Next Article in Journal
Irradiance in Mixed Coherent/Incoherent Structures: An Analytical Approach
Next Article in Special Issue
Special Issue: “Thin Films for Energy Harvesting, Conversion, and Storage”
Previous Article in Journal
Effects of Cobalt Content on the Microstructure, Mechanical Properties and Cavitation Erosion Resistance of HVOF Sprayed Coatings
Previous Article in Special Issue
Tunable Perfect Narrow-Band Absorber Based on a Metal-Dielectric-Metal Structure
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Compositional and Morphological Changes in Water-Induced Early-Stage Degradation in Lead Halide Perovskites

1
Institute of Applied Physics and Materials Engineering, University of Macau, Macau, SAR, China
2
Division of Physics and Applied Physics, School of Physical and Mathematical Sciences, Nanyang Technological University, 21 Nanyang Link, Singapore 637371, Singapore
3
Department of Science, School of Technology, Pandit Deendayal Petroleum University, Gandhiagar 382007, India
4
Institute of Materials Research and Engineering, A∗STAR (Agency for Science, Technology and Research), 2 Fusionopolis Way, Innovis #08-03, Singapore 138634, Singapore
*
Author to whom correspondence should be addressed.
Coatings 2019, 9(9), 535; https://doi.org/10.3390/coatings9090535
Submission received: 20 July 2019 / Revised: 12 August 2019 / Accepted: 14 August 2019 / Published: 22 August 2019
(This article belongs to the Special Issue Thin Films for Energy Harvesting, Conversion, and Storage)

Abstract

:
With tremendous improvements in lead halide perovskite-based optoelectronic devices ranging from photovoltaics to light-emitting diodes, the instability problem stands as the primary challenge in their development. Among all factors, water is considered as one of the major culprits to the degradation of halide perovskite materials. For example, CH3NH3PbI3 (MAPbI3) and CH(NH2)2PbI3 (FAPbI3) decompose into PbI2 in days under ambient conditions. However, the intermediate changes of this degradation process are still not fully understood, especially the changes in early stage. Here we perform an in-situ investigation of the early-stage MAPbI3 and FAPbI3 degradation under high water vapor pressure. By probing the surface and bulk of perovskite samples using near-ambient pressure X-ray photoelectron spectroscopy (NAP-XPS) and XRD, our findings clearly show that PbI2 formation surprisingly initiates below the top surface or at grain boundaries, thus offering no protection as a water-blocking layer on surface or grain boundaries to slow down the degradation process. Meanwhile, significant morphological changes are observed in both samples after water vapor exposure. In comparison, the integrity of MAPbI3 film degrades much faster than the FAPbI3 film against water vapor. Pinholes and large voids are found in MAPbI3 film while only small number of pinholes can be found in FAPbI3 film. However, the FAPbI3 film suffers from its phase instability, showing a fast α-to-δ phase transition. Our results highlight the importance of the compositional and morphological changes in the early stage degradation in perovskite materials.

