Aman Shukla‡
ab,
Dhruba B. Khadka‡
*a,
Chunqing Li
c,
Masahiro Rikukawa
c,
Yuko Takeoka
*c,
Ryoji Sahara
*d,
Masatoshi Yanagida
a and
Yasuhiro Shirai
*a
aPhotovoltaic Materials Group, Center for Green Research on Energy and Environmental Materials, National Institute for Materials Science (NIMS), 1-1 Namiki, Tsukuba, Ibaraki 305-0044, Japan. E-mail: KHADKA.B.Dhruba@nims.go.jp; SHIRAI.Yasuhiro@nims.go.jp
bDepartment of Materials Science & Engineering, National Centre for Flexible Electronics, Indian Institute of Technology Kanpur, Kanpur, Uttar Pradesh 208016, India
cDepartment of Materials and Life Sciences, Sophia University, 7-1, Kioi-cho, Chiyoda-ku, Tokyo 102-8554, Japan. E-mail: y-tabuch@sophia.ac.jp
dComputational Structural Materials Group, Research Center for Structural Materials, National Institute for Materials Science (NIMS), 1-2-1, Sengen, Tsukuba, Ibaraki 305-0047, Japan. E-mail: SAHARA.Ryoji@nims.go.jp
First published on 7th May 2025
Tin-based perovskite solar cells (Sn-PSCs) represent a promising lead-free alternative for photovoltaic applications, however, their oxidation of Sn2+ to Sn4+, induces structural defects and compromises device stability and efficiency. In this study, we introduced fullerene-based multifunctional molecules (F-COOH, F-OH, F-OSO3H) as additives to interact with Sn2+ ions, effectively stabilizing tin in its reduced state. These functional additives affect the growth and optoelectronic properties of tin perovskite film. Among these additives, F-COOH significantly suppresses Sn4+ formation and non-radiative recombination. Consequently, the device with the F-COOH additive exhibits an increased power conversion efficiency (PCE) from 8.20 to 11.22%, along with improved reproducibility and stability. While additives with –OH and –OSO3H functional groups also enhance performance, the superior results with F-COOH are attributed to the localized electron density provided by the –COOH group, facilitated by its connection to the fullerene core through a sp3 hybridized carbon. Device analysis indicated that the F-COOH additive enhances the optoelectronic properties of Sn-PSCs, contributing to a higher diffusion potential while effectively minimizing bulk and interfacial defects. Thus, this work underscores the importance of functional group selection in molecular design to improve the efficiency and stability of Sn-PSCs, paving the way for advanced lead-free solar cell technologies.
The functional additive technique has emerged as a critical approach to enhancing the performance and stability of Sn-PSCs.2,8–11 Reducing agents like SnF2, trivalent doping, hydrazine, or NaBH4 minimize Sn4+ defects by stabilizing the Sn2+ oxidation state.12,13 Additives such as alkylammonium halides, polyethylene glycol, and 2-chloroethylphosphonic acid improve film morphology and crystallization, yielding larger grains and fewer defects, which enhance charge transport.14–16 Solvent additives like DMSO or ethyl acetate control evaporation, ensuring uniform films.17,18 Functional additives also optimize interfacial properties by tuning energy levels, reducing recombination, and enhancing charge extraction.19–21 For example, self-assembled monolayers with specific dipole moments can be applied to tune the work function of the electron transport layer, optimizing the charge extraction process.22 Moreover, hydrophobic additives further improve stability by protecting the perovskite layer from moisture-induced degradation.23
In recent years, fullerene derivatives have come up as particularly effective additives for this purpose.24–28 Their ability to interact with the perovskite material at a molecular level makes them especially suited for the passivation of grain boundaries and surface defects in the film. This passivation reduces non-radiative recombination and enhances charge carrier mobility, resulting in higher photocurrent and PCE.29–31 For instance, Tian et al. reported an enhancement in the efficiency of tin-based perovskite solar cells by incorporating a hexyl ester-containing fullerene derivative as a functional additive.31 This improvement was attributed to the suppression of Sn2+ oxidation by the flexible alkyl chains in the additive, which prevent the perovskite layer from interacting with the oxygen. Similarly, Chen and colleagues demonstrated that the quality of the perovskite layer could be enhanced by forming a bulk heterojunction between the –R–NH2–X group of a fullerene derivative and the perovskite molecules.32 They also showed that a fullerene