Divalent cation replacement strategy stabilizes wide-bandgap perovskite for Cu(In,Ga)Se2 tandem solar cells
  • SJ_Zhang
  • Apr. 19, 2025

Abstract

Despite improvements in the power conversion efficiency (PCE) of perovskite solar cells (PSCs), stability issues due to ion migration and phase separation remain critical concerns. Given the ionic crystal nature of perovskites, the use of multivalent cations is supposed to effectively suppress ionic migration. However, multivalent metal cations produce deep-level trap states, thus impairing device efficiency. Therefore, a multivalent cation replacement strategy that minimizes interstitial defects is desirable. Here we develop a divalent cation replacement strategy that mitigates ionic migration while limiting phase segregation. We demonstrate that the replacement of the A-site cations in the perovskite lattice with methylenediammonium cations (MDA2+) substantially suppresses the above issues in wide-bandgap perovskites. This is mainly due to the bivalent state of MDA2+ generating a strong interaction with the inorganic framework and reducing the mobility of halide ions and the formation of defects. As a result, the stability and efficiency of the fabricated PSCs are substantially improved. We demonstrate a champion PCE of 23.20% (certified 22.71%) for a single-junction PSC with a bandgap between 1.67 eV and 1.68 eV. Furthermore, a PCE of 30.13% is obtained for mechanically stacked perovskite/Cu(In,Ga)Se2 tandem devices, and a PCE of 21.88% for translucent perovskite devices. Finally, we obtain a steady-state PCE of 23.28% (certified 22.79%) for flexible monolithic perovskite/Cu(In,Ga)Se2 tandem cells.

Main

Halide perovskites are attractive materials for photovoltaics owing to their exceptional optoelectronic properties1,2,3,4,5. Perovskite-based tandem solar cells have attracted increasingly more research attention because of their potential to achieve higher power conversion efficiencies (PCEs) than their single-junction counterparts6,7,8,9. The construction of an efficient tandem cell requires the use of a wide-bandgap perovskite material as the light-absorbing layer of the top subcell, generally enabled by the inclusion of more bromide. However, the concomitant instability issue, especially phase separation, has become a major barrier to the commercialization of the perovskite-based tandem photovoltaics10,11,12.

A variety of methods have been reported to enhance the longevity of perovskites and suppress ion migration and phase segregation, including crystallographic modulation12, formation of two-dimensional/three-dimensional heterojunctions13,14, polymer modification15, inorganic layer coating16, utilization of ionic liquids17, chlorine alloying18, inclusion of pseudo-halide anion19,20, interstitial doping by multivalent alkali metal cation21 and so on. Given the ionic migration in this soft ionic crystal, the use of multivalent cations is supposed to be very effective owing to the enhanced electrostatic interactions. However, multivalent metal cations in perovskites, as interstitial defects due to small ionic radii, generally produce deep-level trap states, thus impairing the device efficiency. Although trace amounts may minimize the sacrificial trade-off21, precise control of the dosage is challenging and, thus, not conducive to industrial implementation. Therefore, a multivalent cation replacement strategy that involves entering the perovskite lattice without forming interstitial defects may be desirable.

Here we report the partial replacement of the A-site in the perovskite lattice by using methylenediammonium cations (MDA2+) to inhibit halide migration in wide-bandgap perovskites. We show that MDA2+ bears the strongest interactions with the corner-sharing octahedral sublattice, compared with the univalent methylammonium (MA+), formamidinium (FA+) and guanidinium (GA+) cationic species. This is primarily due to the divalent state of MDA2+, enabling a stronger electrostatic interaction with the inorganic framework. Therefore, the incorporation of MDA2+ significantly increases the energy barrier of ionic movement and, thus, inhibits ion migration and photoinduced halide segregation. The as-resulted wide-bandgap single-junction perovskite solar cells (PSCs) obtained a champion PCE of 23.20% with a bandgap of 1.67–1.68 eV (certified 22.71% efficiency). The unencapsulated target cells maintained approximately 80% of their initial efficiency at maximum power point (MPP) tracking for 750 h at around 45 °C, whereas the control device lost 50% after 700 h. Semitransparent PSCs were fabricated, delivering a champion PCE of 21.88%. We applied them to four-terminal perovskite/Cu(In,Ga)Se2 (CIGS) tandem devices and achieved a high PCE of 30.13%. An efficiency of 23.41% obtained from current density–voltage (JV) sweep with a steady-state efficiency of 23.28% (a certified efficiency of 22.79%) was also demonstrated on flexible monolithic perovskite/CIGS tandem solar cells.