1. Introduction

Halide perovskite materials with excellent optoelectronic properties show extensive potential in applications including photovoltaics [1], photodetectors [2], light-emitting diodes, and so on [3]. However, the inherent instability prevents these materials from long-term usage in devices [4]. The two widely used perovskite materials, methylammonium lead triiodide (MAPbI3) and formamidinium lead triiodide (FAPbI3), are susceptible to degradation by multiple factors, including water, oxygen, UV light, electrical field, and heating [4,5]. Synergistic degradation by the combination of multiple factors were also seen. Recent reports found the degradation process is greatly accelerated when perovskite is exposed to water, oxygen, and light together [6,7]. However, the excess water is still considered as the one of the major culprits causing the degradation of perovskite materials. Without any protection, both MAPbI3 and FAPbI3 films cannot last more than a few days in ambient air, quickly decomposed into PbI2 [8].
Based on these observations, two general strategies are proposed to improve device stability. The first strategy is to prevent perovskite layers from making any contact with water. Methods associated with this strategy include encapsulation [9,10] and surface passivation [11]. The second strategy is to enhance the durability of perovskite against water. Methods associated with this strategy include organic and inorganic doping [12,13,14] and film quality improvement [15]. These efforts successfully extend the lifetime of PSCs to thousands of hours, but still far from fulfilling commercialization requirements [4]. To further extend the lifetime of perovskite materials, new strategy based on thorough understanding of perovskite degradation is needed.
It is clear that MAPbI3 and FAPbI3 decompose into PbI2 at the end, but there are still uncertainties in the process of the degradation. For example, many studies suggest an indirect degradation pathway with sequential formation of two hydrated intermediates, monohydrate (CH3NH3PbI3·H2O) and dihydrate ((CH3NH3)4PbI6·2H2O), during ingress of water [16]. The monohydrate forms first and is reversible, while dihydrate forms only after monohydrate. Prolonged exposure dihydrate with water leads to final decomposition [17]. However, this indirect pathway may not be always valid. Direct PbI2 formation without steps of monohydrate and dihydrate formation was shown in some rigorous studies, raising doubts on the validity of the indirect degradation pathway. Schlipf et al. reported an earlier PbI2 formation than the appearance of monohydrate in their in-situ XRD measurements [18]. Recent near ambient pressure X-ray photoelectron spectroscopy (NAPXPS) study also claimed the surface of a perovskite thin film prepared by thermal deposition quickly decomposed into PbI2 at only 30% of relative humidity level [19]. Therefore, a direct degradation pathway leading to PbI2 may exist, causing premature deterioration of device performance.
Meanwhile, it is also uncertain the role of PbI2 in degradation. Excessive PbI2 appeared beneficial to the device performance, though it may affect the long-term photostability [20,21]. Previous studies usually assumed the PbI2 is formed from the surface, their data usually suggest a linear degradation speed, implying no water-blocking effect from PbI2 layer on surface [22]. Studies using in-vivo XRD measurements confirmed PbI2 formation in sample but was unable to determine whether PbI2 is on the top of surface or not [8,23]. PbI2 layer was observed by surface sensitive techniques such as XPS, but in these studies, perovskite films were usually measured after complete decomposition and missed the critical PbI2 formation period [22,24].
Another uncertainty is the effect of morphological change in the early stage of degradation. Morphology is considered as one of the important factors that affects the performance and lifetime of devices. Interestingly, it can be beneficial if a small fraction of water is introduced during fabrication [25]. Improved performance and lifetime are witnessed in these devices. These improvements are believed to be due to increased grain size as well as trap passivation [26,27]. Morphological change may be detrimental if grains are grown too large and cause film disintegration. However, the process at which the film loses its integrity is largely ignored. Previous studies focused on non-morphological changes such as trap passivation and grain boundary variation at early degradation stage [17,28]. Others studied much later degradation stage, when the perovskite is completely reverted to PbI2 [8,29]. Therefore, the morphology evolution in degradation needs further investigation.
Here we report a study focusing on the critical changes in the early stage of MAPbI3 and FAPbI3 degradation in a precise and controlled condition. Multiple experimental techniques are applied to study water-induced compositional and morphological evolutions under high water vapor pressure at room temperature. Surprisingly, the surface of degraded samples remained stoichiometric to the pristine phase while PbI2 is found in bulk. This finding excludes the possibility of self-passivation by PbI2 and supports the coexistence of both degradation pathways under high water partial pressure condition. Meanwhile, prominent morphological changes such as pinholes and large voids are observed in MAPbI3 sample, revealing loss of film integrity initiated in the early stage degradation. In comparison, FAPbI3 film remains largely intact with much less pinholes, indicating a much slower morphology evolution. However, the FAPbI3 sample suffers from water-induced phase change. XRD data show the dominant δ phase in the degraded FAPbI3 sample, indicating a fast α to δ phase transition in the early stage of the degradation. Our results successfully clarify compositional and morphological changes in the early-stage degradation of perovskite thin films.

2. Materials and Methods

The MAPbI3 sample and FAPbI3 sample are prepared by standard solution preparation methods with anti-solvent treatment. All organic cation salts were purchased from Dyesol (Queanbeyan, Australia) while lead iodide was bought from Acros Organics (Geel, Belgium). MAPbI3 and FAPbI3 perovskite solutions were prepared by mixing precursors in stoichiometric ratio (1 M concentration) in anhydrous dimethylformamide (DMF from Sigma Aldrich, St. Louis, MO, USA). The perovskite thin films were spun-coated on cleaned indium-tin-oxide (ITO) substrates at 5000 rpm for 12 s. The anti-solvent treatment was performed by dripping 100 μL toluene on the spinning substrates 9 s prior to the end of spinning. Subsequently, MAPbI3 samples were annealed at 100 °C for 30 min while FAPbI3 samples were annealed at 160 °C for 10 min. Prepared samples were transferred from glovebox to XPS system through an air-tight container to minimize contact with external atmosphere. XPS measurements were conducted in-situ before and after water vapor dosing in the high-pressure gas cell. SEM and XRD were ex-situ measured before and after water vapor dosing in near-ambient pressure X-ray photoelectron spectroscopy (NAP-XPS) system.
The film crystallinity was measured using Bruker D8 discover X-ray diffractometer (Billerica, MA, USA) with Cu Kα radiation (λ = 1.54 Å). The morphology is measured by JEOL FESEM 6700 (Akishima, Japan). The surface composition was measured by PREVAC NAP-XPS system (Rogow, Poland) with monochromatic Al Kα X-ray source ( = 1486.7 eV).

