Next Article in Journal
Study on Microstructure and Mechanical Properties at Constant Electromigration Temperature of Sn2.5Ag0.7Cu0.1RE0.05Ni-GNSs/Cu Solder Joints
Previous Article in Journal
A Time-Domain Signal Processing Algorithm for Data-Driven Drive-by Inspection Methods: An Experimental Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Growth of Hybrid Perovskite Crystals from CH3NH3PbI3–xClx Solutions Subjected to Constant Solvent Evaporation Rates

1
Faculty of Physics, Babes-Bolyai University, M. Kogalniceanu Str. 1, 400084 Cluj-Napoca, Romania
2
Interdisciplinary Research Institute in Bio-Nano-Sciences, Babes-Bolyai University, Treboniu Laurian 42, 400271 Cluj-Napoca, Romania
3
INCDO-INOE 2000, Research Institute for Analytical Instrumentation, Donath Street 67, 400293 Cluj-Napoca, Romania
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Materials 2023, 16(7), 2625; https://doi.org/10.3390/ma16072625
Submission received: 8 March 2023 / Revised: 22 March 2023 / Accepted: 25 March 2023 / Published: 26 March 2023

Abstract

:
In this work, we subjected hybrid lead-mixed halide perovskite (CH3NH3PbI3–xClx) precursor inks to different solvent evaporation rates in order to facilitate the nucleation and growth of perovskite crystals. By controlling the temperature of perovskite solutions placed within open-air rings in precise volumes, we established control over the rate of solvent evaporation and, thus, over both the growth rate and the shape of perovskite crystals. Direct utilization of diluted lead-mixed halide perovskites solutions allowed us to control the nucleation and to favor the growth of only a low number of perovskite crystals. Such crystals exhibited a clear sixfold symmetry. While crystals formed at a lower range of temperatures (40–60 °C) exhibited a more compact dendritic shape, the crystals grown at a higher temperature range (80–110 °C) displayed a fractal dendritic morphology.

1. Introduction

In recent years, perovskite-based solar cells have reached a certified power conversion efficiency of up to 25.7% [1] due to the optimization of crystalline perovskite microstructure [2,3,4]. Moreover, by altering the crystallization kinetics of quasi-2D perovskites, the external quantum efficiency of light-emitting diodes based on such materials has been shown to exceed 18.15% [5,6,7]. Thus, in order to obtain the most suitable perovskite materials for various energy devices, it is important to control the microstructure of these materials, which is now possible in an ambient environment by utilizing simple deposition methods such as spin coating [8,9], drop casting [10,11,12] or dip coating [13,14,15]. Other scalable deposition methods used to fabricate perovskite active layers of specific crystalline microstructure include spray coating [16,17,18], slot-die coating [19], blade coating [20,21,22], D-bar coating [23,24] and various roll-to-roll processes [25]. However, such rapid deposition methods are not ideal to produce model perovskite structures that could be precisely evaluated with respect to their optoelectronic properties. Moreover, a slower method for the generation of perovskite-based microstructures seems to lead to thin crystalline films exhibiting higher quality and improved properties [26,27,28]. Although various perovskite processing procedures have been designed and developed [29,30,31,32], it is still challenging to fully understand and precisely control the mechanisms of solid-phase nucleation and crystal growth from various perovskite solutions. Nonetheless, this aspect is highly important because perovskite crystals have been shown to display important advantages, such as enhanced optoelectronic properties (high absorption coefficient, extended diffusion length, improved charge carrier mobility, etc.) and stability [33,34,35,36]. Such qualities make them highly suitable for applications like solar cells, light-emitting diodes, photodetectors, lasers, etc.
There are many different methods to obtain crystals of perovskites [37,38,39]. For instance, such methods may rely on cavitation-triggered asymmetrical crystallization [40] or on the utilization of space confinements [41]. While the former method leads to monocrystalline perovskite films ideal for solar cells [40], the latter provides important possibilities for mass production of ultrathin crystalline wafers [41]. Films comprised of perovskite crystals can also be obtained by employing the vapor-phase epitaxial growth [42] or surface-tension-assisted growth [43] methods. Other approaches to generate perovskite crystals are based on reduced solution temperature [44,45], inverse temperature crystallization [46,47,48], meniscus-assisted solution printing [49], anti-solvent vapor-assisted crystallization [34,50,51,52], solvent acidolysis crystallization [53], top-seeded solution growth under an ambient atmosphere [54] and many other methods. Resulting (organic–inorganic hybrid) perovskite crystals can be used for bandgap engineering [34]. Other efficient methods to obtain perovskite crystals may count on the utilization of polymers to control the nucleation process [55] or the use of single-walled carbon nanotube-based additives to retard crystal growth [56]. Furthermore, silk fibroin ca be employed as crystal growth template to align the perovskite crystals, thereby massively improving their electron mobility [57], while the usage of printed hydrophilic–hydrophobic substrates can lead to large-scale perovskite single-crystal arrays for photodetectors and other optoelectronic devices [58].
In this paper, we aim to establish a simple practical method to produce well-defined perovskite crystals that can be used in photovoltaic or emission energy devices. To do so, we utilize perovskite precursor ink solutions of precise volumes. The latter are placed into open-air Teflon rings and heated to a desired temperature, while kept under direct, real-time observation using an optical microscope (OM). By controlling the solution temperature, we dictate a specific yet constant solvent evaporation rate that leads to an increase in perovskite concentration over time, consequently initiating the nucleation and growth of lead-mixed halide perovskite crystals. By employing different solution temperatures, we further control not only the growth rate of perovskite crystals but also their shape and morphology.