derivative with six chlorine atoms could address grain boundary defects by slowing down the crystallization process of the perovskite layer, thereby improving device efficiency. In another study, Liang and colleagues designed a novel fullerene derivative with a porphyrin ring and three pentafluorophenyl groups.33 This innovative additive efficiently interacts with the perovskite material, facilitating defect passivation and significantly extending the device's lifespan. Choi et al. introduced a multifunctional fulleropyrrolidine with triethylene glycol monoethyl ether chains, where the ether component closely interacts with Sn2+, and the fullerene base simultaneously engages with I−.34 This dual interaction prevents the formation of Sn4+ and I3−, resulting in enhanced stability of Sn-based solar cells. Most recently, Chen et al. reported a record-breaking efficiency of 15.14% by using pyridyl-substituted fulleropyrrolidones as functional additives in the perovskite precursor solution.30 This milestone highlights the significant potential of fullerene derivatives in advancing the performance and stability of perovskite solar cells, paving the way for more efficient and durable renewable energy solutions.
Herein, we explored the impact of fullerene derivatives with different functional groups (–COOH, –OH, –OSO3H) on the efficiency of Sn-PSCs. This study aimed to understand how these functional groups with lone pair-bearing oxygen atoms, attached to the bulky fullerene base, interact with the perovskite matrix to influence key parameters such as film morphology, crystallinity, and the oxidation state of tin. It was found that the derivative with a carboxylic group (–COOH) exhibited the most significant enhancement in device performance from 8.20 to 11.22%. These fullerene-based additives with the carboxylic group were found to be capable of moderating the crystallization process of the perovskite film, resulting in a more uniform morphology with fewer defects. The detailed materials and device analysis corroborate that these functional additives effectively suppressed the oxidation of Sn2+ to Sn4+ and the recombination states in Sn-PSCs. This work provides valuable insights into the effect of multifunctional functional groups in Sn-PSCs and their crucial role in improving device performance and stability.
To understand the effect of additives on crystal growth, X-ray diffraction (XRD) patterns of Sn-HP films (Fig. 1c) were measured. The most prominent XRD peaks obtained correspond to (100) and (200) crystallographic planes, which are consistent with the orthorhombic phase, aligning with previously reported data.36 A slightly narrower FWHM of the dominant XRD pattern suggests improvement in crystallinity with Sn-HP film with additives (Fig. S6†). Furthermore, the characteristic XRD peaks of the perovskite layer with different additives do not shift from that of the control layer, suggesting no incorporation of fullerene-functional additives into the lattice of the host crystal.
Similarly, the absorption and PL spectra of Sn-HP with additive films were measured to evaluate the effect of additives on photophysical properties. Fig. 1d exhibits a nuanced impact on characteristic absorption spectra with slight variations in absorbance. Among the different additives, the film with the F-COOH additive demonstrates a slightly higher absorption response, indicating a better optoelectronic response. The characteristic absorption edge, as depicted in the inset, shows a band edge of ∼892 nm. PL spectra of these films are depicted in Fig. 1e. There is no shifting in characteristics PL peak ∼892 nm, which is equivalent to ∼1.395 ± 0.02 eV corresponding to the band edge, indicates non-interference of these additives into the electronic picture of pristine perovskite structure. Importantly, a variation in PL spectra intensity indicates the effect of additives in a passivating defect in the Sn-HP film. An intensified PL spectrum for Sn-HP with F-COOH additive corroborates improved film quality and reduced non-radiative recombination.37 Further, PL spectra of Sn-HP films with varying concentrations of F-COOH additive (Fig. S7†) suggest that an excessive additive concentration induces a nonradiative recombination state within the perovskite films.38,39 With higher additive concentrations, additional defect states or traps in the perovskite film promote non-radiative recombination, where the energy from excited carriers is lost as heat rather than emitted as light, thus diminishing the PL intensity.