3. Results and Discussion

To investigate the degradation on the perovskite surface, it is important to control the hydration level precisely. Many previous works used ambient condition with different relative humidity (RH) values to probe the degradation process [23,30,31]. However, RH is not an absolute unit and can largely vary due to temperature change. Therefore, RH value is not an accurate parameter to measure the content of water. Furthermore, gases in the air such as oxygen may also have implications on the perovskite degradation, making the process more complicated [28,32]. Here we use the gas cell inside a NAPXPS system to study the interaction of perovskite samples to water vapor. The in-situ environment excludes any other external factors and the measured water partial pressure is more accurate than RH value.
MAPbI3 and FAPbI3 samples are exposed to 23 mbar of water vapor pressure in NAPXPS cells separately for two times with 1 h each. The water vapor partial pressure is about 80% RH for the measured cell temperature of 23 °C. The lower exposure pressure at 18 mbar or below shows no distinguished changes. Extending the exposure time to 6 h resulted in further decreased XPS intensities as well as I/Pb, N/Pb, and C/Pb ratios. Therefore, we only discuss the 2 h exposure in detail. To monitor potential surface composition changes, XPS high-resolution spectra from MAPbI3 and FAPbI3 before and after water vapor exposure were acquired (Figure 1). It should be highlighted that measurements were done at an ultrahigh vacuum condition after the system was fully recovered from water exposure. Therefore, signals from monohydrate and dihyrate are not expected. For MAPbI3, the spectra of I 3d5/2, Pb 4f, and N 1s contain single peaks at 619.6, 138.8, and 402.7 eV, respectively. All of them originated from MAPbI3 [33]. In the spectrum of C 1s, two peaks at 286.7 and 285.5 eV can be distinguished. The higher binding energy peak (C1) is from C–N bonding in MA cation. The lower binding energy peak (C2) is attributed to the adventitious carbon and it is not related to perovskite itself. The binding energies of carbon peaks are consistent with previous reports [33,34]. After water vapor exposure, peak intensities of I 3d5/2, Pb 4f, N 1s, and C1 gradually decrease. Only the C2 peak shows a slight increase. Meanwhile, the peak position remains unchanged. The drop in peak intensities are also observed for the FAPbI3 sample (Figure 1e–g). However, the magnitude of the drop is smaller. The relative intensities of different elements are summarized in Figure 2. These intensities in FAPbI3 drop about 10% to 20%, while in MAPbI3, the intensities drop more than 30% for iodine and close to 50% for nitrogen. To clarify if this decrease is related to the decomposition, the normalized atomic ratios between iodine, lead, nitrogen, and C1 are compared, as listed in Table 1. For MAPbI3, the I/Pb ratio is 3.1 in the pristine sample, indicating a slightly iodine rich on the surface. After the first and second water vapor exposure, this ratio further increased to 3.5 and 3.7, respectively. Meanwhile, the N/Pb ratio and C/Pb ratio are also maintained above 1, indicating no sign of organic cation deficiency at surface. Therefore, the atomic ratio change of Pb/I only suggests enrichment of organic cations or deficiency of lead atoms. None of these changes support PbI2 formation. For FAPbI3, the composition change is even smaller. The I/Pb ratio is between 3.1–3.6 and N/Pb ratio is between 2.1–2.2. The C/Pb ratio is slightly lower than 1, but no systematic decrease after water vapor exposure was observed. From the atomic ratio data, it can be concluded that there is no sign of PbI2 formation. It appears to contradict a previous NAPXPS study, in which the perovskite surface is completely decomposed at 9 mbar of water partial pressure [19]. This contradiction can be justified by the difference in sample fabrication. Unlike the solution process method, the perovskite films prepared by vacuum deposition usually have smaller domain sizes and poorer crystalline quality. A lower-quality film may result in much faster degradation. Instead, our results are consistent with studies using solution-processed perovskite samples, where no significant surface degradation were reported [24,35].
Since there is no trace of PbI2 formation on the surface, the decrease in intensities is probably due to reduced film coverage. This is supported by the observation in XPS wide scans (Figure 3). After water vapor exposure, new peaks originated from In 3d and O 1s, indicating the partial exposure of ITO substrates. This is clear evidence that large voids form in MAPbI3 thin film. The In 3d signal is much stronger in the MAPbI3 sample, suggesting greater film area shrinkage. Therefore, it can be concluded that though the surface composition barely changes, the film coverage greatly reduced. Large voids compromise the device integrity and is probably the determining factor to device lifetime.
Besides surface sensitive XPS measurements, XRD is used as a complementary technique to evaluate the changes in the bulk. In Figure 4, the XRD spectra of MAPbI3 and FAPbI3 are shown before and after water vapor exposure. The spectra of both pristine samples only contain peaks from perovskite itself. No PbI2 peak at 12.7° is observed [36]. After water vapor exposure, a small PbI2 peak is observed in both samples. Since there is no trace of PbI2 formation on surface, the signal must come from the site below the very top surface. It appears counterintuitive that PbI2 formation starts from the bulk instead of surface. One probable explanation is due to the fast diffusion of water molecules and organic cations [37,38]. The water diffusion into the bulk may cause decomposition both inside and on the surface. However, when volatile organic cation escapes via the surface, it may partially compensate the loss of cations at the surface region, delaying PbI2 formation on the surface. It is also possible that part of PbI2 crystals formed along the grain boundaries without much exposure to the top surface. Without surface PbI2 formation, it may be beneficial to keep the interface unchanged, but it also means that slowing down the degradation by PbI2 passivation will not occur. In the FAPbI3 sample, the diffraction peak (001) at 13.8° disappeared and a peak at 11.8° emerged. Other high-order peaks also emerged at new two theta degrees. This change is consistent with the α-to-δ phase transition observed in FAPbI3 previously [35]. Therefore, FAPbI3 is more sensitive to water-induced phase change.
The XPS data suggest film coverage is reduced after water vapor exposure. To understand how severe the reduction in surface coverage is, the surface of both samples are ex-situ measured by SEM (Figure 5). The pristine MAPbI3 film is compact with no observable pinholes, while the FAPbI3 film only contains a few. The FAPbI3 film shows similar grain size with small bright grains on the surface. The formation of these bright grains is probably due to the annealing process to convert the film from δ phase to α phase. However, during annealing, grain recrystallization also occurs, leaving small grains scattered on the surface. This phenomenon was observed in the previous study as well [39]. After water vapor exposure, the number of pinholes is greatly increased on the MAPbI3 surface. These pinholes are formed along the grain boundaries and do not affect remaining grains. Meanwhile, a grain-coarsening effect is observed. The average grain size has increased from about 0.13 to 0.23 μm2. All these observations can be attributed to water-induced recrystallization [26]. On the FAPbI3 surface, the increase in pinholes is far less. Instead, needle-like crystals scatter on the surface, probably related to the δ phase of FAPbI3 [40]. In summary, more pinholes are found on the MAPbI3 surface, indicating a poorer morphology stability against water.
Except small pinholes, a much larger-scale reconstruction pattern is also observed on the MAPbI3 surface (Figure 6). This reconstruction is not seen on the FAPbI3 surface. The pattern has a spherulite-like feature with sub-millimeter scale. This pattern was shown in previous works, but its appearance was not discussed [17,41]. This spherulite-like pattern is highly symmetric, where each half assembles an alluvial fan. Inside the pattern, the film has been completely reconstructed, forming long crystals with fused boundary pointing towards the joint of two alluvial fans. Meanwhile, voids are seen between reconstructed patterns or between reconstruction pattern and nearby film. The symmetry of this pattern implies the film reconstruction initiates from the center and then spreads out. The formation of these reconstruction center could be due to either localized surface hydrophilic spots, which make water soaking easy or temporary, and localized humidity level variation, which may cause micro water drops formation on the surface, or both. Interestingly, the relative two-fold symmetric alluvial fan-like structure also indicates non-uniform diffusion on the surface. Along preferential water diffusion direction, adsorbed water diffuses much faster, allowing quick bilateral movement of water to two opposite ends. At the direction perpendicular to the preferential diffusion direction, water diffuses much slower and does not spread out. It is possible that both monohydrate and dihydrate were formed during recrystallization. However, the film is always measured after water exposure and at ultrahigh vacuum condition. Therefore, from the morphology shown in SEM images, perovskite crystals reversed back from hydrates. The void created by this pattern is much larger than the pinholes formed at grain boundaries. Therefore, it poses a much more severe threat to integrity of devices, and may cause premature device failure. The SEM results are consistent with the XPS data in which nearly 30% and 10% of the film area disappeared after water exposure of MAPbI3 and FAPbI3, respectively.