2. Materials and Methods

I101 perovskite precursor ink was purchased from Ossila Ltd. (Sheffield, UK). Such systems have previously been used as efficient solar cell materials [59,60,61] and were produced by the manufacturer by dissolving methylammonium iodide (MAI) and lead chloride (PbCl2) in dimethylformamide (DMF) in a 3:1 ratio [62,63,64]. Further exposure to heat converted I101 perovskite to methylammonium lead halide perovskite (CH3NH3PbI3–xClx). Because MAI:PbCl2 was used as a precursor material, air processing in a low-humidity environment (20% to 35%) was possible.
The following experimental setup was used to study perovskite crystal formation. A Teflon ring with a 3 mm diameter and a 2 mm height was mounted and glued on a standard microscope glass slide. The setup was further treated in a UV ozone cleaner and placed on a Linkam THMS600 hot stage under an OM. The temperature of the hot stage could be adjusted and controlled with high precision between 40 °C and 110 °C (the hot stage is capable of regulating the temperature from 600 °C down to −196 °C when liquid nitrogen is utilized). Finally, 5 µL of the lead-mixed halide perovskite solution was injected in the Teflon rings and evenly spread over the glass substrate, leading to a solution of a thickness of about 700 µm (this value was deduced by considering that a 5 µL of perovskite solution represents a volume of about 5 mm3 and that the diameter of a used Teflon ring is 3 mm). A KERN OKN-177 OM with a 20× objective was used to observe the growth of perovskite crystals in real time and direct space (while maintaining the hot plate at a chosen constant temperature). The whole experimental setup is schematically represented in Figure 1. The solvent evaporation rate was mainly dictated by the processing temperature of the perovskite sample. Nonetheless, other physical parameters, such as the temperature of the perovskite solution and the room temperature and humidity, needed to be to controlled in order to perform comparable and reproducible experiments. To that end, the room temperature was regulated at 24 °C at all time via an air conditioning system, while the perovskite solutions were kept in the fridge at 5 °C prior to use.
A typical perovskite crystal growth experiment is described as follows. A cleaned Teflon ring glued on the glass substrate is positioned on the hot stage. The latter is then placed under the OM. Then, a specific sample processing temperature is set on the hot stage controller, while the OM objective is focused on the glass substrate. Once the substrate reaches the desired temperature, the perovskite solution is placed within the Teflon ring using a micropipette. The solution spreads instantly all over the surface. Then, the OM objective is slowly retracted to focus on the surface of the perovskite solution (or slightly within this solution yet in the vicinity of the surface). At this point, the homogeneous surface of the solution is continuously monitored. Generally, any inhomogeneity appearing after a certain time on the surface indicates the growth of a tiny perovskite crystal. Then, the objective of the OM is aligned so that the surface inhomogeneity is in the middle of the field of view and further focused so that the inhomogeneity is clearly visible. Furthermore, by taking advantage of the camera software installed on the PC, the camera can be programmed to record an optical micrograph (or a movie) at a specific distance in time. The collected micrographs are then analyzed and used to determine the growth rates of the corresponding crystal. Note that the OM objective that is chosen at the beginning of the experiment should be a low magnification (either 10× or 20×) to make it easier to find a tiny crystal/inhomogeneity. Later, when monitoring the growth of a given perovskite crystal, the objective can be switched to one capable of a higher magnification and better spatial resolution.

