Next Article in Journal
Interfacial Characteristics and Mechanical Properties of CuSn15/HT250 Prepared via Additive Manufacturing Combined with an Inconel 718 Interlayer
Next Article in Special Issue
Recent Advances on Thermoelectric Silicon for Low-Temperature Applications
Previous Article in Journal
Cotton Fabric Modified with a PAMAM Dendrimer with Encapsulated Copper Nanoparticles: Antimicrobial Activity
Previous Article in Special Issue
High Temperature Electrical Properties of Co-Substituted La4BaCu5O13+δ Thin Films Fabricated by Sputtering Method
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Two-Step Synthesis of Bismuth-Based Hybrid Halide Perovskite Thin-Films

1
School of Engineering and Materials Science (SEMS), Queen Mary University of London, Mile End Road, London E1 4NS, UK
2
Department of Materials Science and Solar Energy Research Center (MIB-SOLAR), University of Milano-Bicocca, Via Cozzi 55, I-20125 Milan, Italy
3
Department of Physics and Astronomy, Queen Mary University of London, Mile End Road, London E1 4NS, UK
*
Author to whom correspondence should be addressed.
Materials 2021, 14(24), 7827; https://doi.org/10.3390/ma14247827
Submission received: 5 November 2021 / Revised: 6 December 2021 / Accepted: 15 December 2021 / Published: 17 December 2021
(This article belongs to the Special Issue Thin Films for Energy Production and Storage)

Abstract

:
Lead halide perovskites have been revolutionary in the last decade in many optoelectronic sectors. Their bismuth-based counterparts have been considered a good alternative thanks to their composition of earth-abundant elements, good chemical stability, and low toxicity. Moreover, their electronic structure is in a quasi-zero-dimensional (0D) configuration, and they have recently been explored for use beyond optoelectronics. A significant limitation in applying thin-film technology is represented by the difficulty of synthesizing compact layers with easily scalable methods. Here, the engineering of a two-step synthesis in an air of methylammonium bismuth iodide compact thin films is reported. The critical steps of the process have been highlighted so that the procedure can be adapted to different substrates and application areas.

1. Introduction

Bismuth-based hybrid halide perovskites are composed of elements widely available from natural sources and show improved chemical stability and less toxicity compared to their lead-based counterparts [1,2,3,4,5]. For these reasons, bismuth-based perovskites have emerged as a new research field, being explored for use in solar cells; however, the results obtained have not led to improvements in the state-of-the-art, due to the poor substrate coverage and to the charge confinement that limit their use in photovoltaics [3,6]. The stoichiometry, indeed, leads the structure to settle in a quasi-zero dimensional (0D) configuration, where the metal halide octahedra are almost isolated [3,7,8]. Their 0D electronic structure is highly explored for use in photodiodes [9], thermoelectric generators [10], and detectors [11]. Still, the production of a compact and continuous film remains challenging, and the best results are obtained by vapour-assisted deposition, even though the use of cheap and straightforward synthetic methods is a requirement becoming more and more crucial in the device design [3,12]. Unfortunately, the solution-processed thin films that have been reported so far are neither compact nor uniform [3,13,14,15,16]. Encouraging results on (CH3NH3)3Bi2I9 (MABI) deposition were obtained using two-step depositions on mesoporous TiO2 [17], where the BiI3 layer is dipped in a CH3NH3I (MAI) solution to complete the synthesis. However, the growth substrate has a decisive influence on the morphology of the final film: morphologies with vastly different features could be produced when varying the substrate from TiO2 (compact or mesoporous) to NiO, Al2O3 or PEDOT:PSS [3,18,19]. Here, a quartz-coated glass with high wettability and low roughness was used to reduce the substrate impact on the synthesis as much as possible. So far, BiI3 has been deposited by spin-coating [20], epitaxial electrodeposition [21], and thermal evaporation [22]. In this work, in order to produce a compact layer, BiI3 was deposited by thermal evaporation of commercial powders, while MAI was introduced by dipping in a solution in isopropyl alcohol (IPA). Chen et al. followed the same synthetic route [23], employing a spin-coated BiI3 thin film, but the resulting substrate coverage was incomplete [23]. This work defines a procedure developed and performed in air with optimised dipping and annealing times, leading to thick and compact films on flat substrates.