4. Conclusions

In conclusion, we find the degradation of perovskite materials involves multiple changes. For MAPbI3, the initial degradation is dominated by water-induced recrystallization, which causes drastic changes in film morphology. For FAPbI3, the initial change mainly occurs due to the phase transition. For both perovskite films, we reveal the PbI2 formation occurring in the bulk rather than on the surface, probably mainly at grain boundaries. Therefore, PbI2 in the degraded film does not act as a water-blocking barrier to slow down the degradation process. The tendency of PbI2 formation at grain boundaries rather than on the surface is also observed in non-stoichiometric sample with excess PbI2 or stoichiometric samples degraded by heat. Therefore, it can be generally concluded that degradation process of perovskite film is not slowed by the emergence of PbI2. Our results highlight the importance of the external water-blocking layer because PbI2 does not form a continuous film on the surface. Also, a more robust device structure with self-supporting layers may have better resistance to morphology change at the earlier stage of degradation.

Author Contributions

Conceptualization, S.C.; investigation, S.C., A.S., and J.P.; writing—original draft preparation, S.C.; writing—review and editing, S.C., A.S., J.P., and T.C.S., funding acquisition, S.C. and T.C.S.

Funding

This research was funded by Ministry of Education AcRF Tier 2 (MOE2016-T2-1-034), the Singapore National Research Foundation Investigatorship Programme (NRF-NRFI-2018-04) and the University of Macau startup fund (SRG2018-00140-IAPME).