3. Results and Discussion

Figure 2 shows optical micrographs of the CH3NH3PbI3–xClx hybrid crystals nucleated and grown at various temperatures from a lead-mixed halide perovskite precursor solution. At each selected processing temperature, our open-air experimental setup (see Figure 1) permitted us to maintain a corresponding solvent evaporation rate. The latter dictated not only the type of perovskite crystals that would be obtained but also how fast such crystals would grow, i.e., how flawless they would become from a structural point of view. We based our strategy on the well-known correlations between the evaporation and crystallization rates that occur when processing perovskite materials [65,66]. Although other variants of perovskite processing methods based on solvent evaporation were previously reported in the literature [67,68,69], the advantage of our setup lies in the possibility of nucleating only a small number of crystals, which can be grown independent of each other (note that perovskite crystals start “coalescing” only at very late stages of crystallization, when their size increases). As a result, hybrid CH3NH3PbI3–xClx perovskite crystals displaying a sixfold symmetry, represented by the six “tree” branches, nucleated and grew at a processing temperature of 110 °C (Figure 2a–d). Their rather high growth rate, determined to be over 9.7 µm/s (Figure 2e), favored the development of fractal dendrites, i.e., structures that exhibited a density that was substantially lower than that of an “ideal” faceted crystal [70,71]. Moreover, an almost hexagonal envelope formed from the six dendritic “trees” grown along the diagonal directions was observed. The crystal growth rates were generally determined (i) by measuring the distances from the center of a specific crystal to one if its growing tips/trees (the latter are indicated by dotted arrows in Figure 2) at different times after a given time (t0), (ii) by fitting the obtained data points with linear fits and (iii) by identifying the slopes of the linear fits as the corresponding average growth rates.
Furthermore, additional experiments performed at processing sample temperatures higher than 110 °C emphasized that under such conditions, the perovskite solutions were rather degrading, with some inhomogeneous (most often spherical) objects or structures appearing randomly all over the surface. Sometimes, crystals of ill-defined symmetry exhibiting a rather “melted/degraded” appearance were observed to form. More often, only some random objects, possibly irregular aggregates or very poor-quality crystalline structures, were observed on the surface after the total evaporation of the solvent.
At an even lower temperature of 80 °C, similar perovskite crystals also developed (Figure 2f–i). In this case, the only measurable difference was represented by a fewfold lower crystal growth rate (determined to be ~3 µm/s, Figure 2j). Otherwise, the fractal dendritic morphology was represented again by six dendritic “trees” forming a hexagonal envelope. When the temperature was lowered to 60 °C, thereby drastically reducing the solvent evaporation rate, perovskite crystals exhibiting a more compact (i.e., closer to the density of an “ideal” faceted crystal) dendritic shape began to nucleate and grow (Figure 2k–n). Such perfect snowflake-like morphology [72] developed far more slowly, with a growth rate of 0.64 µm/s (Figure 2o). The six dendritic “trees” growing along the diagonal direction with significantly less, almost symmetrical, fractal-like branching, emphasize how close to perfection the sixfold symmetry was in this instance.
Hybrid CH3NH3PbI3–xClx perovskite crystals displaying a sixfold symmetry and a rather compact dendritic structure reminiscent of simple star snowflakes [73] were further grown when a much lower rate of solvent evaporation was employed (Figure 2p–s). Thus, a growth rate of only 0.19 µm/s (Figure 2t) was generated when the solution was annealed at a lower temperature of 40 °C. Here, the branching of all six “trees” was much less prominent. The branching started to become more prominent along one of the six “trees” after a growing time of t0 + 150 s, when the crystal front along this particular “tree” was more prone to become favored by the nearby available perovskite molecules. Therefore, it developed faster by eliminating the competition from the other two “trees” in its vicinity, i.e., by achieving the possibility of more quickly accessing many more available perovskite molecules.
The sixfold symmetry observed for all the aforementioned CH3NH3PbI3–xClx perovskite crystals is a result of a predominant growth along the diagonal directions, with the growing tips closer to the reservoir of available perovskite molecules that could be “captured” and included in the crystal with a higher probability [74,75]. Moreover, the decrease in crystal growth rates observed with a decreased sample temperature can be explained by considering the finite amount of perovskite molecules available in the initial volume of 5 µL of solution and by assuming that at a lower processing temperature, the solvent evaporation rate is low and, consequently, induces only a minimal increase in the concentration of perovskite molecules over time. Thus, because in this case, the number of available molecules ready to attach to a crystal was low, the attachment probability was similarly low. Perovskite molecules had time to attach, detach and reattach until they reached an optimum position within the developing crystal, ultimately leading to more compact (i.e., less branched) crystals. On the other hand, using a high processing temperature causes a high rate of solvent evaporation, which, in turn, rapidly increases the concentration of perovskite molecules in the solution. Consequently, the number of available molecules ready to attach to a crystal “explodes”, along with the attachment probability. Therefore, perovskite molecules attach quickly within the growing crystal, leading to more branched perovskite crystals.
Interestingly, besides the perovskite crystals exhibiting a sixfold symmetry, other “X”-like crystals displaying a fourfold symmetry were also observed (Figure 3). This only occurred in the temperature processing range of 50 °C to 90 °C, and no “X”-like crystals were observed outside this range. While “X”-like crystals were observed to nucleate and grow independently of each other, their branching was rather limited, with more branching visible in crystals grown at 80 °C (Figure 3e–h). Furthermore, because the crystal branching seems to become more visible at later crystal growth stages, when, very often, crystals undergo a coalescence process, a direct correlation between the processing temperature and the degree of crystal branching is not yet possible to establish. Additionally, as determined from Figure 3, the crystal growth rates increased with the increase in processing temperature from ~2.24 µm/s (Figure 3q–t) to ~3.15 µm/s (Figure 3m–p), ~4.73 µm/s (Figure 3i–l), ~5.73 µm/s (Figure 3e–h) and ~6.69 µm/s (Figure 3a–d). These rates, despite the increased processing temperature, are much higher than the growth rates corresponding to the crystals exhibiting sixfold symmetry (shown in Figure 2e,j,o,t). For instance, while a fourfold “X”-like crystal grew at a rate of ~3.15 µm/s at 60 °C, a sixfold crystal grew at a rate of only 0.64 µm/s in the same sample and at the same temperature. Similarly, at 80 °C, the corresponding growth rates were ~5.73 µm/s and ~3 µm/s for the fourfold and sixfold symmetrical perovskite crystals, respectively. This suggests that the sixfold symmetrical crystals are of higher quality and could possibly exhibit puzzling optoelectronic properties, assuming their slower growth and the already demonstrated beneficial effects of the crystalline microstructure of mixed halide perovskites on their optical and photovoltaic properties [76]. To prove this tentative statement, future experiments are necessary to assess the internal structure of such crystals.