Fig. 1f1–f4 display the SEM images of the Sn-HP films, providing insights into how different functional additives influence the film morphology, particularly in terms of surface coverage and defect density. The SEM image of the control film (Fig. 1f1) shows pinholes with poor film coverage, which are detrimental to the overall performance of the perovskite layer.40 While the SEM images (Fig. 1f2–f4) of Sn-HP with F-COOH, F-OH, and F-OSO3H additives demonstrate significant improvements in surface coverage and film uniformity. These additives play a crucial role in enhancing the quality of the perovskite film by reducing the number of pinholes and defects. The presence of fewer defects and more complete surface coverage leads to better charge transport, ultimately improving the optoelectronic properties of the films. Among the additives, F-COOH shows comparatively better film morphology, yielding a film with the most uniform coverage that could be due to the strong interactions between the functional carboxyl (–COOH) groups of the F-COOH additive and the tin perovskite polyhedral. These interactions play a key role in regulating the crystallization process during film formation, promoting more controlled and uniform growth of perovskite crystals. However, the amount of additives also plays a critical role in determining the film quality. SEM images of films with higher concentrations of F-COOH reveal irregularities (Fig. S8†). This suggests that beyond the optimal concentration, the additive begins to interfere with the perovskite crystallization process, possibly by introducing excess nucleation sites or disrupting the film's uniform growth. Consequently, the film becomes more prone to imperfections, undermining the benefits initially provided by the additive.
Device | Scan direction | JSC (mA cm−2) | VOC (V) | FF (%) | PCE (%) | Average (PCE ± SD) |
---|---|---|---|---|---|---|
Control | F | 16.47 | 0.783 | 59.66 | 7.69 | 7.76 ± 0.52 |
R | 17.44 | 0.769 | 61.29 | 8.20 | ||
F-COOH | F | 19.27 | 0.838 | 65.51 | 10.57 | 10.32 ± 0.51 |
R | 19.31 | 0.841 | 69.12 | 11.22 | ||
F-OH | F | 18.48 | 0.819 | 64.59 | 9.77 | 9.38 ± 0.46 |
R | 18.67 | 0.825 | 66.22 | 10.19 | ||
F-OSO3H | F | 17.85 | 0.801 | 65.51 | 9.36 | 8.97 ± 0.47 |
R | 18.36 | 0.813 | 65.42 | 9.76 |
Furthermore, Sn-PSCs with additives containing hydroxyl (–OH) and sulfonic ester (–OSO3H) groups also exhibited an increase in PCE, reaching 10.19% and 9.76%, respectively. These results indicate that both –OH groups and –OSO3H also contribute to improving the optoelectronic properties of the perovskite film, likely through enhanced molecular interaction and better crystallization processes.41,42 These improvements suggest that the additive effectively stabilizes the interface between the perovskite and the charge transport layers and mitigates ion migration and trap-assisted recombination.38,43 We will discuss this in detail in succeeding sections.
Fig. 2d shows the external quantum efficiency (EQE) of the control device and the best Sn-PSC with F-COOH additive. The EQE spectrum for Sn-PSC with F-COOH additive reveals a noticeable enhancement across the entire spectral range, indicating the suppression of recombination activities in bulk and at the interfaces of the device.44 The JSC values (15.67 mA cm−2 for control and 17.50 mA cm−2 for the device with F-COOH) obtained from integrating the EQE spectrum are in the close range of those obtained from J–V curves. Additionally, the band edge of the EQE spectra (Fig. S11†) is estimated to be ∼1.428 and 1.425 ± 0.02 eV for control and with F-COOH additive. These values are in close agreement with the band edge estimation from absorption and PL spectra.