Acknowledgments

Authors acknowledge technical support in experiments from Lim Poh Chong and Bussolotti Fabio in IMRE.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yang, W.S.; Park, B.-W.; Jung, E.H.; Jeon, N.J.; Kim, Y.C.; Lee, D.U.; Shin, S.S.; Seo, J.; Kim, E.K.; Noh, J.H.; et al. Iodide management in formamidinium-lead-halide–based perovskite layers for efficient solar cells. Science 2017, 356, 1376–1379. [Google Scholar] [CrossRef] [PubMed]
  2. Dou, L.T.; Yang, Y.; You, J.B.; Hong, Z.R.; Chang, W.H.; Li, G.; Yang, Y. Solution-processed hybrid perovskite photodetectors with high detectivity. Nat. Commun. 2014, 5, 5404. [Google Scholar] [CrossRef] [PubMed]
  3. Du, B.; Xia, Y.; Wei, Q.; Xing, G.; Chen, Y.; Huang, W. All-inorganic perovskite nanocrystals-based light emitting diodes and solar cells. Chemnanomat 2019, 5, 266–277. [Google Scholar] [CrossRef]
  4. Leijtens, T.; Eperon, G.E.; Noel, N.K.; Habisreutinger, S.N.; Petrozza, A.; Snaith, H.J. Stability of metal halide perovskite solar cells. Adv. Energy Mater. 2015, 5, 1500963. [Google Scholar] [CrossRef]
  5. Li, B.; Li, Y.; Zheng, C.; Gao, D.; Huang, W. Advancements in the stability of perovskite solar cells: Degradation mechanisms and improvement approaches. RSC Adv. 2016, 6, 38079–38091. [Google Scholar] [CrossRef]
  6. Aristidou, N.; Eames, C.; Islam, M.S.; Haque, S.A. Insights into the increased degradation rate of CH3NH3PbI3 solar cells in combined water and O2 environment. J. Mater. Chem. A 2017, 5, 25469–25475. [Google Scholar] [CrossRef]
  7. Norbert, H.N.; Felix, L.; Brus, V.V.; Shargaieva, O.; Rappich, J. Unraveling the light-induced degradation mechanisms of CH3NH3PbI3 perovskite film. Adv. Electron. Mater. 2017, 3, 1700158. [Google Scholar] [CrossRef]
  8. Habisreutinger, S.N.; Leijtens, T.; Eperon, G.E.; Stranks, S.D.; Nicholas, R.J.; Snaith, H.J. Carbon nanotube/polymer composites as a highly stable hole collection layer in perovskite solar cells. Nano Lett. 2014, 14, 5561–5568. [Google Scholar] [CrossRef]
  9. Wei, Z.; Zheng, X.; Chen, H.; Long, X.; Wang, Z.; Yang, S. A multifunctional C + epoxy/Ag-paint cathode enables efficient and stable operation of perovskite solar cells in watery environments. J. Mater. Chem. A 2015, 3, 16430–16434. [Google Scholar] [CrossRef]
  10. Bella, F.; Griffini, G.; Correa-Baena, J.-P.; Saracco, G.; Gratzel, M.; Hagfeldt, A.; Turri, S.; Gerbaldi, C. Improving efficiency and stability of perovskite solar cells with photocurable fluoropolymers. Science 2016, 354, 203–206. [Google Scholar] [CrossRef]
  11. Yang, S.; Wang, Y.; Liu, P.; Cheng, Y.; Zhao, H.; Yang, H. Functionalization of perovskite thin films with moisture-tolerant molecules. Nat. Energy 2016, 1, 15016. [Google Scholar] [CrossRef]
  12. Hu, Y.; Qiu, T.; Bai, F.; Miao, X.; Zhang, S. Enhancing moisture-tolerance and photovoltaic performances of FAPbI3 by bismuth incorporation. J. Mater. Chem. A 2017, 5, 25258–25265. [Google Scholar] [CrossRef]
  13. Wang, Z.; Lin, Q.; Chmiel, F.P.; Sakai, N.; Herz, L.M.; Snaith, H.J. Efficient ambient-air-stable solar cells with 2D-3D heterostructured butylammonium-caesium-formamidinium lead halide perovskites. Nat. Energy 2017, 2, 17135. [Google Scholar] [CrossRef]
  14. Saidaminov, M.I.; Kim, J.; Jain, A.; Quintero-Bermudez, R.; Tan, H.; Long, G.; Tan, F.; Johnston, A.; Zhao, Y.; Voznyy, O.; et al. Suppression of atomic vacancies via incorporation of isovalent small ions to increase the stability of halide perovskite solar cells in ambient air. Nat. Energy 2018, 3, 648–654. [Google Scholar] [CrossRef]
