Jingyu Cai,
Congcong Tian,
Anxin Sun,
Jinling Chen,
Xiling Wu,
Beilin Ouyang,
Jiajun Du,
Ziyi Li,
Rongshan Zhuang,
Ran Li,
Teng Xue,
Tiantian Cen,
Kaibo Zhao,
Yuyang Zhao,
Qianwen Chen and
Chun-Chao Chen*
School of Materials Science and Engineering, Shanghai Jiao Tong University, Shanghai 200240, P. R. China. E-mail: c3chen@sjtu.edu.cn
First published on 16th July 2025
The presence of iodine vacancy defects and iodine impurities on the surface of perovskite can greatly limit the efficiency and stability of perovskite solar cells (PSCs). Therefore, it is crucial to develop a surface iodine management strategy that can passivate surface defects and eliminate harmful I2. This work reports the use of sodium triacetoxyborohydride (STAB) as a reductive anionic passivator for post-treatment of the perovskite surface. The study shows that the multiple electron-rich carbonyl groups of STAB can multidentately coordinate with undercoordinated Pb2+ and suppress iodine vacancy defects. Meanwhile, the reductive nature of STAB prevents the conversion of I− to I2. This alleviates surface iodine loss, reduces the byproduct Pb0, and effectively mitigates the decomposition of perovskite films. Through this synergistic surface iodine management strategy of passivation and reduction, STAB-treated PSCs achieved a power conversion efficiency (PCE) of 25.85%. After 1000 hours of maximum power point tracking at 65 °C, they still retained more than 90% of their initial efficiency. When STAB was applied to narrow-bandgap 1.27 eV perovskite, the PCE increased from 20.02% to 22.97%, demonstrating the universality of this method for different bandgap perovskite systems.
The soft ionic nature of metal halide perovskites leads to a high defect density in the films, with the defect density at the interface being much higher than that in the bulk. This results in severe carrier trapping and non-radiative recombination at the interface, leading to undesirably low voltages in device performance.8,9 Specifically, due to the instability of Pb–I bonds and the volatilization of iodine components during annealing, the upper surface of perovskite films contains a large number of iodine-related defects, such as iodine vacancies (VI) and uncoordinated Pb2+ defects. Post-treatment with ammonium halides can passivate A-position defects while introducing additional iodine ions to compensate for surface VI.10–12 However, the introduction of 2D perovskites results in high exciton binding energy and pronounced quantum confinement effects,13–16 which hinder carrier transport. To avoid the introduction of 2D phases, anion passivators have attracted growing interest.17–20 Halides, especially the iodide ion (I−), are widely adopted anions, which can fill iodine vacancies and suppress surface defects.21–24 Wu et al. introduced CdI2 to compensate for the surface I− ion deficiency, forming stronger Cd–I bonds relative to Pb.21 Subsequently, the interfacial charge recombination loss is reduced, yielding an open-circuit voltage of 1.20 V. However, excessive iodide ions can induce interstitial iodine defects,25 while also causing surface composition disorder,26 which impairs device performance and stability. Compared to halide anions, nonhalide anions usually contain groups with lone pair electrons, such as S, O, and F, which have been proven to strongly coordinate with uncoordinated Pb2+, thus more effectively filling and passivating iodine vacancy defects. Some typical nonhalide anions include CH3COO−, BF4−, and SCN−.27–29 However, these nonhalide anions usually only bind to a single undercoordinated lead site, with limited binding strength. Therefore, anion passivators that can simultaneously bind to multiple uncoordinated Pb2+ sites with stronger passivation capability are worth exploring.
Besides, I− undergoes oxidation reactions under external conditions such as light, thermal stress, and electric fields, generating I2.30,31 I2 has been proven to be highly corrosive, leading to the destruction of the perovskite structure.32 Meanwhile, the generation and volatilization of I2 cause surface iodine loss and create deep defects such as Pb0 on the perovskite surface,33,34 resulting in the degradation of PSC device performance. The use of reducing agents has been demonstrated to eliminate I2. For example, benzylhydrazine hydrochloride (BHC)35 and potassium formate (HCOOK)30 have been introduced into the bulk, mitigating I2-induced film stability issues by reducing I2 to I−. Notably, borohydrides (BH4−) and their derivatives possess broad reducibility36,37 and have been applied in the inhibition of Sn2+ oxidation in tin-lead perovskites,38,39 but there is currently limited research on their use in reducing I2. Liu et al.40 used potassium borohydride (KBH4) on a NiOx substrate, preventing adverse interfacial reactions caused by Ni2+ oxidation and reducing the I2 content within the film, demonstrating the effective reduction of elemental iodine by borohydrides. However, KBH4 is extremely reactive and flammable, making it unsuitable for large-scale production. At present, no suitable borohydride derivative has been discovered to simultaneously reduce I2 and provide interfacial passivation for synergistic improvements in PCE and stability.