2. Materials and Methods

2.1. Materials

Bismuth(III) iodide (99%) and isopropyl alcohol (≥99.5%) were purchased from Sigma-Aldrich. Methylammonium iodide (98%) and ultra-flat quartz coated glass (2 × 1.5 cm) were purchased from Ossia Ltd. All salts and solvents were used as received without any further purification.

2.2. Synthesis of MABI Thin Films

The quartz-coated glasses were cleaned in an ultrasonic bath with IPA for 10 min. BiI3 was deposited by thermal evaporation at 10−6 mbar, with a deposition rate of 2 Å s−1 (crucible temperature of 170 °C). The thin films thus obtained were immersed in a 100 mL beaker containing 20 mL of the MAI solution in IPA (10 mg mL−1). Immersion times have been optimized during the work. Finally, the samples were annealed at 100 °C in a closed petri dish, set on a hot plate under a fume hood. Annealing times have been optimized during the work.

2.3. Characterization

Thin-film thickness was evaluated using a Bruker DektakXT profilometer, the standard deviation calculated on the average of 7 measurements at different points. The optical band gaps were evaluated by measuring the transmission spectra with a PerkinElmer Lambda 950 spectrometer equipped with an integrating sphere. The absorption coefficient (α) was calculated using the relation α(λ) = 1/t ln(1/T(λ)), where t is the sample thickness and T(λ) is the transmittance. The optical band gap was determined by plotting (αhν)2 versus hν and fitting the linear region of the absorption edge (OriginPro 2020b). X-ray diffraction (XRD) measurements were performed on a D5000 X-ray Powder Diffractometer (Siemens), using Cu-Kα radiation. Raman spectroscopy was performed using a Renishaw inVia confocal Raman microscope, equipped with a 785 nm laser; data analysis was performed by using OriginPro 2020b. X-ray photoelectron spectroscopy (XPS) measurements were performed with a Thermo Scientific™ Nexsa™ Surface Analysis System, and the XPS spectra were recorded and processed using the Thermo Avantage software. An FEI Inspect-F scanning electron microscope (SEM) was used for imaging at an accelerating voltage of 10 kV.

3. Results and Discussion

Different thicknesses of BiI3 and different dipping times were tested to obtain compact MABI films. The MAI solution in IPA was fixed at 10 mg mL−1 [23], and the BiI3 dipping was performed at room temperature. Annealing was performed on a hot plate at 100 °C for 90 min in air. The initial BiI3 film thickness and the BiI3 dipping time were varied, as summarized in Table 1. The main challenge of the method was the difficulty in obtaining a uniform transformation of the precursor layer. The homogeneity of the final MABI layer was evaluated by averaging the thicknesses measured with the profilometer. The standard deviation of the thickness was decisive in selecting the optimum growth conditions: with 300 nm—BiI3 films, thick layers were obtained with minimal deviations from the average. Figure 1 shows the Tauc plots of the produced films. The optical band gap was calculated from the linear fitting of the absorption edge. The extrapolated values agree with those reported in the literature: 1.90 eV for BiI3 film [24] and 1.80–1.82 eV for MABI [15]. The characteristic excitonic peak at 2.4 eV [15,24] becomes more or less evident according to the dipping time. A short immersion time (5 and 10 min) does not seem enough for complete conversion to the MABI structure; meanwhile, for immersion times longer than 15 min, the peak is well defined. Then, the subsequent characterisation was carried out to choose between 20 and 25 min.
The structures of the films dipped for 20 and 25 min belong to the hexagonal system with space group P63/mmc (XRD patterns in Figure 2a) [25,26]. Raman analysis can be used to probe changes in Bi-I bond stretching when MAI coordinates it. Figure 2b shows in dark brown the single stretching mode visible at 116 cm−1 for pristine BiI3. When it is coordinated in the MABI structure, in addition to the peak at 116 cm−1, the stretching mode at 146 cm−1 emerges [27]. Signals at 115–117 cm−1 and 146 cm−1 have been assigned to the (Bi-I) stretching mode, and the second peak appears when CH3NH3I is coordinated to the metal centre [3,27]. The area ratio of 116 cm−1 to 146 cm−1 components (76/24 for 20 min of dipping; 84/16 for 25 min) suggests that the maximum number of coordinated metal centres is reached for 20 min immersion. On this basis, 20 min was taken to be the optimal immersion time.
The oxidation states of the chemical species were investigated by XPS, and the high-resolution core-level spectra for N, I, C, and Bi are presented in Figure 3. The N 1s spectra are composed of a broad peak around 402 eV (Figure 3a), confirming the presence of free amine group NH2, bonded to C, and the N-C bond [28]. The peak around 619 eV belongs to I 3d5/2 (Figure 3b). In the sample dipped for 25 min, a shoulder appearing at higher energy can be related to iodine oxidation at the surface [28]. The C 1s spectrum (Figure 3c) has two components, one given by the C-C bond at 284.6 eV and the other by the C-N in MABI at 286.0 eV [28]. XPS Bi 4f7/2 core-level shows that metallic bismuth (peaks at 157.0 eV) is present (Figure 3d). The reduction of metallic bismuth has been reported to be a consequence of X-ray irradiation [29,30]. The most likely mechanism for this is analogous to reported Pb0 evolution from MAPbI3 under the XPS environment, that is X-ray induced photolysis of any metal halide salts present in the film (PbI2 in the reported case, or BiI3 in our case). The emergence of Bi0 can therefore be linked to unreacted BiI3 in the films. Therefore, the annealing time was varied to improve the MABI stability in a challenging environment.
Figure 4a shows the core-level spectra of Bi 4f7/2 after annealing at 100 °C for 90, 105, and 120 min: the sample treated for 105 min shows shallow Bi0 content. The shorter annealing time (90 min) probably leaves unreacted BiI3 in the film, which can be reduced to Bi0 under the X-ray beam. McGettrick et al. proved that in CH3NH3PbI3, the X-rays employed during the XPS measurement can trigger the photolysis of Pb2+ to Pb0, through light-induced generation of PbI2 [31]. The longer annealing time (120 min) results in a prominent Bi0 peak, which is likely due to MAI loss driving the formation of BiI3, which can then be reduced to Bi0 under the X-ray beam. In addition, the XRD pattern of the 120 min annealed sample, Figure 4b, shows that the peaks at 8.2°, 16.3°, 24.6° have reduced to almost background levels. Meanwhile, a broad peak appears at 9.7°, suggesting the growth of an amorphous phase. Comparing the XRD patterns of the annealed sample for 90 and 105 min (Figure 4b), the (1 0 1) and (2 0 2) peaks’ intensities in the spectrum of the sample annealed for 105 min are more significant, and, as the layer thickness is similar (875 ± 75 nm for 105 min; 850 ± 60 nm for 90 min), an enhanced crystallinity is possible and a more preferential alignment. This is confirmed by the SEM images (Figure 5): hexagonal crystals, which were not present in the sample annealed for 90 min, are evident in the film annealed for 105 min.

4. Conclusions

A two-step synthesis in air of methylammonium bismuth iodide thin-films is reported, using a method based on the conversion of a thermally evaporated bismuth iodide layer by immersion into an MAI solution. The critical process steps, namely the dipping time in a methylammonium iodide solution and the final annealing, were controlled to obtain a compact thin-film of about 1 µm thickness. The fine-tuning of the dipping and annealing times can allow the synthesis to be moved from an inert glove box environment to a fume hood. The described perovskite layer is suitable for use in thin-film technologies; therefore, the reported results may enable the use of bismuth-based perovskites in the next technological device generation, in which quantum phenomena are exploited.