  15. Kim, J.H.; Williams, S.T.; Cho, N.; Chueh, C.-C.; Jen, A.K.-Y. Enhanced environmental stability of planar heterojunction perovskite solar cells based on blade-coating. Adv. Energy Mater. 2015, 5, 1401229. [Google Scholar] [CrossRef]
  16. Song, Z.; Abate, A.; Watthage, S.C.; Liyanage, G.K.; Phillips, A.B.; Steiner, U.; Graetzel, M.; Heben, M.J. Perovskite solar cell stability in humid air: Partially reversible phase transitions in the PbI2–CH3NH3I–H2O system. Adv. Energy Mater. 2016, 6, 1600846. [Google Scholar] [CrossRef]
  17. Leguy, A.M.A.; Hu, Y.; Campoy-Quiles, M.; Alonso, M.I.; Weber, O.J.; Azarhoosh, P.; van Schilfgaarde, M.; Weller, M.T.; Bein, T.; Nelson, J.; et al. Reversible hydration of CH3NH3PbI3 in films, single crystals, and solar cells. Chem. Mater. 2015, 27, 3397–3407. [Google Scholar] [CrossRef]
  18. Schlipf, J.; Hu, Y.H.; Pratap, S.; Bießmann, L.; Hohn, N.; Porcar, L.; Bein, T.; Docampo, P.; Muller-Buschbaum, P. Shedding light on the moisture stability of 3D/2D hybrid perovskite heterojunction thin films. ACS Appl. Energy Mater. 2019, 2, 1011–1018. [Google Scholar] [CrossRef]
  19. Ke, J.C.-R.; Walton, A.S.; Lewis, D.J.; Tedstone, A.; O’Brien, P.; Thomas, A.G.; Flavell, W.R. In situ investigation of degradation at organometal halide perovskite surfaces by X-ray photoelectron spectroscopy at realistic water vapour pressure. Chem. Commun. 2017, 53, 5231–5234. [Google Scholar] [CrossRef] [Green Version]
  20. Bi, D.; Tress, W.; Dar, M.I.; Gao, P.; Luo, J.; Renevier, C.; Schenk, K.; Abate, A.; Giordano, F.; Baena, J.-P.C.; et al. Efficient luminescent solar cells based on tailored mixed-cation perovskites. Sci. Adv. 2016, 2, e1501170. [Google Scholar] [CrossRef]
  21. Jesper, J.; Juan-Pablo, C.B.; Elham, H.A.; Bertrand, P.; Samuel, D.S.; Marine, E.F.B.; Wolfgang, T.; Kurt, S.; Joel, T.; Jacques, E.M.; et al. Unreacted PbI2 as a double-edge sword for enhancing the performance of perovskite solar cells. J. Am. Chem. Soc. 2016, 138, 10331–10343. [Google Scholar] [CrossRef]
  22. Wang, C.C.; Gao, Y.L. Stability of perovskites at the surface analytic level. J. Phys. Chem. Lett. 2018, 9, 4657–4666. [Google Scholar] [CrossRef] [PubMed]
  23. Wei, W.; Hu, Y.H. Catalytic role of H2O in degradation of inorganic-organic perovskite (CH3NH3PbI3) in air. Int. J. Energy Res. 2017, 41, 1063–1069. [Google Scholar] [CrossRef]
  24. Philippe, B.; Park, B.-W.; Lindblad, R.; Oscarsson, J.; Ahmadi, S.; Johansson, E.M.J.; Rensmo, H. Chemical and electronic structure characterization of lead halide perovskites and stability behavior under different exposures—A photoelectron spectroscopy investigation. Chem. Mater. 2015, 27, 1720–1731. [Google Scholar] [CrossRef]
  25. You, J.; Yang, Y.; Hong, Z.; Song, T.B.; Meng, L.; Liu, Y.; Jiang, C.; Zhou, H.; Chang, W.H.; Li, G.; et al. Moisture assisted perovskite film growth for high performance solar cells. Appl. Phys. Lett. 2014, 105, 183902. [Google Scholar] [CrossRef] [Green Version]
  26. Zhang, W.; Xiong, J.; Li, J.; Daoud, W.A. Mechanism of water effect on enhancing the photovoltaic performance of triple-cation hybrid perovskite solar cells. ACS Appl. Mater. Interfaces 2019, 11, 12699–12708. [Google Scholar] [CrossRef]
  27. Solanki, A.; Lim, S.S.; Mhaisalkar, S.; Sum, T.C. Role of water in suppressing recombination pathways in CH3NH3PbI3 perovskite solar cells. ACS Appl. Mater. Interfaces 2019, 11, 25474–25482. [Google Scholar] [CrossRef]
  28. Yang, J.M.; Yuan, Z.C.; Liu, X.J.; Braun, S.; Li, Y.Q.; Tang, J.X.; Gao, F.; Duan, C.G.; Fahlman, M.; Bao, Q.Y. Oxygen- and water-induced energetics degradation in organometal halide perovskites. ACS Appl. Mater. Interfaces 2018, 10, 16225–16230. [Google Scholar] [CrossRef]