Impedance of ion migration in wide-bandgap perovskite

As the simplest and smallest diammonium cation, MDA2+ was previously reported to be used in the FAPbl3 perovskite to inhibit the transition of α-FAPbl3 to its δ-phase22,23. However, the potential of MDA2+ in other perovskite systems has remained largely unexplored. MDA2+ features a high charge number, rich –N–H bonds and an acceptable ionic size, which may allow substitutional doping at the A-site and the formation of stronger interactions with the inorganic sublattices over MA+ and FA+, thus holding the potential to suppress ion migration and photoinduced halide segregation in wide-bandgap perovskites. We compared the energy barriers of ion migration among the cases of MA+, FA+, GA+ and MDA2+ (see chemical structures and ionic radii in Supplementary Fig. 1) cationic dopants using density functional theory (DFT) calculations based on a MAFAPbIxBr3−x perovskite system. Figure 1a depicts a schematic illustration of the substitution of various organic cations at the A-site within the perovskite lattice. Figure 1b shows the most preferred migration path of iodide ion. Based on this path, we compared the energy barriers of migration in the case of four types of organic A-cation (Fig. 1c). The energy barrier for ion migration was significantly increased upon MDA2+ replacement as compared with the other three cationic substitution cases. Figure 1d showcases the crystal model for DFT calculations. The detailed manner of interactions between the four types of organic cation and the lead–halide frameworks are shown in Fig. 1e. Among them, the hydrogen on the MDA2+ has the closest interaction distance with the halogen atom, suggesting a stronger bonding. The total binding energy with respect to each kind of cation is compared in Fig. 1f, and MDA2+ features the strongest interaction with the inorganic octahedron compared with its counterparts (−6.5, −6.3, −6.7 and −8.4 eV for FA+, MA+, GA+ and MDA2+, respectively).

 

Fig. 1: Theoretical studies of the interactions between different organic A-cations with the perovskite.

figure 1

a, A diagram of the replacement of different organic cations (MA+, FA+, GA+ and MDA2+) at the A-site in the perovskite lattice based on a mixed-cation mixed-halide perovskite system. b, Theoretical models of the iodide ion migration pathway in the perovskite lattice. c, Relative energy landscapes of the system during the iodide ion migration. The x axis denotes the steps of an iodide ion hopping to the nearby iodine vacancy. The data points are the steps with the highest relative energy during the ion migration. d, The DFT model used to calculate the interaction energy of organic A-cations with the inorganic framework. e, DFT-modelled effective hydrogen bonds between the organic cations and perovskite lattice for FA+, MA+, GA+ and MDA2+, respectively. f, Interaction energies of the four types of organic A-cation with the inorganic cage.

We prepared perovskite thin films with MDA2+. The stability and solubility of MDA2+ in solvents are discussed in Supplementary Texts 1 and 224,25. Fourier transform infrared (FTIR) spectrometry and X-ray photoelectron spectrometry (XPS) were performed to study the interactions of MDA2+ with the perovskites. The absorption peaks ascribed to the stretching vibration mode and bending vibration mode of the C–N and C–H bonds of MDA2+ shift from 1,223 cm−1 and 1,403 cm−1 to 1,209 cm−1 and 1,409 cm−1, respectively, indicating the interactions between MDA2+ and the perovskite (Supplementary Fig. 6). The electron orbitals of Pb-4f, I-3d and Br-3d all shift significantly towards lower binding energies according to the XPS spectra, which also indicates strong interactions between MDA2+ and the inorganic framework (Supplementary Fig. 7a–c). The XPS signal of the Cl was also acquired in the target film, which was negligible in the control sample (Supplementary Fig. 7d). Time-of-flight secondary ion mass spectroscopy (ToF-SIMS) unveiled that the amount of Cl increased significantly in the perovskite film after addition of MDACl2 and mainly distributed in the upper domain of the thin layer (Supplementary Fig. 8). Given the divalent state of MDA2+, the introduced Cl may exist at both the X-sites and the interstitial sites of the lattice to satisfy electrical neutrality. We calculated the energy of Cl at the interstitial position of the perovskite lattice with and without MDA2+ substitution (Supplementary Fig. 9). The interstitial formation energy is dramatically decreased upon MDA2+ replacement, supporting that MDA2+ could facilitate the insertion of Cl into the interstitial spaces, consistent with a previous report22. We also calculated the formation energy of interstitial Cl in a pure FAPbI3 lattice, without the inclusion of MDACl2, which is also much higher than in the case of MDA2+ replacement (Supplementary Fig. 10). The signals from FA+ and MDA2+ were hard to distinguish due to their similar chemical structures (Supplementary Fig. 11).

Suppressed halide segregation and migration

We conducted temperature-dependent conductivity measurements following previous reports to verify the suppressed ion migration in the MDA2+-incorporated perovskite film (Fig. 2a)16,26,27. The change of conductivity with temperature is determined by two regions. In the lower-temperature region, the conductivity is dominated by the conductance of electrons. At higher temperatures, the conductivity is instead dominated by ionic conductance. The transition of the two regions occurs at a higher temperature (245 K) for the MDA2+-doped sample compared with its undoped counterpart (192 K), indicating that ion migration is inhibited in the doped film. Furthermore, by linearly fitting the temperature-dependent conductivity data in the two regions to the Nernst–Einstein relation, we obtained the corresponding activation energies of electron transport and ion migration, respectively. The ion migration activation energy of MDA2+-doped perovskite films is 0.75 eV, much higher than that of undoped samples (0.43 eV). This is compelling evidence that MDA2+ cations stabilize halide perovskites and inhibit ionic movement.