Author Contributions

Conceptualization, V.T.; methodology, V.T.; software, V.T.; validation, O.F. and S.B.; formal analysis, V.T., S.L. and E.S.S.G.; investigation, V.T., S.L. and E.S.S.G.; resources, W.P.G., O.F. and S.B.; data curation, V.T.; writing—original draft preparation, V.T., G.T. and S.R.; writing—review and editing, V.T., G.T., S.R., W.P.G. and O.F.; visualization, V.T.; supervision, O.F. and S.B.; project administration, V.T. and O.F.; funding acquisition, V.T. and O.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the “European Union’s Horizon 2020 research and innovation programme under the Marie Skłodowska-Curie grant agreement, grant number 798271”, by the Royal Society University Research Fellowship UF/40372, by the Engineering and Physical Sciences Research Council (UK) under the Centre for Doctoral Training in Plastic Electronics (EP/L016702/1).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hu, H.; Dong, B.; Zhang, W. Low-toxic metal halide perovskites: Opportunities and future challenges. J. Mater. Chem. A 2017, 5, 11436–11449. [Google Scholar] [CrossRef]
  2. Jin, Z.; Zhang, Z.; Xiu, J.; Song, H.; Gatti, T.; He, Z. A critical review on bismuth and antimony halide based perovskites and their derivatives for photovoltaic applications: Recent advances and challenges. J. Mater. Chem. A 2020, 8, 16166–16188. [Google Scholar] [CrossRef]
  3. Trifiletti, V.; Asker, C.; Tseberlidis, G.; Riva, S.; Zhao, K.; Tang, W.; Binetti, S.; Fenwick, O. Quasi-zero dimensional halide perovskite derivates: Synthesis, status, and opportunity. Front. Electron. 2021, 2, 758603. [Google Scholar] [CrossRef]
  4. Chu, L. Pseudohalide anion engineering for highly efficient and stable perovskite solar cells. Matter 2021, 4, 1762–1764. [Google Scholar] [CrossRef]
  5. Chu, L.; Ahmad, W.; Liu, W.; Yang, J.; Zhang, R.; Sun, Y.; Yang, J.; Li, X. Lead-free halide double perovskite materials: A new superstar toward green and stable optoelectronic applications. Nano-Micro Lett. 2019, 11, 16. [Google Scholar] [CrossRef] [Green Version]
  6. Zhang, L.; Wang, K.; Zou, B. Bismuth halide perovskite-like materials: Current opportunities and challenges. ChemSusChem 2019, 12, 1612–1630. [Google Scholar] [CrossRef] [PubMed]
  7. Scholz, M.; Flender, O.; Oum, K.; Lenzer, T. Pronounced exciton dynamics in the vacancy-ordered bismuth halide perovskite (CH3NH3)3Bi2I9 observed by ultrafast uv–vis–nir transient absorption spectroscopy. J. Phys. Chem. C 2017, 121, 12110–12116. [Google Scholar] [CrossRef]
  8. Sarkar, A.; Acharyya, P.; Sasmal, R.; Pal, P.; Agasti, S.S.; Biswas, K. Synthesis of ultrathin few-layer 2d nanoplates of halide perovskite Cs3Bi2I9 and single-nanoplate super-resolved fluorescence microscopy. Inorg. Chem. 2018, 57, 15558–15565. [Google Scholar] [CrossRef] [PubMed]
  9. Ji, K.; Anaya, M.; Abfalterer, A.; Stranks, S.D. Halide perovskite light-emitting diode technologies. Adv. Opt. Mater. 2021, 9, 2002128. [Google Scholar] [CrossRef]
  10. Long, X.; Pan, Z.; Zhang, Z.; Urban, J.J.; Wang, H. Solvent-free synthesis of organometallic halides CH3NH3PbI3 and (CH3NH3)3Bi2I9 and their thermoelectric transport properties. Appl. Phys. Lett. 2019, 115, 072104. [Google Scholar] [CrossRef]
  11. Tie, S.; Zhao, W.; Xin, D.; Zhang, M.; Long, J.; Chen, Q.; Zheng, X.; Zhu, J.; Zhang, W.H. Robust fabrication of hybrid lead-free perovskite pellets for stable X-ray detectors with low detection limit. Adv. Mater. 2020, 32, e2001981. [Google Scholar] [CrossRef]
  12. Zhang, M.; Xin, D.; Zheng, X.; Chen, Q.; Zhang, W.-H. Toward greener solution processing of perovskite solar cells. ACS Sustain. Chem. Eng. 2020, 8, 13126–13138. [Google Scholar] [CrossRef]
  13. Park, B.W.; Philippe, B.; Zhang, X.; Rensmo, H.; Boschloo, G.; Johansson, E.M. Bismuth based hybrid perovskites A3Bi2I9 (a: Methylammonium or cesium) for solar cell application. Adv. Mater. 2015, 27, 6806–6813. [Google Scholar] [CrossRef] [PubMed]
  14. Singh, T.; Kulkarni, A.; Ikegami, M.; Miyasaka, T. Effect of electron transporting layer on bismuth-based lead-free perovskite (CH3NH3)3 Bi2I9 for photovoltaic applications. ACS Appl. Mater. Interfaces 2016, 8, 14542–14547. [Google Scholar] [CrossRef] [PubMed]
  15. Abulikemu, M.; Ould-Chikh, S.; Miao, X.; Alarousu, E.; Murali, B.; Ngongang Ndjawa, G.O.; Barbé, J.; El Labban, A.; Amassian, A.; Del Gobbo, S. Optoelectronic and photovoltaic properties of the air-stable organohalide semiconductor (CH3NH3)3Bi2I9. J. Mater. Chem. A 2016, 4, 12504–12515. [Google Scholar] [CrossRef] [Green Version]
  16. Chatterjee, S.; Payne, J.; Irvine, J.T.S.; Pal, A.J. Bandgap bowing in a zero-dimensional hybrid halide perovskite derivative: Spin–orbit coupling versus lattice strain. J. Mater. Chem. A 2020, 8, 4416–4427. [Google Scholar] [CrossRef]
  17. Ahmad, K.; Ansari, S.N.; Natarajan, K.; Mobin, S.M. A (CH3NH3)3Bi2I9 perovskite based on a two-step deposition method: Lead-free, highly stable, and with enhanced photovoltaic performance. ChemElectroChem 2019, 6, 1192–1198. [Google Scholar] [CrossRef]
  18. Baranwal, A.K.; Masutani, H.; Sugita, H.; Kanda, H.; Kanaya, S.; Shibayama, N.; Sanehira, Y.; Ikegami, M.; Numata, Y.; Yamada, K.; et al. Lead-free perovskite solar cells using sb and bi-based A3B2X9 and A3BX6 crystals with normal and inverse cell structures. Nano Converg. 2017, 4, 26. [Google Scholar] [CrossRef] [Green Version]
  19. Trifiletti, V.; Cannavale, A.; Listorti, A.; Rizzo, A.; Colella, S. Sequential deposition of hybrid halide perovskite starting both from lead iodide and lead chloride on the most widely employed substrates. Thin Solid Films 2018, 657, 110–117. [Google Scholar] [CrossRef]
  20. Hamdeh, U.H.; Nelson, R.D.; Ryan, B.J.; Bhattacharjee, U.; Petrich, J.W.; Panthani, M.G. Solution-processed bii3 thin films for photovoltaic applications: Improved carrier collection via solvent annealing. Chem. Mater. 2016, 28, 6567–6574. [Google Scholar] [CrossRef] [Green Version]
  21. Banik, A.; Bohannan, E.W.; Switzer, J.A. Epitaxial electrodeposition of bii3 and topotactic conversion to highly ordered solar light-absorbing perovskite (CH3NH3)3Bi2I9. Chem. Mater. 2020, 32, 8367–8372. [Google Scholar] [CrossRef]
  22. Coutinho, N.F.; Cucatti, S.; Merlo, R.B.; Silva Filho, J.M.C.; Villegas, N.F.B.; Alvarez, F.; Nogueira, A.F.; Marques, F.C. The thermomechanical properties of thermally evaporated bismuth triiodide thin films. Sci. Rep. 2019, 9, 11785. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Chen, M.; Wan, L.; Kong, M.; Hu, H.; Gan, Y.; Wang, J.; Chen, F.; Guo, Z.; Eder, D.; Wang, S. Influence of rutile-TiO2 nanorod arrays on pb-free (CH3NH3)3Bi2I9-based hybrid perovskite solar cells fabricated through two-step sequential solution process. J. Alloys Compd. 2018, 738, 422–431. [Google Scholar] [CrossRef]
  24. Ran, C.; Wu, Z.; Xi, J.; Yuan, F.; Dong, H.; Lei, T.; He, X.; Hou, X. Construction of compact methylammonium bismuth iodide film promoting lead-free inverted planar heterojunction organohalide solar cells with open-circuit voltage over 0.8 v. J. Phys. Chem. Lett. 2017, 8, 394–400. [Google Scholar] [CrossRef] [PubMed]
  25. Eckhardt, K.; Bon, V.; Getzschmann, J.; Grothe, J.; Wisser, F.M.; Kaskel, S. Crystallographic insights into (CH3NH3)3(Bi2I9): A new lead-free hybrid organic-inorganic material as a potential absorber for photovoltaics. Chem. Commun. 2016, 52, 3058–3060. [Google Scholar] [CrossRef] [Green Version]
  26. Lyu, M.; Yun, J.-H.; Cai, M.; Jiao, Y.; Bernhardt, P.V.; Zhang, M.; Wang, Q.; Du, A.; Wang, H.; Liu, G.; et al. Organic–inorganic bismuth (III)-based material: A lead-free, air-stable and solution-processable light-absorber beyond organolead perovskites. Nano Res. 2016, 9, 692–702. [Google Scholar] [CrossRef] [Green Version]
  27. Öz, S.; Hebig, J.-C.; Jung, E.; Singh, T.; Lepcha, A.; Olthof, S.; Jan, F.; Gao, Y.; German, R.; van Loosdrecht, P.H.M.; et al. Zero-dimensional (CH3NH3)3Bi2I9 perovskite for optoelectronic applications. Sol. Energy Mater. Sol. Cells 2016, 158, 195–201. [Google Scholar] [CrossRef]
  28. Bresolin, B.-M.; Hammouda, S.B.; Sillanpää, M. Methylammonium iodo bismuthate perovskite (CH3NH3)3Bi2I9 as new effective visible light-responsive photocatalyst for degradation of environment pollutants. J. Photochem. Photobiol. A Chem. 2019, 376, 116–126. [Google Scholar] [CrossRef]
  29. Hoefler, S.F.; Rath, T.; Fischer, R.; Latal, C.; Hippler, D.; Koliogiorgos, A.; Galanakis, I.; Bruno, A.; Fian, A.; Dimopoulos, T.; et al. A zero-dimensional mixed-anion hybrid halogenobismuthate(III) semiconductor: Structural, optical, and photovoltaic properties. Inorg. Chem. 2018, 57, 10576–10586. [Google Scholar] [CrossRef]
  30. Tang, W.; Zhang, J.; Ratnasingham, S.; Liscio, F.; Chen, K.; Liu, T.; Wan, K.; Galindez, E.S.; Bilotti, E.; Reece, M.; et al. Substitutional doping of hybrid organic–inorganic perovskite crystals for thermoelectrics. J. Mater. Chem. A 2020, 8, 13594–13599. [Google Scholar] [CrossRef]
  31. McGettrick, J.D.; Hooper, K.; Pockett, A.; Baker, J.; Troughton, J.; Carnie, M.; Watson, T. Sources of Pb(0) artefacts during xps analysis of lead halide perovskites. Mater. Lett. 2019, 251, 98–101. [Google Scholar] [CrossRef]
Figure 1. Tauc plots of the 300 nm BiI3 layer, and MABI films obtained after 5, 10, 15, 20, and 25 min of dipping in the MAI solution in IPA.
Figure 1. Tauc plots of the 300 nm BiI3 layer, and MABI films obtained after 5, 10, 15, 20, and 25 min of dipping in the MAI solution in IPA.
Materials 14 07827 g001
Figure 2. XRD patterns (a) and Raman spectra (b) of the 300 nm BiI3 layer, and MABI films obtained after 20 and 25 min dipping in the MAI solution in IPA.
Figure 2. XRD patterns (a) and Raman spectra (b) of the 300 nm BiI3 layer, and MABI films obtained after 20 and 25 min dipping in the MAI solution in IPA.