  29. Christians, J.A.; Miranda Herrera, P.A.; Kamat, P.V. Transformation of the excited state and photovoltaic efficiency of CH3NH3PbI3 perovskite upon controlled exposure to humidified Air. J. Am. Chem. Soc. 2015, 137, 1530–1538. [Google Scholar] [CrossRef]
  30. Shirayama, M.; Kato, M.; Miyadera, T.; Sugita, T.; Fujiseki, T.; Hara, S.; Kadowaki, H.; Murata, D.; Chikamatsu, M.; Fujiwara, H.; et al. Degradation mechanism of CH3NH3PbI3 perovskite materials upon exposure to humid air. J. Appl. Phys. 2016, 119, 115501. [Google Scholar] [CrossRef]
  31. Li, D.; Bretschneider, S.A.; Bergmann, V.W.; Hermes, I.M.; Mars, J.; Klasen, A.; Lu, H.; Tremel, W.; Mezger, M.; Butt, H.J.; et al. Humidity-induced grain boundaries in MAPbI3 perovskite films. J. Phys. Chem. C 2016, 120, 6363–6368. [Google Scholar] [CrossRef]
  32. Ralaiarisoa, M.; Salzmann, I.; Zu, F.S.; Koch, N. Effect of water, oxygen, and air exposure on CH3NH3PbI3xClx perovskite surface electronic properties. Adv. Electron. Mater. 2018, 4, 1800307. [Google Scholar] [CrossRef]
  33. Chen, S.; Goh, T.W.; Sabba, D.; Chua, J.; Mathews, N.; Huan, C.H.A.; Sum, T.C. Energy level alignment at the methylammonium lead iodide/copper phthalocyanine interface. APL Mater. 2014, 2, 081512. [Google Scholar] [CrossRef] [Green Version]
  34. Hawash, Z.; Raga, S.R.; Son, D.Y.; Ono, L.K.; Park, N.G.; Qi, Y.B. Interfacial modification of perovskite solar cells using an ultrathin MAI layer leads to enhanced energy level alignment, efficiencies, and reproducibility. J. Phys. Chem. Lett. 2017, 8, 3947–3953. [Google Scholar] [CrossRef] [PubMed]
  35. Yamanaka, S.; Hayakawa, K.; Cojocaru, L.; Tsuruta, R.; Sato, T.; Mase, K.; Uchida, S.; Nakayama, Y. Electronic structures and chemical states of methylammonium lead triiodide thin films and the impact of annealing and moisture exposure. J. Appl. Phys. 2018, 123, 165501. [Google Scholar] [CrossRef]
  36. Smecca, E.; Numata, Y.; Deretzis, I.; Pellegrino, G.; Boninelli, S.; Miyasaka, T.; La Magna, A.; Alberti, A. Stability of solution-processed MAPbI3 and FAPbI3 layers. Phys. Chem. Chem. Phys. 2016, 18, 13413–13422. [Google Scholar] [CrossRef] [PubMed]
  37. Sun, P.P.; Chi, W.J.; Li, Z.S. Effects of water molecules on the chemical stability of MAGeI3 perovskite explored from a theoretical viewpoint. Phys. Chem. Chem. Phys. 2016, 18, 24526–24536. [Google Scholar] [CrossRef] [PubMed]
  38. Tong, C.; Geng, W.; Tang, Z.; Yam, C.; Fan, X.; Liu, J.; Lau, W.; Liu, L. Uncovering the veil of the degradation in perovskite CH3NH3PbI3 upon humidity exposure: A first-principles study. J. Phys. Chem. Lett. 2015, 6, 3289–3295. [Google Scholar] [CrossRef]
  39. Liu, T.; Zong, Y.; Zhou, Y.; Yang, M.; Li, Z.; Game, O.S.; Zhu, K.; Zhu, R.; Gong, Q.; Padture, N.P. High-performance formamidinium-based perovskite solar cells via microstructure-mediated δ-to-α phase transformation. Chem. Mater. 2017, 29, 3246–3250. [Google Scholar] [CrossRef]
  40. Tseng, W.S.; Jao, M.H.; Hsu, C.C.; Huang, J.S.; Wu, C.I.; Yeh, N.C. Stabilization of hybrid perovskite CH3NH3PbI3 thin films by graphene passivation. Nanoscale 2017, 9, 19227–19235. [Google Scholar] [CrossRef]
  41. Yang, J.L.; Fransishyn, K.M.; Kelly, T.L. Comparing the effect of mesoporous and planar metal oxides on the stability of methylammonium lead iodide thin films. Chem. Mater. 2016, 28, 7344–7352. [Google Scholar] [CrossRef]
Figure 1. High-resolution spectra of I 3d5/2, N 1s, Pb 4f and C 1s obtained from MAPbI3 (ad) and FAPbI3 (eh) samples before and after water vapor exposure to different periods, respectively.