Moreover, the statistical data (Fig. 1e), provided in Table S4† and graphically illustrated in Fig. S12,† offer a detailed comparison of batches of Sn-PSCs with additives. The control device demonstrates an average PCE of 7.76%, while those values for devices with additives are higher. Importantly, the statistical data reveal that the distribution of device parameters across batches has narrowed for devices with additives compared to the control device, suggesting higher reproducibility of device parameters and, hence device performance. Enhanced reproducibility is critical for ensuring that solar cells perform consistently, which is a key requirement for large-scale manufacturing and commercialization.
To study the effect on device stability, we collected the operational stability of unencapsulated control and device with F-COOH additive under maximum power point tracking (MPPT) conditions and air ambient, adopting stability assessment ISOS-L-1.45 The normalized efficiencies of respective Sn-PSCs are presented in Fig. 2f. The control device experiences a significant drop in efficiency to 35% of its initial value after 500 hours. This rapid decline in performance suggests that the control device is highly susceptible to degradation, likely due to factors such as moisture, oxygen ingress, or intrinsic instability within the perovskite layer.46 In contrast, the device incorporating the F-COOH additive exhibits markedly improved stability, retaining more than 51% of its initial efficiency even after 500 hours. This enhanced stability highlights the beneficial effect of the F-COOH additive in mitigating the degradation mechanisms commonly observed in perovskite solar cells.46,47 It is likely attributed to the additive's ability to interact with the perovskite structure during film formation, leading to better crystallization, reduced defect density, and stronger resistance to environmental factors. The water contact angles of the Sn-HP film with fullerene derivatives (Fig. S15†) show a higher water contact angle compared to the control film, suggesting a higher hydrophobicity film. The increase in hydrophobicity with fullerene additive also supports the superior device stability of Sn-PSC with F-COOH additive. This improvement implies that the F-COOH additive appears to play a protective role by passivating surface defects, stabilizing grain boundaries, and inducing water resistivity, which are often points of vulnerability in perovskite films where degradation initiates. It reduces degradation over time and contributes to a more stable and durable solar cell. However, further research is warranted to explore additional methods to enhance stability even more.
![]() | ||
Fig. 3 XPS spectra: (a and b) Sn 3d, (c and d) I 3d of the surface of the Sn-HP films without and with F-COOH additive. |
In fundamental chemical aspects, the ability of these additives to mitigate Sn2+ oxidation can be attributed to their inherent nature and interactions with the Sn2+ ion.37 Functional groups capable of donating electron density or coordinating with tin atoms play a crucial role in stabilizing Sn2+. Particularly, the carboxylate group in F-COOH can bind to tin via coordination, offering a protective barrier that helps to maintain a higher proportion of Sn2+. Importantly, the –COOH group in F-COOH is not directly attached to the fullerene ring but instead is connected through a sp3-hybridized carbon atom (C60(C(COOH)2)n). This structural feature localizes the electron density on the carboxylate group, making it more available for interaction with Sn2+, contributing to the improved stability and quality of the resulting perovskite film. On the other hand, the hydroxyl group (–OH) in F-OH also contributes electron density to the fullerene ring through hydrogen bonding, which can help stabilize Sn2+. Although the electron-donating power of the hydroxyl group is generally stronger than that of the carboxylate group, in this specific molecular framework, the –OH group is directly attached to an sp2-hybridized carbon of the fullerene ring, which is more acidic than sp3 hybridized carbon.49 This configuration results in higher acidity of the carbon and partial delocalization of the electron density over the fullerene ring, reducing the availability of electron density for Sn2+ stabilization as compared to F-COOH. The sulfate ester group (–OSO3H) in F-OSO3H also provides some stabilization of Sn2+ through conjugation, but the electron-withdrawing nature of the –SO3H group limits the electron donation to the tin ion. As a result, F-OSO3H is less effective in stabilizing Sn2+ compared to F-COOH or F-OH. The hierarchy of electron-donating ability and defect passivation effectiveness among the derivatives follows the trend: F-COOH > F-OH > F-OSO3H, in agreement with prior experimental observations and device performance trends. This observation is in line with a report on the effect of the interaction of HCOO− anions and Sn2+ cations by Wang and co-workers.50 To validate this hypothesis further, these fullerene derivatives were investigated through DFT-calculated electrostatic surface potential (ESP) analysis. The ESP mapping (Fig. S16a–d†) reveals that F-COOH exhibits significant negative potential around the –COOH group, facilitating effective coordination with undercoordinated Sn2+ sites. In contrast, the ESP maps of F-OH & F-OSO3H display relatively less negative potentials, reflecting weaker passivation capability compared to F-COOH. Hence, the superior performance of F-COOH arises from its optimal molecular structure that facilitates efficient electron donation and robust Sn2+ stabilization, as confirmed by ESP analysis. These findings validate the proposed molecular interaction mechanism and underscore the critical importance of sp3-hybridized anchoring points and carefully modulated electrostatic environments in engineering highly efficient and stable tin-based perovskite solar cells.
![]() | ||
Fig. 4 (a) TRPL decay spectra of Sn-HP films (control and with F-COOH additive). (b) TPV decay curves, and (c) TPC decay curves of devices (control and with F-COOH additive). |
Moreover, to gain insights into the photocarrier dynamics, we conducted transient photo characteristics, transient photovoltage (TPV), and transient photocurrent (TPC) measurements. The TPV curve (Fig. 4b) for the F-COOH device exhibits a notably slower decay compared to the control device, indicating a significantly longer carrier lifetime of 15.44 μs from 10.62 μs. This extended carrier lifetime is typically a sign of fewer trap states within the device, which would otherwise act as recombination centers for carriers.51,52 It corroborates that Sn-PSC with F-COOH additive reduces trap-assisted recombination and contributes to enhanced performance, supporting the conclusion that this functional additive is key to achieving better device stability and efficiency. Similarly, the TPC curve (Fig. 4c) assesses the effect on the interface quality of the device. Sn-PSC with F-COOH additive shows a faster current decay time of 5.56 μs, compared to the control (6.38 μs). This faster decay is indicative of more efficient charge carrier extraction in the Sn-PSC with F-COOH, one of the critical factors for improving solar cell performance. These results further support the benign effect of F-COOH additives on material quality and device properties. Furthermore, the capacitance measurements were conducted to gain critical insights into the device's charge carrier dynamics, interface states, trap states, and defect density profile.53,54 Fig. 5a shows the capacitance–frequency (C–f) spectra of devices. The control device exhibits a noticeably higher capacitance compared to the F-COOH-modified device, which is attributed to higher charge accumulation and ionic motion in the control device. This suggests that the additive might be suppressing ionic movement or mitigating its effects on charge accumulation.55 The device with the F-COOH additive demonstrates a lower capacitance over the frequency range, indicating that the additive has likely reduced the density of trap states.43,46
![]() | ||
Fig. 5 Capacitance characteristics of devices: (a) C–f spectra under darkness, (b) Mott–Schottky plots, (c) carrier profile of the device without and with F-COOH additive. |
To evaluate the effects on the defect profile, we analyzed the capacitance spectra with Mott–Schottky (M–S) plots and carrier profile as given by where NCV represents carrier density calculated from the capacitance–voltage (C–V) curve, C is the capacitance per unit area, ε0 is the permittivity of free space, εs is the dielectric constant of the perovskite material. The MS plot, as shown in Fig. 5b, compares the control device with the device containing the F-COOH additive. The device with the F-COOH additive reveals a higher diffusion potential (VD) of 0.824 V compared to the control (0.702 V). This increment in VD aligns well with the enhancement in VOC in Sn-PSC. This result suggests that the F-COOH additive strengthens the separation of electron–hole pairs, reduces recombination losses, and ultimately contributes to higher VOC.51
Fig. 5c depicts the spatial distribution of charge carriers across the device calculated from the C–V measurements. This analysis provides insights into how charge carriers are distributed within the bulk of the active layer and at the interface regions, where recombination and charge transport play a crucial role in device performance. It has been documented that the NCV profile accounts for the carrier distribution (free carrier and defect density)47,56,57 and ion or charge accumulation at the interface53 in thin-film solar cells. In the control device, the bulk carrier density (NBCV) is estimated to be 3.18 × 1015 cm−3. While the device with the F-COOH additive showed a reduced bulk carrier density of 1.47 × 1015 cm−3. This reduction suggests that the F-COOH additive is effective in mitigating defects within the bulk perovskite layer that can capture and recombine charge carriers. This trend of reduced carrier profile extends to the interface region as well. The control device exhibits an interfacial carrier density (NIFCV) of 7.42 × 1016 cm−3, while the F-COOH-modified device shows a significantly lower density of 1.65 × 1016 cm−3. This reduction at the interface implies that the F-COOH additive is also effective at passivating interfacial defects, which are often hotspots for charge recombination due to discontinuities in the crystal structure and imperfect layer alignment. The decrease in the C–V carrier profile correlates with improved carrier lifetimes resulting from defect passivation.
Theoretical insights were obtained by performing DFT calculations considering a slab model with a SnI2-terminated perovskite surface as described in our earlier report58 to investigate the interaction between fullerene derivatives and the tin-based perovskite. The charge density difference of fullerene functional derivatives on the defective Sn-perovskite surface (Fig. 6a–c) indicates mitigation of the density of defect states.37 The adsorption energies (Fig. 6a–c and S16d–f†) reveal a clear trend: F-COOH exhibits the strongest binding with the Sn-perovskite surface (−0.555 eV), followed by F-OH (−0.408 eV) and F-OSO3H (−0.267 eV). The stronger adsorption energy of F-COOH suggests a more robust chemical interaction with undercoordinated Sn2+ sites, leading to more effective defect passivation compared to F-OH and F-OSO3H. This inference is further supported by the density of states (DOS) calculation as depicted in Fig. 6d–g, which shows a reduction in defect states near the Fermi level up to some extent upon fullerene functionalization, particularly in the case of F-COOH. The stronger binding and superior defect mitigation effect of F-COOH correlate well with experimental observations, including reduced trap-assisted recombination from capacitance measurements, longer carrier lifetimes from TRPL analysis, and improved photovoltage stability from TPV measurements.59
Thus, theoretical and experimental results reveal that the fullerene-based functional additives induce a strong adsorption on the Sn-perovskite surface. This interaction plays a crucial role in mitigating the oxidation of Sn2+ to Sn4+ and enhancing the perovskite layer's material chemistry integrity. The fullerene derivatives demonstrate a notable improvement in both efficiency and stability of the devices. A comprehensive device analysis combined with theoretical insights substantiates the experimentally observed performance enhancements, highlighting the potential of fullerene derivatives as effective functional additives for advancing Sn-based perovskite photovoltaics.
For the exchange–correlation function, the Perdew–Burke–Ernzerhof function,38 was used. A 2√2 × 2√2 × 1 slab supercell of (001) surface, containing 5 layers, was built from a bulk tetragonal phase of FASnI3 (space group: P4/mbm), with a vacuum region of about 22 Å was added in the z direction. The kinetic energy cutoff of 400 eV and the convergence criterion of 10−4 eV for the self-consistent loop were employed. To explore stable adsorption sites of the molecule, a SnI2-terminated surface with Sn-vacancy (VSn) defect was used, on which a molecule was placed, based on the insight from the previous work.62 Gamma point sampling was employed for the Brillouin zone integration. The adsorption energy of the molecule was evaluated as Eads = Esystem with molecule − Esystem without molecule − μmol where Esystem with molecule and Esystem without molecule are energies of the surfaces with and without a molecule additive, respectively, and μmol is the chemical potential of the molecule. The total energy computed for an isolated gas phase was used for μmol.
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4ta08566c |
‡ These authors contributed equally to this work. |
This journal is © The Royal Society of Chemistry 2025 |