In this work, a mild borohydride derivative, sodium triacetoxyborohydride (STAB), was introduced as a reductive anion passivator for post-treatment of perovskite. Density functional theory (DFT) calculations revealed that in the presence of iodine vacancy (VI) defects, STAB, containing three electron-rich carbonyl groups (CO), simultaneously coordinates with multiple undercoordinated Pb2+ defect sites on the perovskite surface. As a result, TAB− achieved higher adsorption energy compared to commonly used anions (I−, Br−, and Cl−), significantly passivating surface iodine vacancy defects. Additionally, the strong reducing B–H bond in STAB can reduce I2 to I−. UV-vis spectra show that the reducing effect of the B–H bond inhibits the generation of oxidative I3− in FAI and significantly reduces the extraction of I2 from the perovskite film under light and thermal stress. Meanwhile, the reduction of I2 in the film by STAB helps to alleviate the surface iodine deficiency and eliminates harmful Pb0 byproducts. Therefore, the defect passivation and I2 reduction effects brought by STAB post-treatment promote less interface recombination loss and a more uniform, high-quality surface. As a result, the inverted PSCs treated by STAB achieved a champion efficiency of 25.85% and a high open-circuit voltage of 1.185 V. Meanwhile, the target devices retained over 90% of their initial PCE after 1000 hours of thermal aging at 85 °C in a N2 atmosphere or 1000 hours of maximum power point tracking (MPPT) at 65 °C. Moreover, we applied this strategy to narrow-bandgap (1.27 eV) PSCs and increased the PCE from 20.01% to 22.97%, while achieving better thermal stability. This demonstrates the good universality of this strategy for various bandgap perovskite solar cells.
Further detailed characterization studies of the interaction mechanism between STAB and perovskite films were conducted. First, time-of-flight secondary-ion mass spectrometry (ToF-SIMS) depth profiles were performed to clarify the distribution of STAB in perovskite films,43 as shown in Fig. S1.† Since B+ accumulated at the surface, this indicates that the anions of STAB are mainly present at the surface of the perovskite. Na+, in contrast, was distributed throughout the perovskite film and was enriched both at the surface and at the buried interface.44 It has been reported that the presence of Na+ in the bulk and at the interfaces suppresses ion migration and reduces hysteresis.45,46 Then, the detailed spectroscopic tests were conducted to further investigate the interaction between STAB and perovskite components. The FTIR spectra of a series of samples, including STAB powder, STAB + PbI2 mixture, FAI, and FAI + STAB mixture, were analyzed, as shown in Fig. 2a. After the addition of PbI2, the CO peak of the C
O bond in STAB shifted from 1683.1 cm−1 to 1650.7 cm−1. In addition, the FTIR spectra of the control and STAB-treated perovskite films were measured, as shown in Fig. S3.† The stretching vibration of C
O of STAB appeared in STAB-treated perovskite films and shifted from 1683.1 to 1653.7 cm−1. These results indicate a strong interaction between STAB and perovskite, which can be attributed to the coordination of carbonyl oxygen in TAB− with Pb2+.47 Additionally, the peaks of NH and CH in FAI moved from 3338.5 cm−1 and 1692.5 cm−1 to 3350.2 cm−1 and 1698.4 cm−1, respectively, with the peak of NH shifting to 3350.2 cm−1 after adding STAB (Fig. S4†), which may be attributed to the hydrogen bonding between the highly electronegative oxygen atoms in STAB and FAI. The proton nuclear magnetic resonance (1H NMR) spectra (Fig. S5†) showed that the chemical shift of N–H in FA+ exhibited an upfield shift after mixing with STAB, further confirming the presence of hydrogen bonding. This hydrogen bonding has been proven to stabilize the surface structure, which is beneficial for reducing the generation of VI.7 X-ray Photoelectron Spectroscopy (XPS) tests were performed on perovskite films to elucidate the changes in the chemical environment on the film surface. Fig. S6† shows the signal of B 1s. The presence of boron on the surface of STAB-treated films, which was absent in the control films, indicates that STAB was successfully introduced onto the perovskite surface. As shown in Fig. 2b, the binding energies (BE) corresponding to the Pb 4f5/2 and Pb 4f7/2 electronic states in the control film were 143.35 eV and 138.49 eV, respectively. In contrast, the Pb 4f spectrum of the STAB-treated film shifted towards lower binding energy by 0.16 eV, further confirming that undercoordinated Pb2+ formed coordination bonds with carbonyl groups, leading to an increase in electron density around Pb. Interestingly, the control film showed two shoulder peaks at 141.72 eV and 136.62 eV, indicating the presence of trace amounts of Pb0 on the surface.48 Pb0 is known as a primary deep defect state that severely degrades the performance and durability of perovskite optoelectronic devices. The peaks of Pb0 were eliminated after STAB treatment, indicating that the multi-carbonyl coordination of STAB reduced Pb-based defects in the film.
Due to the reducing nature of the BH bond in STAB, the inhibition of oxidation of I− by STAB is investigated here. The above result shows that the inhibition of VI by STAB can block the generation of VI-mediated I0 species.49 Therefore, the presence of STAB should be beneficial in suppressing the formation of iodine in the film, thereby improving the film's photostability. To verify this hypothesis, STAB was first added to an IPA solution containing I2 and left to stand for 5 minutes, observing that the solution changed from dark brown to transparent and colorless (Fig. 2d). FTIR was used to examine the changes in chemical groups after mixing STAB with I2, as shown in Fig. 2d. The scissoring bending vibration of B–H at 1106 cm−1 disappeared after mixing with I2,50,51 proving the role of the reducing B–H bond in STAB in eliminating I2. Furthermore, FAI was dissolved in IPA solvent and subjected to light aging at 55 °C for two hours. As shown in Fig. 2e, the aged FAI solution showed a new peak at 360 nm,40 indicating the oxidation of I− to I2. In contrast, the FAI solution with added STAB did not show a peak at this position, indicating that STAB could prevent the oxidation of I− under light.23,30 The I/Pb ratio and Pb0 content in the perovskite films were calculated based on XPS (Fig. 2f), and the results showed that the STAB-treated films had a higher I/Pb ratio and less content of Pb0. At the same time, the XPS spectra of the I 3d signal (Fig. 2c) shifted (initially at 630.9 and 619.4 eV) to 630.7 and 619.2 eV, respectively. These results indicate that the loss of I− due to the oxidation to I2 in STAB-treated samples was reduced,48 and the by-product Pb0 was also reduced.15 To study the inhibitory effect of STAB on I2 in perovskite films, two perovskite films (control and STAB-treated) were immersed in toluene under 1 sun continuous illumination and tested for the presence of I2 produced at different times using UV-vis absorption spectra, as shown in Fig. 2g and h. After 12 hours of illumination, the I2 absorption peak (∼500 nm) of the control film was significant, attributed to the degradation of the perovskite film during light aging. In contrast, the STAB-treated film showed a weak peak. The intensity of the I2 absorption peak at different times was statistically analyzed (Fig. 2i). It can be seen that STAB significantly reduced the escape of I2 from the perovskite film, delaying the degradation of the perovskite film and improving the film's photostability.
The effects of STAB post-treatment on the crystal structure and surface morphology of perovskite films were investigated. The results of Grazing Incidence Wide Angle X-ray Scattering (GIWAXS) and X-ray diffraction (XRD) (Fig. 3a and S7†) both demonstrated that the crystal structure remained unchanged after STAB treatment. The crystallinity slightly improved, which can be attributed to the passivation of under-coordinated Pb defects and post-treatment annealing.42,52 Additionally, the reduction of PbI2 was observed in GIWAXS, XRD, and SEM (Fig. 3b). As a photo-unstable substance, the reduction of PbI2 will benefit the film's tolerance to light.53 The UV-vis absorption spectra of the perovskite film (Fig. S8†) indicated that STAB treatment almost did not alter the bandgap width of the perovskite and had no significant impact on the light absorption properties. Atomic force microscopy (AFM) and Kelvin probe force microscopy (KPFM) were used to study the films (Fig. S9† and 3c). The STAB film showed a lower root mean square roughness (RMS) compared to the control group, decreasing from 21.8 nm to 17.4 nm. Furthermore, the average contact potential difference (CPD) of perovskite films significantly decreased from 50.58 mV to 42.93 mV and became more uniform after STAB treatment (Fig. 3d). These observations indicated that a smoother and more uniform surface is formed by STAB treatment, which could be beneficial for achieving better interfacial contact and promoting efficient interfacial charge extraction.11
![]() | ||
Fig. 3 (a) GIWAXS patterns, (b) SEM images and (c) KPFM images of the control and STAB perovskite films. |
To investigate the passivation effect of STAB, steady-state PL and time-resolved photoluminescence (TRPL) were measured on glass/perovskite samples to study the carrier dynamics of the films. As shown in Fig. 4a and b, the STAB-treated film exhibited significantly enhanced PL intensity compared to the control perovskite film, indicating suppressed non-radiative recombination due to effective passivation of surface defects. TRPL demonstrated that STAB treatment significantly increased the average carrier lifetime (τave) of the perovskite from 546.21 ns to 1725.10 ns (bi-exponential function fitting was applied; detailed parameters are listed in Table S1†), which is consistent with the PL result. To further investigate the effects of STAB on charge transport, a semi-complete device with a glass/perovskite/PCBM structure was prepared. The control sample showed a strong PL peak, while the intensity of the PL peak of the STAB-treated sample was only about 33% of the control sample, indicating fast fluorescence quenching. Meanwhile, the TRPL results showed that the τ1 of the STAB-treated sample is shorter (14.55 ns) than that of the control sample (18.76 ns), indicating faster charge extraction from the STAB-treated perovskite surface to the PCBM layer.4,54 At the same time, the τ2 of the STAB-treated sample is longer than that of the control sample (273.17 ns vs. 223.22 ns), suggesting suppressed non-radiative recombination at the interface. These results demonstrate that STAB promotes electron extraction at the perovskite/PCBM interface and hinders non-radiative recombination within the perovskite.28 The c-AFM tests (Fig. 3e) were conducted to further investigate the charge transport properties of the film surface. The control film had a lower average current, indicating high carrier transport loss.55 Moreover, the grain boundaries appeared brighter, suggesting the presence of leakage current.56,57 After STAB treatment, the average surface current was increased from 2.46 pA to 2.93 pA, and the contrast between grain boundaries and grains was reduced, indicating reduced carrier recombination at the grain boundaries. The frequency distribution of the current is plotted in Fig. 3f. The STAB-treated film showed stronger current signals and a narrower current distribution, indicating that good defect passivation effects enhanced both conductivity and uniformity of the film surface. These results collectively demonstrate the excellent defect passivation and carrier extraction promotion capabilities of STAB.
UPS (Ultraviolet Photoelectron Spectroscopy) was used to study how STAB affects the surface band structure of perovskite films. Fig. 4g shows the secondary electron cut-off edge and the Fermi edge for both the control and STAB-treated films. The STAB-treated film has a slightly lower work function (WF) of 4.15 eV compared to the control's 4.17 eV, which aligns with the lower surface potential measured by KPFM. Based on the film's bandgap (1.55 eV) from absorption spectra, the full band alignment was calculated, as shown in Fig. 4i. The STAB film has an EF (Fermi level) further away from the valence band minimum (VBM) (1.26 eV vs. 1.12 eV), indicating stronger n-type surface characteristics. Additionally, the conduction band minimum (CBM) of the STAB-treated film is −3.86 eV, closer to the LUMO of PCBM (−3.92 eV) than the control's −3.74 eV, suggesting better band-alignment compatibility. These findings are consistent with earlier PL (photoluminescence) and TRPL (time-resolved photoluminescence) studies, showing the significant effect of STAB treatment to obtain a high-quality surface with suppressed non-radiative recombination and enhanced electron extraction.
Given the improved performance of STAB-modified films, we built inverted (p–i–n) perovskite solar cells (PSCs) to test the effect of STAB post-treatment on device performance. The device structure is FTO/MeO-4PACz/perovskite/STAB/PCBM/BCP/Ag (Fig. 5a). The STAB concentration was first optimized for the best results. Fig. 5b shows that 1 mg ml−1 is the optimal concentration, offering the highest average efficiency and reproducibility. Higher concentrations (2 mg ml−1) can lead to excessive STAB deposition at the interface, hindering charge transport. The J–V curves in both reverse and forward scanning directions were measured under standard AM 1.5G sunlight. Fig. 5c and Table S2† display the J–V curves and photovoltaic parameters of the best-performing control and STAB-treated devices. The champion STAB-treated device achieved a PCE of 25.85%, which was significantly higher than the 22.21% of the control device. Notably, the open-circuit voltage (VOC) increased from 1.122 V to 1.185 V, and the hysteresis index (HI) improved dramatically, dropping from 4.23% to 1.81%. This is due to the ability of STAB to suppress iodine vacancies and passivate undercoordinated Pb2+ ions, thus reducing ion migration. The introduction of Na+ can also contribute to the reduction of hysteresis.23 Due to better film quality and charge extraction, the short-circuit current density (JSC) increased from 24.90 mA cm−2 to 25.25 mA cm−2, and the fill factor (FF) also improved from 79.50% to 86.36%. The integrated current density from the EQE (External Quantum Efficiency) spectrum (Fig. 5d) matched the JSC value at 25.08 mA cm−2, confirming the accuracy of J–V testing. Fig. 5e shows the results of the steady-state power output (SPO) test. The control device's steady-state PCE is 21.53%, while the champion device achieves a steady-state PCE of 25.47%. Meanwhile, the device with a large area (1.028 cm2) achieved a PCE of 24.39%, a VOC of 1.179 V, a JSC of 25.06 mA cm−2, and an FF of 82.55% (Fig. S10†). The improved performance of the device at a larger scale can be attributed to the uniformity of the perovskite surface passivated by the STAB treatment, demonstrating the potential of this method for large-area fabrication. To demonstrate the method's versatility, STAB treatment was also applied to tin–lead perovskite solar cells with a 1.27 eV bandgap. The PCE of inverted devices based on Cs0.25FA0.75Pb0.5Sn0.5I3 increased from 20.02% to 22.97%, with a JSC of 32.74 mA cm−2, consistent with EQE results (Fig. S11 and S12,† photovoltaic parameters in Table S3†). The performance enhancement can be attributed to STAB's reducibility, which prevents oxidation of Sn and I components, and its excellent defect passivation of iodine vacancies.39 The 20 tested devices also showed that STAB-treated ones had the narrowest efficiency distribution, indicating high reproducibility (Fig. S13†). The 1.27 eV device treated with STAB exhibited a more stable photocurrent output of 30.04 mA cm−2 within 300 s, as shown in Fig. S15.† These results highlight STAB's universal effectiveness as a reducible anion passivator in enhancing device performance for both 1.55 eV and 1.27 eV bandgap perovskite solar cells.
Then, a series of performance characterization studies were conducted to explore the mechanisms behind the excellent optoelectronic performance. First, dark-state J–V testing was performed. The devices based on STAB exhibited lower dark current (Fig. 5f, ESI†) compared to the control group. The saturated dark current density J0 decreased by approximately one order of magnitude (from 1.18 × 10−5 mA cm−2 to 1.03 × 10−6 mA cm−2), indicating that shallow-level defects were significantly suppressed.58 Next, transient photovoltage (TPV) and transient photocurrent (TPC) were used to investigate the recombination dynamics of the devices (Fig. 5g and h). After STAB treatment, the photovoltage lifetime increased significantly (from 794.64 ns to 1020.28 ns), which demonstrates good agreement with the improved photovoltaic performance due to defect passivation.59 Meanwhile, the photocurrent lifetime decreased significantly (from 499.13 ns to 103.28 ns), indicating faster carrier extraction in the STAB-treated devices.60 Additionally, the dependence of VOC on light intensity was tested in the range of 10–100 mW cm−2. The light intensity (I) and VOC follow the following equation:
The stability of the devices was evaluated under different settings. First, the unpackaged devices were placed on a hot plate at 85 °C in the dark to study their thermal stability. As shown in Fig. 6a, the performance of the control device deteriorated rapidly, with the PCE decreasing to 61.98% of its initial value after 1000 hours. In contrast, the STAB-treated device retained 92.31% of its initial PCE after 1000 hours, showing significantly improved thermal stability. Fig. 6c displays the cross-sectional SEM images of the control and STAB-treated devices after heat aging at 85 °C for 10 days. Many voids can be seen in the control device, which could be attributed to the loss of components in the perovskite film during heat aging.20 In contrast, the STAB-treated device remained more intact than the control films, which is consistent with the efficiency trend. Moreover, the maximum power point tracking (MPPT) test was performed under continuous one sun illumination in N2 at 65 ± 5 °C to estimate the photostability of the devices. As shown in Fig. 6b, the STAB device retained 90.21% of its initial efficiency after 1000 hours of MPPT, while the control device retained only 70.52% of its initial efficiency after 600 hours, indicating a significant improvement in the photostability of the STAB device. Additionally, the humidity stability of the device was enhanced, owing to a more hydrophobic surface (Fig. S17 and S18†). The stability gain of the device under light and heat conditions comes from STAB inhibiting the generation of VI and hindering the conversion of I− to I2, thereby preventing the degradation of the perovskite film. Furthermore, the 1.27 eV Sn–Pb solar cells were also subjected to 85 °C thermal aging, and it was found that STAB significantly enhanced the thermal stability of the 1.27 eV Sn–Pb device (Fig. 6d). The STAB-treated 1.27 eV devices retained 81.2% of its initial efficiency after 600 hours of thermal aging, while the control device only retained 41.9%. Overall, these results highlight the universal enhancement of the stability of inverted PSCs by the STAB post-treatment method due to the management of iodine vacancy defects and reduction ability towards I2.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5ta04007h |
This journal is © The Royal Society of Chemistry 2025 |