Figure 1. High-resolution spectra of I 3d5/2, N 1s, Pb 4f and C 1s obtained from MAPbI3 (ad) and FAPbI3 (eh) samples before and after water vapor exposure to different periods, respectively.
Coatings 09 00535 g001aCoatings 09 00535 g001b
Figure 2. The relative intensities change after different water vapor exposure period. (a) Elemental intensity changes of MAPbI3 sample. (b) Elemental intensity changes of FAPbI3 sample.
Figure 2. The relative intensities change after different water vapor exposure period. (a) Elemental intensity changes of MAPbI3 sample. (b) Elemental intensity changes of FAPbI3 sample.
Coatings 09 00535 g002
Figure 3. XPS wide scans obtained from MAPbI3 and FAPbI3 samples before and after water vapor exposure. (a) MAPbI3 sample; (b) FAPbI3 sample.
Figure 3. XPS wide scans obtained from MAPbI3 and FAPbI3 samples before and after water vapor exposure. (a) MAPbI3 sample; (b) FAPbI3 sample.
Coatings 09 00535 g003
Figure 4. XRD data of MAPbI3 and FAPbI3 sample before and after water vapor exposure. PbI2 spectra are shown as a reference. Peaks marked with * is from indium-tin-oxide (ITO) substrate. (a) XRD spectra of the MAPbI3 sample before and after water exposure. (b) XRD spectra of the FAPbI3 sample before and after water exposure.
Figure 4. XRD data of MAPbI3 and FAPbI3 sample before and after water vapor exposure. PbI2 spectra are shown as a reference. Peaks marked with * is from indium-tin-oxide (ITO) substrate. (a) XRD spectra of the MAPbI3 sample before and after water exposure. (b) XRD spectra of the FAPbI3 sample before and after water exposure.
Coatings 09 00535 g004
Figure 5. SEM images of MAPbI3 and FAPbI3 samples before and after water vapor exposure for 2 h. (a) Pristine MAPbI3 sample; (b) MAPbI3 sample after degradation; (c) pristine FAPbI3 sample; (d) FAPbI3 sample after degradation.
Figure 5. SEM images of MAPbI3 and FAPbI3 samples before and after water vapor exposure for 2 h. (a) Pristine MAPbI3 sample; (b) MAPbI3 sample after degradation; (c) pristine FAPbI3 sample; (d) FAPbI3 sample after degradation.
Coatings 09 00535 g005
Figure 6. Large-scale morphology changes on the MAPbI3 sample after water exposure. (a) Spherulite-like pattern on the MAPbI3 surface. (b) Zoomed image of the center of spherulite-like pattern.
Figure 6. Large-scale morphology changes on the MAPbI3 sample after water exposure. (a) Spherulite-like pattern on the MAPbI3 surface. (b) Zoomed image of the center of spherulite-like pattern.
Coatings 09 00535 g006
Table 1. Atomic ratio of different elements obtained by XPS from MAPbI3 and FAPbI3 before and after water vapor exposure.
Table 1. Atomic ratio of different elements obtained by XPS from MAPbI3 and FAPbI3 before and after water vapor exposure.
Exposure TimeMAPbI3FAPbI3
I/PbN/PbC/PbI/PbN/PbC/Pb
0 h3.11.61.53.12.20.8
1 h3.51.01.53.62.10.7
2 h3.71.52.43.32.10.9

Share and Cite

MDPI and ACS Style

Chen, S.; Solanki, A.; Pan, J.; Sum, T.C. Compositional and Morphological Changes in Water-Induced Early-Stage Degradation in Lead Halide Perovskites. Coatings 2019, 9, 535. https://doi.org/10.3390/coatings9090535

AMA Style

Chen S, Solanki A, Pan J, Sum TC. Compositional and Morphological Changes in Water-Induced Early-Stage Degradation in Lead Halide Perovskites. Coatings. 2019; 9(9):535. https://doi.org/10.3390/coatings9090535

Chicago/Turabian Style

Chen, Shi, Ankur Solanki, Jisheng Pan, and Tze Chein Sum. 2019. "Compositional and Morphological Changes in Water-Induced Early-Stage Degradation in Lead Halide Perovskites" Coatings 9, no. 9: 535. https://doi.org/10.3390/coatings9090535

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop