As a result, we achieved a PCE of 23.49% and a VOC of 1.291 V. By integrating the optimized wide-bandgap PSCs on textured silicon, we pioneered an all-solution two-step sequential deposition method for preparing 1.68 eV perovskite/silicon tandem solar cells, which exhibited a PCE of 31.12%, which is the highest among those of two-step prepared tandem cells. This means that a huge breakthrough has been achieved in the preparation of wide-bandgap perovskite films using an all-solution two-step sequential deposition method. |
However, the high Br content (>20%) required to obtain a wide-bandgap perovskite generally leads to the well-known problem of increased ion migration, caused by a marked increase in the iodide vacancy concentration18,19 and phase segregation, resulting in a large number of defects that reduce PCE and compromise the long-term stability of the devices.20,21 These energy losses and instabilities are mainly due to the unbalanced nucleation rates between the I− and Br− components22 and the fact that iodide is preferentially oxidized in a mixed-halide perovskite.23 The perovskite films with uneven halide distribution show high defect density and low crystallinity.24 As such, large open-circuit voltage (VOC) losses are usually observed in wide-bandgap PSCs with an Eg of >1.65 eV.25,26 Generally speaking, improving the uniformity of halide distribution and the crystallinity of perovskite films by adjusting the crystallization kinetics of perovskite films is critical to further improve device performance.27 To address these issues, researchers have proposed various strategies to control the crystallization of mixed halide wide-bandgap perovskites, including component engineering, ligand engineering, Lewis's base additives, etc.20,28,29 Chen et al. developed a novel strategy using a cation alloy that not only tailors the lattice properties and crystallization but also effectively passivates the defects, enabling homogeneous halide distribution and substantially reducing defect density. These improvements have led to a remarkable power conversion efficiency (PCE) of 19.50% with a record open-circuit voltage of 1.35 V for 1.79 eV perovskite solar cells.30 Jiang et al. combined the rapid bromide crystallization with a gentle gas-quench method to prepare highly textured columnar 1.75 eV wide-bandgap perovskite films with reduced defect density, resulting in a PCE over 20% and an VOC of 1.33 V.27 Yang et al. reported a series of pyridine derivatives containing amino and carboxylic groups that were applied to modify the perovskite surface, which can interact with undercoordinated Pb2+ through coordination bonds and suppress halide ion migration via hydrogen bonding.24
Most strategies for preparing wide-bandgap perovskite films rely on the one-step spin-coating method, which often leads to non-uniform Br− distribution during the crystallization process.31 In contrast, there are limited reports on the fabrication of wide-bandgap PSCs using an all-solution two-step method. This method involves first depositing a lead halide film and then introducing an organic ammonium salt solution to form a perovskite film, offering potential pathways to control over the crystallization rate of wide-bandgap perovskites. For instance, Zhang et al. utilized a two-step method by incorporating formamidinium iodide (FAI) and rubidium acetate (RbAc) into a PbI2/PbBr2 solution (with a molar ratio of 0.83/0.17) to modify the crystallization kinetics of 1.63 eV perovskite films. This approach promoted the oriented growth of the perovskite layer, ultimately allowing the perovskite/silicon TSCs to exhibit a PCE of 27.64%.32
Herein, we presented a novel all-solution two-step method to regulate the crystallization kinetics of 1.68 eV wide-bandgap perovskite films by leveraging a simultaneous interaction of tris(pentafluorophenyl)borane (BCF) with halide anions and organic cations. Fourier-transform infrared (FTIR) spectroscopy, nuclear magnetic resonance (NMR) analyses and density functional theory (DFT) calculations revealed that the BCF molecules, with their electron-deficient boron center and multiple fluorine substituents, interacted with perovskite precursor components through boron–halide and hydrogen bonding effects. These interactions might regulate the reaction kinetics between the organic ammonium salt solution and the inorganic layer. In situ absorption spectroscopy studies demonstrated that the boron–halide interactions effectively stabilized uniform halide distribution within the lead halide film, while the chemical reaction and crystallization rates between the organic ammonium salt solution and the inorganic layer were delayed via boron–halide and hydrogen bonding interactions. This resulted in the growth of pinhole-free, stress-free and less defect perovskite films. The results of time-resolved photoluminescence and transient photocurrent measurement indicate that after the introduction of BCF, the electron–hole recombination time extended, slowing down nonradiative recombination and facilitating the rapid extraction of charges. As a result, the optimized single-junction PSCs exhibited a PCE of 23.49% and a VOC of 1.291 V, with a remarkably low VOC loss of only 0.39 V. The devices also exhibited exceptional long-term stability, retaining 90% of their initial efficiency after 750 hours of maximum power point (MPP) tracking. By integrating the optimized wide-bandgap PSCs with crystalline silicon (c-Si), we pioneered an all-solution two-step sequential deposition method for preparing 1.68 eV perovskite/silicon tandem solar cells on textured silicon, exhibiting a PCE of 31.12%, which is the highest among those of two-step prepared tandem cells, retaining >90% of initial performance after 500 hours of continuous operation.
:
1 molar ratio) precursor supplemented with 0.06 M CsI and 0.015 M RbCl. Tris(pentafluorophenyl)borane (BCF) was introduced into the inorganic layer as the strategic modifier, since it features highly electronegative fluorine substituents enhancing the electron-deficient boron center's interaction with halide anions (I−/Br−) via boron–halide bonding.33 Additionally, the fluorine substituents form hydrogen bonds with organic cations (e.g., methylammonium or formamidinium), making BCF a bifunctional passivator that modulates both halide anions and organic cations (Fig. S1). These interactions are anticipated to regulate halide distribution and crystallization kinetics, suppressing ion migration and yielding high-quality perovskite films with large grains and low defect density. For clarity, films processed without and with BCF are labeled control and target, respectively, while inorganic layers are denoted as PbX2 and PbX2 + BCF (X = I or Br).
We firstly experimentally adopted Fourier-transform infrared spectroscopy (FTIR) spectroscopy measurements to elucidate the chemical interactions of BCF with the perovskite precursors. The FTIR spectra of BCF, BCF + PbI2, and BCF + PbBr2 are shown in Fig. 1a. The blueshift in the B–C stretching vibration of BCF molecules from 974 to 987 cm−1 is consistent with the boron–halide binding effect as supported by theoretical calculations. The shift was attributed to the increased electron cloud density of the B–C bond due to the contribution of the electron-rich halides, confirming the boron–halide interaction. Fig. S2a presents the FTIR spectra of the BCF + FAI composite and pure FAI. The characteristic peak corresponding to N–H bonds exhibited a redshift as compared to that of pure FAI, which was attributed to the formation of hydrogen bonds (N–H⋯F) between FA and the fluorine atoms in BCF. A similar N–H vibration redshift supporting the formation of hydrogen bonds (due to a decrease in the N–H bond strength) was observed when comparing the FTIR spectra of the BCF + MABr mixture and pure MABr (Fig. S2b). Thus, the observed spectral changes confirm the existence of strong interactions between BCF and the perovskite precursor components.
Furthermore, the influence of BCF on the PbX2 films was investigated by X-ray photoelectron spectroscopy (XPS) measurements. As illustrated in Fig. S2c, the Pb 4f core-level XPS spectra revealed only a minor shift in the PbX2 + BCF film compared to the PbX2 film, suggesting that lead cations (Pb2+) and BCF likely do not interact directly. Instead, the electronic environment of the [PbX6]4− octahedra appears to be indirectly modified by the boron–halide interaction. This boron–halogen interpretations were further supported by the observed shifts in the XPS spectra of I 3d and Br 3d bands for PbX2 + BCF films due to the decrease in the electron cloud density of the halogen (Fig. 1b and c). These shifts provide additional evidence of the interactions between BCF and the halide anions (I− and Br−), confirming that BCF influences the electronic structure of the inorganic framework.
To further explore these interactions, nuclear magnetic resonance (NMR) spectroscopy was employed which provided additional insights into the chemical interactions of BCF with the perovskite precursor components. First, the 19F NMR spectra revealed some shifts of the signals corresponding to fluorine substituents in para- and meta-positions of the phenyl rings, which are the most distant from the boron center and, hence, they are sterically the most accessible for intermolecular hydrogen bonding and halogen–halogen interactions. These shifts were observed with comparable magnitudes when BCF interacted with PbI2, PbBr2, CsI, FAI, MABr and MACl (Fig. S3).
The 1H NMR spectra have revealed spectacular changes for FAI, MABr and MACl when they were mixed with BCF. Firstly, one could notice that the CH group signal of FAI undergoes splitting to a quintet, which implies that all four neighboring N–H protons become symmetrically non-equal and they do not exchange on the timescale of the NMR measurement (Fig. 1d). Furthermore, the broad peak at 8.83 ppm corresponding to all four NH protons of the pristine FAI undergoes splitting in two distinct signals: a singlet at 9.01 ppm and a doublet at 8.66 ppm, thus further confirming the loss of symmetry of the FA+ cation. Similarly, the peak corresponding to the –NH3+ protons of the methylammonium cation in MABr and MACl becomes much broader upon interaction with BCF (Fig. 1e and f). At the same time, the singlet corresponding to the methyl group (CH3–) of the methylammonium cation shows a spectacular splitting to a quartet, which is only possible if all three protons of –NH3+ attain some fixed geometry and become symmetrically non-equal. All these findings taken together provide univocal evidence for the occurrence of hydrogen bonding interactions mediated by BCF. Very similar situation happens with CsI, which also shows a notable shift in the 133Cs NMR spectra evidencing the change of the coordination environment for the Cs+ cation (Fig. S4).
To theoretically analyze the interactions between BCF and the key perovskite precursor components the density functional theory (DFT) calculations were performed (see the SI for details). According to the DFT calculations, BCF covalently binds halides X− which leads to the formation of complex anions [BCF–X]− (X− = I−, Br−, Cl−), named boron–halide interactions. The bond enthalpies corresponding to the gas-phase binding enthalpy between the halide and BCF are ΔH° = −1.54 eV, −2.01 eV, −2.23 eV and the binding Gibbs free energy under standard conditions are ΔG° = −1.16 eV, −1.62 eV and −1.85 eV for I, Br and Cl, respectively (Fig. 1g–i). The bonds between BCF and MA as well as FA are weaker, with the gas-phase binding enthalpy values ΔH° = −0.67 and −0.53 eV under standard conditions and relatively modest Gibbs free binding energies ΔG° = −0.24 eV and −0.15 eV, respectively, which indicate that the cations would be much less restricted in their motion in the presence of BCF. It is also worth pointing out that the dimerization of BCF with itself has ΔH° = −0.88 eV and ΔG° =–0.17 eV, which means that some energy is required to break apart these noncovalent interactions when mixing it with the salts (Fig. 1j–l). Thus, the organic counter-ions FA+ and MA+ become confined to the corresponding bulky organoboron anions but can likely move relatively freely among them, which results in the formation of hydrogen-bonded complexes with a fixed unsymmetrical geometry. Furthermore, the minimal-energy computed molecular structures of BCF adducts with MABr and FAI were also calculated (Fig. 1m and n). In particular, the complex formation between MABr and BCF is thermodynamically favorable and proceeds with the energy gain of 6.4 kcal mol−1. The Br− anion in this complex forms a short contact with the B center of BCF, which confirms the halide-boron bonding type of interactions. Furthermore, the CH3NH3+ cation forms hydrogen bonds with the Br− counterion and F-substituent of BCF (Fig. 1m). The simulated FTIR spectra (Fig. S5a) showed that the most intense N–H bond stretching peak shifts from 3404 to 3377 cm−1 upon MABr complexation with BCF, which is consistent with the experimental data (Fig. S2b). The computed 1H NMR chemical shifts for the MABr–BCF complex also align well with the experimental results. Thus, the CH3 group is predicted to appear as a quartet with a chemical shift of 2.57 ppm (exp. 2.37 ppm), whereas the NH3+ group protons have a computed chemical shift of 7.72 ppm (exp. 7.57 ppm). The computed difference in the chemical shifts of C–H and N–H protons almost matches the experimental value (5.15 vs. 5.21 ppm, respectively).
The interaction patterns of BCF with FAI appeared to be more sophisticated as compared to the MABr + BCF system. DFT calculations predicted very small energy gain upon 1
:
1 complex formation between BCF and FAI, which cannot compensate for the decreased entropy of the system. However, the association of 2 FAI with 2 BCF molecules produces a very stable complex (Fig. 1n) with the energy gain of 46 kcal mol−1. The I− anions form short contacts with boron atoms of BCF, which confirms the boron–halide bonding effect. Two NH2 groups of the FA cations become strongly non-equivalent in the complex. Two protons of one NH2 group form short contacts with the I− anions, while the protons of another NH2 group participate in the hydrogen bond with the F-substituent of BCF. This arrangement first results in the distinct chemical shifts for protons of two NH2 groups: ca. 9.01 ppm (NH2 bonded to I−) and ca. 8.65 ppm (NH2 bonded to F). For the latter peak, the 1H–19F interactions lead to the signal splitting to the doublet (8.676 and 8.646 ppm). As a consequence of these amine proton bonding effects, the CH proton of the FA cation appears in the NMR spectra as a triplet of triplets. The computed chemical shifts for two types of N–H (9.09 and 8.25 ppm) and CH (7.70 ppm) protons agree reasonably well with the experimental values (9.01, 8.65 and 7.86 ppm, respectively). Finally, the simulated FTIR spectrum of the FAI–BCF interaction product (Fig. S5b) reveals the major N–H stretching band at ∼3256 cm−1, which is shifted to lower wavenumbers with respect to its position for pristine FAI modeled by its tetramer (3315–3325 cm−1), which is consistent with the experimental FTIR spectra shown in Fig. S2a.
The theoretical calculation results perfectly match the experimental FTIR and NMR spectral data, thus establishing reliable structures for the complexes of BCF with FAI, MABr and MACl. These findings collectively confirmed the existence of boron–halide interactions and hydrogen-bonding within the BCF complexes with the perovskite precursor components, which represents the strategic mechanism to regulate the two-step process of crystallization and growth of a wide-bandgap perovskite, thereby producing high-quality absorber films.
Upon dynamically dripping the organic ammonium salt solution onto the lead halide film, rapid nucleation and initial crystal growth were observed in the control perovskite films, as evidenced by an absorption shift to approximately 738 nm (the characteristic absorption edge of the 1.68 eV perovskite) within 2.3 seconds. In contrast, the absorption edge of the target perovskite film redshifted to approximately 738 nm more slowly, taking 3.8 seconds. This delay indicated that the BCF–PbX2 intermediates and hydrogen-bonded complexes slowed the intermolecular exchange process between the organic salt and lead halide, allowing for better controlled crystallization.
During the subsequent thermal annealing process at 100 °C, the target perovskite film displayed a uniform and stable absorption spectrum, whereas the control perovskite film showed slightly weaker absorption features. This discrepancy is attributed to the presence of residual lead halide and Pb0 in the control film as proved by XRD and XPS measurements, which hindered its light absorption efficiency. It is worth noting that a slight redshift from 738 nm to 743 nm occurred at the absorption edge of the control film, which is attributed to the formation of an iodine-rich phase on the film. The in situ UV-vis absorption spectroscopy results demonstrated that the boron–halide interaction and hydrogen-bonded complexes effectively stabilized the uniform distribution of iodide and bromide ions to suppress phase separation, and delays the reaction and crystallization kinetics between the organic ammonium salt and lead halide. This controlled delay was expected to promote the formation of larger grains in the perovskite films with fewer grain boundaries and enhance the overall quality of the perovskite films.
Fig. 2e presents the X-ray diffraction (XRD) patterns of the PbX2 and PbX2 + BCF films. The intensity of the diffraction peaks corresponding to the PbX2 crystal planes decreased with the introduction of BCF, attributed to the formation of a fluffy, porous morphology upon incorporating the BCF additive into the PbX2 film. To further investigate the morphological changes induced by BCF, the scanning electron microscopy (SEM) and atomic force microscopy (AFM) measurements were carried out. The top-view SEM showed that the addition of BCF results in a significant decrease in the grain size of PbX2, which consequently leads to an increase in the areal density of grain boundaries and voids (Fig. 2f and g). Furthermore, the cross-sectional SEM images clearly demonstrated that the addition of BCF resulted in the formation of numerous voids and interlaced patterns within the PbX2 + BCF layer (Fig. S6), which is not observed in the PbX2 films. These findings confirm the strong influence of BCF addition on the morphology of the PbX2 films, which facilitated the sufficient diffusion of the organic ammonium salts during the perovskite formation process. This observation was supported by the increase in root mean square (RMS) roughness from 6.7 nm for the PbX2 film to 7.8 nm for the PbX2 + BCF film, as measured by AFM, further confirming the morphological changes induced by BCF (Fig. S7). Infrared scattering-type scanning near field optical microscopy (IR s-SNOM) was applied to further characterize the PbX2 + BCF films. This technique enables simultaneous AFM topography measurement with the collection of local spectral data from the area comparable to the tip radius of the cantilever (20–30 nm).34,35 Thus, IR s-SNOM could reveal the distribution of the BCF molecules within the target PbX2 + BCF films. The major part of BCF was accumulated at the grain boundaries, and also observed on the surface of the PbX2 grains (Fig. S8), which confirms that the boron–halide interactions control the growth of the PbX2 + BCF films.
To investigate the influence of the BCF modifier within the PbX2 precursor on the perovskite films crystallinity, we analyzed the X-ray diffraction (XRD) patterns of the perovskite film. Both types of the perovskite films processed without and with BCF revealed a peak at ≈13°, attributed to residual unreacted PbX2 (Fig. 2h). The intensity of this PbX2 peak was significantly lower for the target perovskite film as compared to the control film. This reduction suggested that the incorporation of BCF facilitated PbX2 conversion in the reaction with the organic ammonium salts and improved interdiffusion of the organic and inorganic components. This optimization promoted the growth of high-quality perovskite crystals. Notably, trace amounts of lead halide remained in the target perovskite film, which was expected to passivate surface and grain boundary defects, thereby improving device performance. Furthermore, the intensity of the (111) diffraction peak was notably stronger in the target perovskite film, indicating that BCF promoted the growth of the (111) crystal plane.
To further evaluate the effects of BCF on perovskite film quality, surface morphological characterization was conducted using scanning electron microscopy (SEM) and atomic force microscopy (AFM). The SEM images (Fig. 2i and j) revealed that the target perovskite films consist of larger grains with fewer grain boundaries (GBs) compared to the control films, which was attributed to the slower chemical reaction and crystallization kinetics between the organic ammonium salts and the lead halides due to the boron–halide and the hydrogen bonding interactions.36,37 This improvement of the perovskite grains was corroborated by cross-sectional SEM images (Fig. S9), which showed no distinct grain stacking in the vertical direction in the case of target films, facilitating efficient charge carrier transport. Additionally, the target perovskite film exhibited a higher water contact angle (72.40°) than the control film (53.47°) (Fig. S10), demonstrating that BCF treatment enhanced the film's ability to resist water intrusion, thereby improving its stability under ambient conditions and thus enhancing its potential for real-world applications.
AFM measurements (Fig. S11) further confirmed that the surface roughness (RMS) of the target perovskite films decreased from 27.8 nm to 20.3 nm, promoting better interfacial contact and charge transfer with the subsequently deposited ETL. Kelvin probe force microscopy (KPFM) was employed to explore the alignment of the interface energy levels. As shown in Fig. S12 and Fig. S13, the surface contact potential distribution (CPD) of the target perovskite film is more uniform, within a range of ±10 mV, compared to the control film, which exhibits a broader range of ±15 mV. This uniformity was attributed to the more homogeneous surface structure and a more controllable crystallization and growth process of the target perovskite films. The KPFM results also suggested that the target perovskite film exhibited a more positive surface potential, indicating that the Fermi level on the perovskite surface is closer to the conduction band to form a more n-type surface electric potential of the film. This alignment is favorable for efficient electron extraction and transfer, further enhancing device performance.
In order to evaluate the influence of BCF additives on the residual stress of the perovskite films, the surface residual stress from 100 nanometers depth below the surface of the perovskite film was measured by deep-resolution grazing incidence X-ray diffraction (GIXRD) combined with the 2θ-sin2
ψ method (Fig. 2k and l). As is well known, the negative slope of the 2θ-sin2
ψ method linear fitting indicated the residual tensile stress in the perovskite film, and the positive slope of the 2θ-sin2
ψ method linear fitting indicated the compressive strain in the perovskite film. Obviously, due to the formation of the boron–halide interactions and hydrogen-bonded complexes between BCF molecules and perovskite precursor components, the slope of 2θ-sin2
ψ variation changed from negative to positive, indicating that stress changed from tensile stress (183.1 MPa) of the control perovskite film to compressive stress (−25.4 MPa) of the target sample. These findings suggested that BCF released the strain during thermal annealing and exhibited an excellent ability to regulate grain growth and morphological structure, and improved device performance and mechanical stability (Fig. 2m and Fig. S14).
An additional insight into the effects of the BCF additive on the nanoscale structure of the grown perovskite films was gained using IR s-SNOM. Using this technique, we could individually follow the nanoscale distribution of organic cations (MA+, FA+) and the BCF additive in the perovskite films. The obtained results are presented in Fig. 3. For the control perovskite film, we observed very uniform distribution of the methylammonium and formamidinium cations in the film. In contrast, the target perovskite films showed some minor local gradients in the concentrations of MA+ and FA+ cations thus creating a “perovskite bulk heterojunction” system, which we have shown recently is highly beneficial for the PSC performance.38 Furthermore, we demonstrated that BCF effectively covered the perovskite film surface and showed the highest abundance at the grain boundaries, thus contributing to the passivation of defects and suppressing the nonradiative trap-assisted recombination of charge carriers.
Thus, the obtained IR s-SNOM imaging data corroborated with the results provided by other techniques and proved unambiguously that the introduction of BCF had a highly positive influence on the nanoscale morphology and structure of the perovskite absorber films grown using the two-step deposition method.
The charge transfer and recombination kinetics of the perovskite films were investigated using steady-state photoluminescence (PL) and time-resolved photoluminescence (TRPL) measurements. The target perovskite film exhibited higher PL intensity compared to the control film, indicating reduced density of defect states and improved film quality due to effective defect passivation (Fig. 4d). TRPL results (Fig. 4e) were fitted using eqn (S1) and S2, revealing an increase in the average carrier lifetime from 731.25 ns to 1180.70 ns with a slightly enhanced rapid decay lifetime (τ1) and a significantly enhanced slow decay lifetime (τ2), which indicates that the introduction of BCF not only significantly passivated the bulk defects in the perovskite film but also inhibited non-radiative recombination at the interface (Table S1). On the bare glass substrate without the charge extraction pathway, the perovskite films only exhibited the characteristic slow PL decay due to the dominance of bulk defect-mediated recombination.39
Transient photocurrent (TPC) decay testing was used to characterize the carrier extraction and transport capabilities within perovskite solar cells. As shown in Fig. 4f, the target device exhibited a shorter decay lifetime of 268.9 ns, compared to 520.9 ns for the control device. The faster TPC decay indicated that the addition of BCF can facilitate the transport and transfer of photogenerated charges, demonstrating the stronger carrier extraction ability of the target device. The device transient photovoltage (TPV) decay test was employed to further reveal the nonradiative recombination of charge carriers within and at the interface of perovskite films when the device is in an open-circuit state (Fig. 4g). The decay lifetime of TPV for the target device is 381.2 μs, longer than that of the target device (352.4 μs). The addition of BCF effectively suppresses nonradiative recombination within and at the interface of perovskite films, thereby prolonging the carrier lifetime in the target devices. These findings collectively affirm that the introduction of BCF molecules can promote carrier extraction and reduce nonradiative recombination losses, thereby realizing high-performance photovoltaic devices.40
To quantify the nonradiative recombination loss at the perovskite film interface, we measured the photoluminescence quantum yield (PLQY) to estimate the quasi-Fermi level splitting (QFLS) relevant to the theoretically feasible VOC (Fig. 4h and eqn (S3)). The PLQY increased from 2.544% to 3.602%, indicating an improvement in the efficiency of treating perovskite with BCF on glass substrates. The results showed that the average QFLS increased from 1.306 eV to 1.315 eV, which was directly related to the increment of VOC in PSCs. Significantly, the PVK/ETL films exhibited a remarkable PLQY drop, which was attributed to additional defects introduced by the ETL. After the addition of the ETL, the QFLS of the control film decreased to 1.265 eV, indicating a 41 mV increase in nonradiative recombination loss, while the QFLS of the BCF/ETL remained stable at 1.298 eV, decreasing by only 17 mV (Table S2). These results emphasized that BCF effectively reduced VOC loss by suppressing nonradiative recombination.
Furthermore, the defect density of states (Nt) in the perovskite films was quantitatively evaluated using eqn (S4). Hole-only devices (ITO/NiOx/Me-4PACz/Perovskite/Spiro-OMeTAD/Ag) were fabricated for space charge-limited current (SCLC) measurements. As can be seen in Fig. 4i, Nt of the target film decreased from 6.73 × 1015 cm−3 to 4.01 × 1015 cm−3 compared to the control film (Table S3). This reduction in Nt enhanced charge mobility, further supporting the role of BCF in improving perovskite crystal quality.
Photoinduced phase segregation, primarily driven by ion migration through halide vacancies, is a critical factor which limits the performance and longevity of wide-bandgap perovskite films. The evolution of PL spectra under one-sun illumination was analyzed to evaluate halide segregation. In particular, the control perovskite film exhibited a significant redshift in its PL spectra, indicating the formation of narrow bandgap I-rich defect states due to phase segregation (Fig. S16a). In contrast, the BCF-treated film showed suppressed photoinduced phase segregation (Fig. S16b), suggesting that BCF contributes to the formation of more stable wide-bandgap perovskites. To quantify these changes, we calculated the PL peak shifts after 90 minutes of illumination. The BCF-treated wide-bandgap perovskite films showed a smaller redshift by 9 nm, compared to the 21 nm of the control film. This smaller shift demonstrated BCF's ability to inhibit photoinduced phase segregation, which was likely due to defect passivation and halide immobilization on the perovskite grain surface by boron–halogen interactions.
Strong evidence for the remarkable stabilizing effect of BCF was obtained using unique in situ IR s-SNOM measurements that to the best of our knowledge have been performed for lead halide perovskite films for the first time. The tested samples were irradiated with blue laser light (460 nm, ca. 100 mW cm−2) directly at the microscope stage. Running a continuous measurement produced a series of IR s-SNOM images for control and target samples, which could be compared on the same timescale. From the Video S1, it is clearly seen that the target perovskite film demonstrates much higher stability under light exposure than the control sample. Indeed, a strong depletion of organic formamidinium cations is observed in control films after 2 h of light-induced aging, while the target films remain almost intact under the same conditions. These findings directly prove a strong stabilization effect of BCF and its ability to suppress light-induced ion migration and degradation of inorganic cations.
The enhanced stability of wide-bandgap perovskites must correlate with the ability of BCF to suppress the photoinduced segregation of Br-rich and I-rich phases. Mechanistically, the halide phase segregation process involves the formation of Pb0 and I3− species, their accumulation at the grain boundaries, and subsequent recombination (Pb0 + I3− → [PbI3]−) with the formation of an I-rich phase.41,42 We have shown that BCF significantly improves the photostability of PbX2 films (I
:
Br = 3
:
1) and suppresses the formation of metallic lead in the samples exposed to 100 mW cm−2 illumination at a temperature of 77 ± 3 °C in a pure nitrogen atmosphere (Fig. S17). This finding provides clear evidence that BCF controls the behavior of PbX2 films, suppresses Br–I segregation, and enhances inorganic phase photostability, which finally leads to the stabilization of wide-bandgap perovskites.
To gain a further mechanistic insight, we modelled the behavior of the PbI2–BCF system using DFT calculations for a simple cluster model composed of four interconnected ribbons of PbI2 (Fig. S18). This cluster has a series of coordinationally unsaturated Pb2+ ions surrounded by 3, 4 or 5 iodide anions, which can be considered as a realistic model for the PbI2 grain surface. DFT calculations demonstrate that the Pb0 atom appearing within the PbI2 structure tends to pair with one of the Pb2+ cations and forms a [Pb–Pb]2+ molecule-like fragment. The maximal energy gain of 46.2 kcal mol−1 is reached when the [Pb–Pb]2+ fragment is localized on the surface of the PbI2 cluster (Fig. S18a). Interestingly, the presence of more Pb0 atoms within PbI2 makes them clustering together forming Pb2, Pb4 and even bigger species, which forms the initial stages of the metallic lead phase nucleation process.
However, the presence of BCF in the system drastically changes its dynamics since BCF was shown to bind to Pb0 with the energy gains varied from 4 to 14 kcal mol−1 depending on the Pb0 localization. The most realistic structure shown in Fig. S19 is characterized by the surface localization of metallic lead with Pb0 and BCF binding energies of 44.2 and 13.1 kcal mol−1, respectively.
In the further modelling, we introduced an I2 molecule into the system, forming a triiodide (I3−) fragment on the surface of the PbI2 cluster. Thus, the set of structures presented in Fig. S19 represents the intermediate state with the photogenerated I3− and Pb0 species, which could evolve in several ways. First, I3− could release I2 as a vapor into the gas phase and then the material decomposition becomes irreversible with the accumulation of metallic lead as a distinctive feature. Second, Pb0 could nucleate as small metallic clusters, then react with I3− and form a separate iodide-rich phase, thus leading to the I–Br phase segregation. However, BCF effectively binds to Pb0 and suppresses the metallic lead nucleation process. Furthermore, our calculations show that when the distance between I3− and Pb0 species reduces, as illustrated by the transition from the structure A to structure B in Fig. S19, the BCF binding energy also decreases dramatically from 11.6 to 5.4 kcal mol−1. Simultaneously, the I2 binding energy also becomes twice smaller (20.2 and 10.3 kcal mol−1 for structures A and B), which provides ideal conditions for BCF desorption from the surface and recombination of Pb0 with I2 leading to the regeneration of PbI2.
Thus, our modelling unravels the unique mechanism, which involves reversible binding of BCF to Pb0 on the PbX2 surface (blocking metallic lead phase nucleation) and subsequent release of Pb0 when I3− diffuses closer, thus promoting their recombination and PbI2 formation. This extraordinary BCF behavior makes it a strategic stabilizer of perovskite absorber films with potential of showing tremendous improvements in the PSC operational lifetime.
Additionally, the integral short circuit current density (JSC) derived from external quantum efficiency (EQE) measurements for the target PSCs was 21.21 mA cm−2, closely matching the JSC value of 21.87 mA cm−2 obtained from J–V measurements under simulated AM 1.5G illumination (Fig. 5c). Statistical analysis of the photovoltaic parameters (Fig. 5d and e) confirms the reproducibility of these BCF-optimized devices. The performance enhancement is attributed to improved trap passivation and crystallinity facilitated by the BCF additive. The target device exhibited a stable power output (SPO) efficiency of 23.2% when measured at the maximum power point (Vmax = 1.12 V) (Fig. 5f), consistent with the PCE obtained from the J–V curve.
Furthermore, the relationship between VOC and the logarithm of the illumination intensity yielded fitted slopes of 1.45kT/q for the control device and 1.26kT/q for the target device, as calculated using eqn (S5) (Fig. 5g). Dark J–V measurements (Fig. 5h) revealed that the BCF-treated devices exhibited a lower dark saturation current and reduced leakage current compared to the control device, indicating suppressed leakage pathways and improved diode ideality. This suggests a reduction in trap-assisted recombination, further corroborating the higher PCE of the BCF-modified device.
To investigate the influence of the BCF additive on the optical properties of the perovskite photoactive layer, UV-vis absorption spectra were measured. Fig. S24 demonstrates that the target perovskite film exhibited enhanced light absorption intensity, primarily due to improved perovskite crystal quality. Tauc plots derived from the UV-vis absorption data confirmed a bandgap (Eg) of 1.68 eV for the perovskite absorber layer. Ultraviolet photoelectron spectroscopy (UPS) was also employed to analyze the optoelectronic properties of the perovskite films (Fig. S25). The UPS results indicated a decrease in work function (WF) from 4.31 eV (control) to 4.25 eV (target), consistent with Kelvin probe force microscopy measurements (Table S9). The upward shift in WF suggests that the perovskite surface exhibits n-type doping characteristics after BCF incorporation, which is expected given that BCF is a Lewis base. The energy level diagram for the PSC functional layers is presented in Fig. 5i. The introduction of BCF optimizes the energy level alignment at the perovskite/ETL interface, enhancing carrier extraction and contributing to the higher PCE achieved in the photovoltaic devices. These findings underscore the significant role of BCF in improving the performance of wide-bandgap PSCs through enhanced crystallinity, trap passivation, and optimized energy level alignment.46
Comprehensive stability studies were then carried out in accordance with the stability protocols for PSCs. In ambient stability tests, the unencapsulated devices were stored in air with a relative humidity of 55% at 25 ± 5 °C and their performance was regularly monitored, as can be seen in Fig. 5j. After ambient storage for approximately 50 days, the control devices maintain, in average, only 50% of their initial PCE, while the target series of PSCs maintain, in average, 90% of their initial PCE, respectively, which is attributed to the high-quality perovskite films and a higher exposure ratio of the (111) crystal plane of target devices. The heat stability of unencapsulated devices at 85 °C with day/night cycles was assessed. Remarkably, PSCs based on BCF retained above 90% of their initial efficiency after 300 h, compared with 60% retention (Fig. 5k). Operating stability was assessed under harsh conditions in N2 at a temperature of approximately 50 °C (unencapsulated, continuous 1 sun illumination, and negative electric bias) (Fig. 5l). The PSCs employing BCF as a modifier retain 90% of their initial PCE value after 750 h, in sharp contrast to the low 50% PCE retention observed for the control devices.
Supplementary information (SI) is available. See DOI: https://doi.org/10.1039/d5ee03984c.
| This journal is © The Royal Society of Chemistry 2025 |