Materials 14 07827 g002
Figure 3. XPS high-resolution core-level spectra, related to the MABI films obtained after 20 and 25 min dipping in the MAI solution in IPA, for (a) N 1s; (b) I 3d5/2; (c) C 1s; (d) Bi 4f7/2.
Figure 3. XPS high-resolution core-level spectra, related to the MABI films obtained after 20 and 25 min dipping in the MAI solution in IPA, for (a) N 1s; (b) I 3d5/2; (c) C 1s; (d) Bi 4f7/2.
Materials 14 07827 g003
Figure 4. (a) XPS high-resolution core-level spectra for Bi 4f7/2, recorded for the MABI films obtained after 20 dipping in the MAI solution in IPA, and annealed at 100 °C for 90, 105, 120 min in air; (b) XRD patterns of the MABI films obtained after 20 dipping in the MAI solution in IPA, and annealed at 100 °C for 90, 105, 120 min in air.
Figure 4. (a) XPS high-resolution core-level spectra for Bi 4f7/2, recorded for the MABI films obtained after 20 dipping in the MAI solution in IPA, and annealed at 100 °C for 90, 105, 120 min in air; (b) XRD patterns of the MABI films obtained after 20 dipping in the MAI solution in IPA, and annealed at 100 °C for 90, 105, 120 min in air.
Materials 14 07827 g004
Figure 5. SEM images of (a) 300 nm BiI3 layer; (b) MABI film obtained with immersion time in CH3NH3I solution of 20 min and annealed at 100 °C for 90 min in air; (c) MABI film obtained with immersion time in CH3NH3I solution of 20 min and annealed at 100 °C for 105 min in air. Scale: 20 μm top and 2 μm bottom images.
Figure 5. SEM images of (a) 300 nm BiI3 layer; (b) MABI film obtained with immersion time in CH3NH3I solution of 20 min and annealed at 100 °C for 90 min in air; (c) MABI film obtained with immersion time in CH3NH3I solution of 20 min and annealed at 100 °C for 105 min in air. Scale: 20 μm top and 2 μm bottom images.
Materials 14 07827 g005
Table 1. BiI3 film thickness, dipping times in MAI in IPA solution, and the obtained MABI film thicknesses.
Table 1. BiI3 film thickness, dipping times in MAI in IPA solution, and the obtained MABI film thicknesses.
BiI3—Thickness
(nm)
Deviation
(%)
Dipping Time
(min)
MABI—Thickness
(nm)
Deviation
(%)
50251008
502101408
502151809
502202509
502252759
1002515010
10021020010
10021526010
10022027510
10022535010
250355009
2503105509
25031570010
25032080010
25032585010
300356008
3003107007
3003158007
3003208507
3003259008
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Trifiletti, V.; Luong, S.; Tseberlidis, G.; Riva, S.; Galindez, E.S.S.; Gillin, W.P.; Binetti, S.; Fenwick, O. Two-Step Synthesis of Bismuth-Based Hybrid Halide Perovskite Thin-Films. Materials 2021, 14, 7827. https://doi.org/10.3390/ma14247827

AMA Style

Trifiletti V, Luong S, Tseberlidis G, Riva S, Galindez ESS, Gillin WP, Binetti S, Fenwick O. Two-Step Synthesis of Bismuth-Based Hybrid Halide Perovskite Thin-Films. Materials. 2021; 14(24):7827. https://doi.org/10.3390/ma14247827

Chicago/Turabian Style

Trifiletti, Vanira, Sally Luong, Giorgio Tseberlidis, Stefania Riva, Eugenio S. S. Galindez, William P. Gillin, Simona Binetti, and Oliver Fenwick. 2021. "Two-Step Synthesis of Bismuth-Based Hybrid Halide Perovskite Thin-Films" Materials 14, no. 24: 7827. https://doi.org/10.3390/ma14247827

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