Fig. 2: Suppressed ion migration and halide segregation enabled by MDA2+.

figure 2

a, Arrhenius plots of the conductivity of the undoped and MDA2+-doped wide-bandgap perovskite films, where the Ea is calculated from the slope of the curves. Data are presented as mean values ± s.d. Sample size n is defined as the number of conductivity values calculated from each it test under constant applied voltage, n = 30 for all cases. b,c, Time-dependent PL spectra of the 1.68-eV-bandgap perovskite absorbers without (b) or with (c) MDA2+ doping under light irradiation. d,e, In situ PL images before and after applying a constant bias voltage for the lateral devices based on the control film (d) and the MDA2+-doped film (e).

We also performed time-dependent in situ photoluminescence (PL) emission spectroscopy tests of the MDA2+-doped and control perovskite films under continuous photon radiation (Fig. 2b,c). The initial emission peaks of the control and target films are located at 736 nm and 734 nm, respectively. The blue-shifted emission of the MDA2+-doped film is attributed to the co-doped Cl, indicating that the perovskite bandgap is slightly increased with inclusion of 3 mol% MDACl2. This is in line with the X-ray diffraction (XRD) and ultraviolet (UV)–visible results presented in Supplementary Text 3. The PL emission peak of the undoped film suffered a dramatic redshift from the initial 736 nm to 781 nm and then remained stable (Fig. 2b), suggesting a severe photoinduced phase separation. By stark contrast, the PL spectra of the MDA2+-doped perovskite preserved almost invariability during an equal period of time (Fig. 2c), indicating a substantially suppressed halide segregation. Cl can also suppress halide separation18. To rule out the influence of Cl, we simultaneously examined the time-dependent PL spectrum evolution behaviour of the perovskite film with MDAI2 incorporation (3 mol%). The PL emission spectra of MDAI2-modified perovskite were unchanged (Supplementary Fig. 17), thus suggesting that the markedly enhanced photostability of the perovskite was mostly attributable to the presence of MDA2+ rather than Cl. To figure out the effect of FA+ and GA+ on photoinduced phase separation of the perovskites, we also compared the PL spectrum evolution over time of the perovskite thin films (Supplementary Fig. 18). The results indicated that the films were less stable than those incorporating MDA2+, potentially due to weaker interactions with the perovskite as demonstrated by the theoretical calculation result (Fig. 1f). For further discussion on the photo- and thermostability of the perovskites over doping amounts, see Supplementary Text 4.

To gain insight into the stability of perovskite films under electric field, confocal fluorescence microscopy was used to record the changes of their PL maps between two adjacent gold electrodes in real time. The control film underwent significant applied potential induced changes as revealed by the in situ PL images within 6 min (Fig. 2d). This appeared to be the result of a directional migration of ions driven by the electrical field. By contrast, we did not observe a similar phenomenon in the MDA2+-doped sample, which instead held steady over the same period of time (Fig. 2e). This result also convincingly demonstrated a stabilized halide perovskite with restrained ionic migration in the presence of MDA2+.

Improved optoelectronic properties

To gain insights into how the A-site substitution by MDA2+ would impact the defects in halide perovskite, we carried out DFT calculations to compare the formation energies of several types of point defect in the presence of MA+, FA+ and MDA2+ at the A-site, respectively (Fig. 3a and Supplementary Fig. 23). The defect formation energies of Pb–I antisite, I–FA antisite and iodide vacancy defects show the most notable increase with MDA2+ as compared with MA+ and FA+, suggesting that these types of imperfection in perovskite would be inhibited. Pb–I and I–FA antisite defects are deep-level defects that behave as non-radiative recombination centres28. The iodide vacancy defects are characterized by low formation energy that induce only shallow trap states, which, however, help to ion migration because vacancy-mediated halide migration is considered a main ionic transport mechanism in halide perovskites29.

Fig. 3: Improved optoelectronic properties of the perovskite films.

figure 3

a, Formation energies of several types of intrinsic point defect in the perovskite in the case of different organic cations at the A-site. b, PL spectra of the perovskite thin films prepared on glass substrates with or without MDA2+ inclusion. c, tDOS spectra of the control and target perovskite devices. d,e, Fluorescence lifetime maps of the control (d) and target (e) perovskite films on glass substrates.

Steady-state PL measurements were conducted on perovskite films on glass substrates (Fig. 3b). Stronger PL emission intensity of the MDA2+-doped sample suggests greater inhibition of non-radiative recombination channels in the perovskite compared with those in the control film due to lower defect state density. We measured the trap density of states (tDOS) with the energetic distribution in the devices based on the control and target perovskite films using thermal admittance spectroscopy (TAS) analysis (Fig. 3c). The target PSC has a lower tDOS in the deeper trap region (0.35–0.41 eV) compared with the reference one, demonstrating that MDA2+ doping reduces the deep-level defects. We used fluorescence lifetime imaging microscopy (FLIM) to visualize the difference in spatially resolved photocarrier lifetimes between the doped and undoped films. As shown in Fig. 3d,e, the photogenerated carrier lifetime was significantly longer in MDA2+-doped perovskites than in the control. Supplementary Fig. 24 provides a statistical distribution of carrier lifetimes for both samples. The lifetimes of the maximum number of charge carriers in the two perovskite layers are 77 ns and 114 ns, respectively. These results verified that MDA2+-doped perovskites have lower defect density and, thus, higher optoelectronic quality.

Photovoltaic performance and device stability

We prepared single-junction wide-bandgap (1.67–1.68 eV) PSCs with an inverted architecture (Fig. 4a). Figure 4b shows the JV curves of the champion PSCs without or with MDA2+ doping with an optimal doping concentration (see the detailed results of optimization in Supplementary Text 5). The champion PCE of the MDA2+-doped solar cell (0.75 mol% MDA2+) is 23.20% obtained by reverse voltage sweep with an open-circuit voltage (VOC) of 1.267 V, short-circuit current density (JSC) of 21.52 mA cm−2 and fill factor (FF) of 85.10%. It is one of the highest efficiencies of PSCs based on perovskite absorbers with 1.67–1.69 eV bandgaps (Supplementary Fig. 28 and Supplementary Table 4). The control devices yield a maximum efficiency of 22.22% with a VOC of 1.246 V, JSC of 21.26 mA cm−2 and FF of 83.89%. Reverse- and forward-scanned JV curves are shown in Supplementary Fig. 29. We sent it for certification (National Institute of Metrology), and a certified efficiency of 22.71% was obtained (Supplementary Fig. 30). Supplementary Fig. 31 shows the external quantum efficiency (EQE) spectra of the devices. A statistical comparison of photovoltaic parameters of the control and target devices is shown in Fig. 4c. Upon incorporation of MDA2+, the VOCJSC and FF of the PSCs are overall increased with a narrow distribution (Supplementary Table 5).

 

Fig. 4: Photovoltaic performance and device stability.

figure 4

a, The device configuration of the single-junction wide-bandgap perovskite photovoltaic cell (not to scale), where the SAM refers to self-assembled molecular layer. bJV curves of the champion PSCs with or without MDA2+ inclusion. c, Statistics of the photovoltaic metrics (PCE, VOCJSC and FF) of the control and target single-junction wide-bandgap PSCs (statistics derived from 18 and 26 devices of the undoped and MDA2+-doped group, respectively. whiskers: maxima and minima; bounds of box: 25th and 75th percentile; centre: mean). The blue and red data points and boxes represent the photovoltaic metrics of the undoped and MDA2+-doped PSCs, respectively. d, The MPP tracking stability of the unencapsulated control and target devices under simulated one-sun illumination in an N2 glovebox. e, A schematic illustration of the four-terminal perovskite/CIGS tandem solar cell (not to scale). fJV characteristics of the semitransparent perovskite device, stand-alone CIGS and filtered CIGS cells. gJV curve of a flexible two-terminal perovskite/CIGS tandem solar cell with MDA2+ incorporation.

To evaluate the stability of photovoltaic cells, we conducted MPP tracking and thermal stability tests. The unencapsulated solar cell with MDA2+ retained ~80% of its initial efficiency after output for 750 h at its MPP under one-sun illumination at around 45 °C. By contrast, the control device lost ~50% of its initial efficiency after 700 h of operation (Fig. 4d). The result demonstrates that A-site substitution by MDA2+ is capable of substantially improving the operational stability of the perovskite devices. In addition, we evaluated the stability of the PSCs at an elevated temperature. We continuously aged the unencapsulated devices in an N2 glovebox on a hotplate at 85 ± 1 °C and monitored their efficiency variations for about 1,200 h. An enhanced resistance to thermal stress of the perovskite photovoltaics by inclusion of MDA2+ was demonstrated (Supplementary Text 6).

We further prepared semitransparent wide-bandgap PSCs and achieved a champion PCE of 21.88% under reverse scan (see Supplementary Text 7 for results and discussions). We combined the high-performance semitransparent PSCs with CIGS photovoltaics to obtain mechanically stacked perovskite/CIGS tandem solar cells and measured their efficiencies. Figure 4e shows the configuration of the four-terminal perovskite/CIGS tandem device. When measuring the efficiency of the CIGS subcell, the semitransparent perovskite device without a silver electrode was placed onto the CIGS, behaving as a light filter. The gap between them was filled with a refractive index matching fluid or liquid to reduce the light reflection loss. We obtained an efficiency as high as 30.13% of the four-terminal perovskite/CIGS tandem device. The stand-alone CIGS solar cell with 120-nm-LiF antireflective coating has an efficiency of 18.63%, with a VOC of 0.676 V, JSC of 38.11 mA cm−2 and FF of 72.31% (see EQE spectra of the CIGS cells before and after antireflective coating in Supplementary Fig. 40). When the filter was applied, the CIGS device delivered a PCE of 8.25%, with a VOC of 0.643 V, JSC of 17.64 mA cm−2 and FF of 72.73%. The sum of the PCEs of the two subcells led to an efficiency of 30.13% for the four-terminal perovskite/CIGS tandem photovoltaics (Fig. 4f and Supplementary Table 7). We summarized the efficiencies of four-terminal perovskite/CIGS tandem solar cells reported in recent years (Supplementary Fig. 41)30,31. The integrated JSC of the two subcells from EQE spectra (Supplementary Fig. 42) are in line with their JV values.

Flexible monolithic perovskite/CIGS tandem solar cells are a type of promising photovoltaic technology characterized by flexibility, light weight and high efficiency. They bear advantages over perovskite/silicon tandem devices in flexibility and light weight, and advantages over perovskite/organic tandem devices and all perovskite tandem devices in stability. We therefore fabricated flexible two-terminal perovskite/CIGS tandem solar cells (see photograph of the tandem device in Supplementary Fig. 43). An efficiency of 23.41% obtained from JV sweep (Fig. 4g) with a steady-state efficiency of 23.28% after MPP output (VMPP = 1.38 V) for 10 min in ambient air (Supplementary Fig. 44) was demonstrated with MDACl2 inclusion. We obtained a certified efficiency of 22.79% under reverse scan and 21.73% under forward sweep (Supplementary Fig. 45), which is among the highest reported PCEs so far for flexible monolithic perovskite/CIGS tandem cells (Supplementary Fig. 46 and Supplementary Table 8)32,33. The EQE spectra of the tandem device are shown in Supplementary Fig. 47. The surface morphology of the perovskite layer on CIGS and the cross-sectional image of the tandem device are shown in Supplementary Fig. 48 and Supplementary Fig. 49, respectively.

Conclusions

We demonstrate a divalent cation replacement strategy, partially substituting the A-site using MDA2+ cations, capable of stabilizing wide-bandgap perovskites with intrinsically inhibited ion migration and photoinduced halide segregation. This is attributed primarily to the bivalent state of MDA2+, resulting in strong interactions with the inorganic sublattice and, thus, raising the energy barrier of ionic movement. In addition, the photoelectric properties of perovskite films were also improved due to the reduced defect density. Single-junction PSCs wih 1.67–1.68 eV bandgap afforded a champion efficiency of 23.20% (certified 22.71%) with enhanced operational stability. We also achieved 30.13% and 23.41% efficiencies on four-terminal perovskite/CIGS tandem and flexible monolithic perovskite/CIGS tandem photovoltaics, respectively. Overall, these represent state-of-the-art four-terminal and flexible monolithic perovskite/CIGS tandem photovoltaic technologies. Besides, they may inspire broader explorations of the potential of MDA2+ in the perovskite community and represent a pioneering work on multivalent cation replacement strategies to suppress ion migration in halide perovskites, thus opening new avenues for stabilizing perovskites for a wide range of optoelectronic applications beyond photovoltaics.

Methods

Materials

Lead iodide (PbI2), lead bromide (PbBr2), caesium iodide (CsI), nickel oxide particle (NiOx), lithium fluoride (LiF), bathocuproine (BCP) and C60 were purchased from Advanced Election Technology. Formamidinium iodide (FAI), methylammonium bromide (MABr), methylamine chloride (MACl), methylenediammonium dichloride (MDACl2), methylenediammonium diiodide (MDAI2), formamidinium chloride (FACl) and guanidinium chloride (GACl) were bought from Xi’an Polymer Light Technology; 4- trifluorophenylethylammoniumiodide (CF3-PEAI) was obtained from Xi’an E-Light New Material; and 4-(3,6-dimethyl-9H-carbazol-9-yl)butyl]phosphonic acid (Me-4PACz) was purchased from TCI. Anhydrous solvents involving dimethylformamide (DMF), dimethyl sulfoxide (DMSO), isopropanol (IPA), ethyl acetate (EA) and ethanol were purchased from Sigma-Aldrich.

Device fabrication

Opaque wide-bandgap PSCs

The prepatterned indium tin oxide (ITO) glass substrates were cleaned with a detergent in an ultrasonic bath and then washed several times by ultrapure water, followed by sonication in ethanol. The cleaned glass/ITO substrates were dried under a N2 flow and treated by UV–ozone for 20 min. The NiOx thin layers were prepared by spin-coating the NiOx/H2O dispersive liquid (10 mg ml−1) on the glass/ITO substrates at 3,000 rpm for 30 s and annealed at 150 °C for 10 min in ambient air. Then, they were transferred into an N2 glovebox. For Me-4PACz deposition, 0.6 mg ml−1 Me-4PACz in ethanol was spin-coated onto the NiOx layer at 3,000 rpm for 30 s, followed by heating at 110 °C for 10 min. To deposit wide-bandgap perovskite films, a pristine precursor solution (1.4 M) was prepared by mixing PbI2 (1.05 M, containing 5 mol% excess PbI2), PbBr2 (0.42 M), CsI (0.07 M), FAI (1.12 M), MABr (0.21 M) and MACl (0.14 M, as additive) in mixed solvents of DMF and DMSO with a volume ratio of 4:1 with incorporation of 1.5% H2O in volume ratio. The aim of adding H2O is to enable rapid dissolution of MDACl2 in the perovskite solution. For MDA2+ doping, MDACl2 powder was weighed separately in a new sample bottle. Then, a certain amount of completely dissolved pristine perovskite solution (vigorously shaken for 2–4 h) was transferred into the bottle that contained MDACl2 to ensure a specific molar ratio of MDACl2 relative to Pb. Because MDA2+ is relatively unstable at neutral pH, aprotic polar solvents such as DMSO, DMF and the as-prepared perovskite solution are typically shaken for a few minutes for use after introducing MDACl2 or MDAI2 into the well-prepared pristine precursor solution. A two-step rotation procedure of 1,000 rpm@10 s followed by 5,000 rpm@40 s was used to spin-coat the precursor solution. During spinning, 200 μl of EA was dropped onto the film 20–18 s before the end of the second step. The as-cast pristine films were annealed at 110 °C for 30 min in the glovebox. After the perovskite films cooled down, 15 μl of CF3-PEAI solution (1.5 mg ml−1 in IPA) was dynamically spin-coated onto the perovskite film at 6,000 rpm for 30 s, then annealed at 90 °C for 3 min. Subsequently, 1 nm of LiF (0.1 Å s−1), 25–40 nm of C60 (0.2 Å s−1) and 6 nm (0.1 Å s−1) of BCP were thermally deposited in sequence. Then, 120 nm (1.5 Å s−1) of silver electrodes were prepared by thermal evaporation with shadow masks, defining a working area of 0.107 cm2 for each device. Finally, 130 nm LiF (1 Å s−1) was evaporated as antireflective coating on the glass sides.

Semitransparent wide-bandgap PSCs

The fabrication processes for the semitransparent perovskite devices are essentially the same as described above, except that the thickness of the C60 layer was fixed at 25 nm and the BCP was replaced by SnO2 and indium zinc oxide (IZO). Besides, the silver electrode adopted a ring-type electrode with a thickness of 400 nm. Finally, 140 nm of LiF on the top side and 130 nm of LiF on the glass side were thermally evaporated. The preparation of the SnO2 and IZO layers is described as follows.

The SnO2 layers were prepared by atomic layer deposition using a Veeco system. The temperature of the substrates was maintained at 80 °C in the vacuum chamber. Tetrakis(dimethylamino)tin(IV) (TDMASn) was used as the Sn precursor by holding at 60 °C in a stainless-steel container during operation. Water was adopted as the oxidant without active heating. The precursor delivery manifold was heated to 150 °C. The TDMASn/wait/H2O/wait times were 0.015 s/10 s/0.05 s/10 s with a nitrogen flow of 20 sccm. A total of 140 cycles were carried out. The IZO transparency electrodes (40 nm) were deposited by radiofrequency sputtering (Angstrom EvoVac sputtering system) from a 3-inch IZO ceramic target (a radiofrequency power of 50 W) with shadow masks (an aperture area of 0.16 cm2). The chamber pressure was controlled at 0.6 mTorr with an argon flow rate of 15 sccm and oxygen flow rate of 0.06 sccm.

CIGS solar cells

First, the Mo back contact layer (~1 μm) was sputter-deposited on the cleaned stainless-steel substrate. A three-stage co-evaporation process was used to grow the CIGS absorber layer (~2 μm) in a vacuum chamber. An alkali postdeposition treatment was subsequently performed. The CdS buffer layer (∼35 nm) was grown via chemical bath deposition, and the intrinsic ZnO (i-ZnO) layer (~50 nm) and ZnO:Al layer (~200 nm) were prepared by radio frequency (RF) sputtering in sequence, followed by nickel/aluminium/nickel grid deposition for current collection.

Flexible monolithic perovskite/CIGS tandem solar cells

A 20 nm IZO layer was first deposited on the CIGS/CdS/i-ZnO/AZO substrate by using RF sputtering (Angstrom EvoVac sputtering system). Before sputtering, the samples were treated with UV–ozone for 20 min. Then, the NiOx thin layers were prepared by spin-coating the NiOx/H2O dispersive liquid (10 mg ml−1) on the UV–ozone-treated substrates at 3,000 rpm for 30 s and annealing at 110 °C for 10 min in ambient air. The samples were subsequently transferred to an N2 glovebox. For Me-4PACz deposition, 0.8–1.0 mg ml−1 Me-4PACz in ethanol was spin-coated onto the NiOx layer at 3,000 rpm for 30 s, followed by heating at 110 °C for 10 min. The perovskite layers were prepared by spin-coating 1.4 M precursor solution at 1,000 rpm@10 s followed by 5,000 rpm@40 s. During spinning, 200 μl of EA was dropped onto the film 20–18 s before the end of the second step. The as-cast pristine films were annealed at 110 °C for 10 min in the glovebox. After the perovskite films cooled down, 15 μl of CF3-PEAI solution (1.5 mg ml−1 in IPA) was dynamically spin-coated onto the perovskite film at 6,000 rpm for 30 s, then annealed at 90 °C for 3 min. After that, 1 nm of LiF (0.1 Å s−1) and 25 nm of C60 (0.2 Å s−1) were thermally deposited in sequence using an Angstrom EvoVac system. The SnO2 layers were then prepared by atomic layer deposition using a Veeco system with 140 cycles. The transparency electrodes (40 nm) were deposited by RF sputtering IZO (Angstrom EvoVac sputtering system) with an O2/Ar ratio of 0.4%. To complete the device preparation, 300 nm ring silver electrodes (defining an active area of 0.12 cm2) were deposited by thermal evaporation with a deposition rate of 3 Å s−1, followed by 140 nm LiF antireflective coating.

Sample characterizations

Steady-state PL

Steady-state PL measurements were conducted using a FLS1000 fluorescence spectrometer (Edinburgh Instruments). A 450 nm monochromatic light was used as the excitation source by filtering the white light from a 450 W xenon arc lamp. The intensity of the excitation light can be adjusted by changing the slit width. PL emission spectra with wavelengths ranging from 650 nm to 900 nm were collected. For time-dependent PL characterization to examine the photostability of perovskite films, steady-state PL tests were repeatedly performed on a perovskite film with a step size of 1 nm and a dwell time of 0.2 ms. The test interval between each spectrum was about 5 s. The slit width for the excitation source was fixed at 4 nm to ensure the same irradiation intensity for different samples. The power density of the 450 nm excitation light source was measured to be 27 mW cm−2. The total duration of the measurement for each sample was about 30 min, and about 22 steady-state PL spectra were recorded.

Temperature-dependent electric conductivity

The temperature-dependent electric conductivity measurements were performed on a low-temperature probe electrical measurement system. Lateral devices with a configuration of Au/perovskite (150 μm)/Au were fabricated, and the perovskite layers were deposited on glass substrates. The temperature of the lateral device was controlled by a cryogenic probe station (Lakeshore CRX-4K) in a vacuum chamber. Before testing, the temperature was first cooled to 160 K for 30 min. The current response was recorded at different temperatures by a semiconductor parameter analyser (Keithley 4200-SCS) under an applied voltage of 10 V. The temperature was raised step by step from 160 K to 340 K. Each step was stabilized at that temperature for 10 min before the current measurement was conducted.

The activation energy of ion migration could be extracted by fitting the raw data points using the Nernst–Einstein relation: σ(T) = (σ0/T)exp(−Ea/kT), where k is the Boltzmann constant, σ0 is a constant and T is the temperature. The ln(σT)–1/kT plot consists of two linear regions. Linear fitting of these two regions results in slopes corresponding to activation energy for ion migration (at higher temperature) and charge carrier transport (at lower temperature), respectively.

PL mapping

PL imaging characterizations were carried out using a confocal Raman imaging microscope (WITec alpha300 R). For in situ observation of the variation of perovskite films under an applied electric field by PL mapping, lateral devices (Au/perovskite/Au) were prepared with a channel width of 150 μm. A continuous-wave laser with a wavelength of 532 nm was used as the excitation source. A 50× objective lens with a numerical aperture of 0.55 (Zeiss LD EC Epiplan-Neofluar) was selected to both focus the laser excitation onto the sample and collect the PL spectra from the sample. Spatial information was obtained by moving the microscope stage, and therefore the sample, in the XY directions and acquiring an array of PL spectra that were then used to generate two-dimensional PL maps. The PL spectrum was detected using an electron-multiplying charge-coupled device, which is similar to a regular charge-coupled device but contains an additional on-chip electron-multiplying amplification stage that enables faster data readouts while preserving the high sensitivity. The scanning range was set to 300 μm × 300 μm, and the centre wavelength for detection was 730 nm. When the device is biased at 9 V, successive PL maps were acquired with each image acquisition time of about 80 s.

PL lifetime mapping

Fluorescence lifetime imaging was performed using a Leica Stellaris 8 FALCON confocal fluorescence microscope. A Leica 63× oil immersion objective (numerical aperture 1.40) was used for the measurement. Samples were placed on a cover glass with a thickness of 0.17 mm with the front surface downwards during the measurement. The excitation light source was a 448 nm laser line (repetition rate 2.5 MHz) extracted from a pulsed supercontinuum laser using an acousto-optical beam splitter (AOBS). With the excitation intensity fixed at 1.5%, the power was maintained at approximately 0.07 µW. The scanning speed was set to 100 Hz (lines s−1), with each line being accumulated 16 times. The image size was about 10.25 µm × 10.25 µm with a resolution of 512 × 512 pixels. The data were analysed and processed using Leica application suite (LAS) X FLIM software.

Trap density measurement

The distribution of trap density in energy was revealed through TAS analysis. TAS tests were performed on an electrical measurement system using an impedance analyser (Keysight E4990A). The devices were mounted on a temperature-control platform with low vacuum. The applied a.c. voltage signal had an amplitude of 20 mV, with a frequency range of 20 Hz to 1 MHz, and the d.c. bias was 0 V. The tDOS (Nt) is calculated by the following equation: ?t(??)=−?bi??d?d????, where VbiwCωqk and T are the build-in potential, depletion width, capacitance, angular frequency, elementary charge, Boltzmann constant and temperature, respectively. The frequency axis is transformed into an energy axis by Eω=kTln(ω0/ω), where ω0 is the attempt-to-escape angular frequency that equals 2πν0T2. The reduced attempt-to-escape frequency ν0 is determined by a preceding Arrhenius evaluation of ln(T2/ω) = Et/kT − ln(2πν0) based on temperature-dependent capacitance–frequency (Cf) measurements.

Other characterizations

SEM images were obtained using a scanning electron microscope (Zeiss Gemini 450) with an accelerating voltage of 5 kV and a beam current of 200 pA. XRD measurements were performed on an X-ray diffractometer (Bruker D8 Advance) with Cu Kα radiation (λ = 1.5418 Å) operating at 40 kV and 40 mA. The UV–visible absorption spectra were acquired using a Shimadzu UV 2700 spectrophotometer. XPS spectra were recorded on an X-ray photoelectron spectrometer (Thermo Fisher Escalab Xi+) by using a monochromatized Al Kα source (hv = 1,486.6 eV). The FTIR spectra were acquired using a FTIR spectrometer (Nicolet iS50, Thermo Scientific). The ToF-SIMS measurements were conducted on the ULVAC-PHI instrument with a Bi3+ (30 keV) primary ion beam for analysis and a gas (Ar/O2) gun (2 keV) for sputtering. Data were acquired for both positive ions and negative ions. The dynamic light scattering (DLS) measurements were performed using a light scattering/zeta potential analyser (BI-200SM/NanoBrook ZetaPALS). 1H NMR spectra were acquired on a Bruker 600 MHz solution NMR spectrometer (room temperature probe).

Device measurements

JV characteristics

The JV characteristics of single-junction PSCs were measured in an N2-filled glovebox at room temperature using a Keithley 2450 sourcemeter under simulated AM1.5 G irradiation (100 mW cm−2) from a xenon lamp solar simulator (Enlitech Class AAA). The light intensity was calibrated by a silicon reference cell before measurement, and no preconditioning was applied. The scanning step was 20 mV with a scanning rate of 100 mV s−1 and a delay time of 10 ms. The active areas of the opaque and semitransparent cells were defined by metal masks with aperture areas of 0.084 cm2 and 0.0912 cm2, respectively. To test the efficiencies of CIGS and four-terminal tandem solar cells, the JV characteristics were conducted in ambient conditions. The sourcemeter and solar simulator are the same as the ones in the glovebox. The CIGS photovoltaics are flexible with a structure of metal/Mo/CIGS/CdS/i-ZnO/AZO/Ag (finger electrode). The CIGS solar cells were measured with a 1 cm2 metal mask either stand-alone or below the semitransparent perovskite device under precalibrated one-sun illumination. When measuring the efficiency of the CIGS subcell, the semitransparent perovskite device without deposition of silver electrode was placed onto the CIGS to act as a light filter. The gap between them was filled with a refractive index matching fluid or liquid to reduce light reflection loss. For testing the monolithic perovskite/CIGS tandem, a metal mask with each aperture area of 0.092 cm2 was used.

Stabilized power output

The stabilized power outputs of PSC MPP were recorded by monitoring the current output over time, where the cells were biased at the MPP voltage extracted from the reverse-scanned JV curves.

EQE measurement

EQE measurements were carried out in ambient conditions using an EQE system (EnliTech). The devices were irradiated with monochromatic light chopped at a frequency of 210 Hz, and their photocurrent was measured with a lock-in amplifier. For the EQE measurements of the monolithic tandems, the EQE of the perovskite front subcell was measured using a light-emitting diode (LED) with an emission peak of 850 nm as the bias illumination. The CIGS subcell was measured by saturating the perovskite top cell using an LED with an emission peak of 550 nm. Before measurements, the EQE setup was calibrated by standard silicon and germanium photodetectors with a known spectral response.

Stability test

The stability (thermal and light stability) of the wide-bandgap perovskite films was examined by placing the perovskite films on a hotplate at 85 ± 1 °C or under simulated one-sun illumination (LED solar simulator) in an N2 glovebox for different periods of time, and their XRD patterns and/or PL spectra were repeatedly recorded for comparison. To assess the thermal stability of the wide-bandgap PSCs, unencapsulated devices were aged on an 85 ± 1 °C hotplate in a glovebox. To avoid the rapid degradation of the PSCs at such a relative high temperature, each device was covered by a glass substrate. The photovoltaic performances of these cells were monitored periodically through JV measurements. The operational stability of the unencapsulated devices was measured in an N2 glovebox using the MPP tracking mode on a custom-built multichannel ageing test system (Shenzhen Puri Materials). The illumination was provided by a white-light LED with intensity calibrated to match the one-sun condition. MPP was tracked by a perturb and observe algorithm that updates the MPP voltage for bias every 60 s, and an interval of 1 h was set for data collection. No active cooling was implemented during the test.

DFT calculations

Computational methodology

Our approach involved using first-principles calculations utilizing DFT with a plane-wave basis set and the projected augmented wave method. To account for atomic exchange-correlation effects, we applied the revised Perdew–Burke–Ernzerhof generalized gradient approximation, known as PBEsol. In addition, Grimme’s DFT-D3 scheme was integrated to address dispersion forces. Initial geometries were derived on the basis of the experimental lattice parameters of FAPbI3. For self-consistent field calculations and geometry optimizations, a plane-wave basis set with a cut-off energy of 400 eV and a 2 × 2 × 2 gamma-centred k-mesh were utilized. Ionic positions and cell dimensions underwent relaxation via a conjugate gradient algorithm until residual forces were below 0.02 eV Å−1. Unless otherwise specified, all DFT calculations were performed using theVienna Ab-initio simulation package (VASP) following the described methodology.

Iodine migration calculations

To evaluate migration barriers, we analysed the energy changes associated with iodine movement along the reaction pathway. The process involved optimizing both initial and final structures, followed by using a linear interpolation method to generate intermediate structures along the migration trajectory. Energy profiles were established using the nudged elastic band method and constrained energy minimization, considering 16 grid points during the calculations. The energy barriers for iodine migrations were computed by assessing the energy difference between the lowest energy state and the saddle point along the pathway.