4. Conclusions

In this study, we grew crystals of an organic–inorganic halide perovskite precursor solution at different processing temperatures using an open-air setup that allowed us to change the solvent evaporation rate and to nucleate only a small number of crystals. We demonstrated that perovskite crystals grown at lower temperatures and, therefore, at lower solvent evaporation rates, displayed a clear sixfold symmetry with a shape reminiscent of rather compact dendrites. In contrast, six-branched crystals exhibiting a fractal dendritic morphology were obtained at higher temperatures. Fourfold symmetrical “X”-like crystals were also observed to nucleate and grow in the 50–90 °C temperature range but at higher growth rates as compared to their sixfold symmetry counterparts. This work is useful for providing model crystalline structures that can be employed in the future to correlate the optoelectronic properties of certain perovskites with their microstructure.

Author Contributions

I.P. and O.T.-B. contributed equally to this work by designing the concept of the paper, conducting the experimental investigation and writing the original draft. I.B. and L.D. supervised I.P. and O.T.-B. and edited the final manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was carried out through the Core Program within the National Research Development and Innovation Plan 2022–2027, carried out with the support of MCID, project no. PN 23 05.

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. Min, H.; Lee, D.Y.; Kim, J.; Kim, G.; Lee, K.S.; Kim, J.; Paik, M.J.; Kim, Y.K.; Kim, K.S.; Kim, M.G.; et al. Perovskite Solar Cells with Atomically Coherent Interlayers on SnO2 Electrodes. Nature 2021, 598, 444–450. [Google Scholar] [CrossRef] [PubMed]
  2. Xiong, Z.; Chen, X.; Zhang, B.; Odunmbaku, G.O.; Ou, Z.; Guo, B.; Yang, K.; Kan, Z.; Lu, S.; Chen, S.; et al. Simultaneous Interfacial Modification and Crystallization Control by Biguanide Hydrochloride for Stable Perovskite Solar Cells with PCE of 24.4%. Adv. Mater. 2022, 34, 2106118. [Google Scholar] [CrossRef] [PubMed]
  3. Sharma, R.; Sharma, A.; Agarwal, S.; Dhaka, M.S. Stability and Efficiency Issues, Solutions and Advancements in Perovskite Solar Cells: A Review. Sol. Energy 2022, 244, 516–535. [Google Scholar] [CrossRef]
  4. Lee, D.-K.; Park, N.-G. Materials and Methods for High-Efficiency Perovskite Solar Modules. Sol. RRL 2022, 6, 2100455. [Google Scholar] [CrossRef]
  5. Guo, Z.; Zhang, Y.; Wang, B.; Wang, L.; Zhou, N.; Qiu, Z.; Li, N.; Chen, Y.; Zhu, C.; Xie, H.; et al. Promoting Energy Transfer via Manipulation of Crystallization Kinetics of Quasi-2D Perovskites for Efficient Green Light-Emitting Diodes. Adv. Mater. 2021, 33, 2102246. [Google Scholar] [CrossRef]
  6. Ai, B.; Fan, Z.; Wong, Z.J. Plasmonic–Perovskite Solar Cells, Light Emitters, and Sensors. Microsyst. Nanoeng. 2022, 8, 5. [Google Scholar] [CrossRef] [PubMed]
  7. Rakshit, S.; Piatkowski, P.; Mora-Seró, I.; Douhal, A. Combining Perovskites and Quantum Dots: Synthesis, Characterization, and Applications in Solar Cells, LEDs, and Photodetectors. Adv. Opt. Mater. 2022, 10, 2102566. [Google Scholar] [CrossRef]
  8. Xiao, M.; Huang, F.; Huang, W.; Dkhissi, Y.; Zhu, Y.; Etheridge, J.; Gray-Weale, A.; Bach, U.; Cheng, Y.-B.; Spiccia, L. A Fast Deposition-Crystallization Procedure for Highly Efficient Lead Iodide Perovskite Thin-Film Solar Cells. Angew. Chem. Int. Ed. 2014, 53, 9898–9903. [Google Scholar] [CrossRef]
  9. Liu, X.; Tan, X.; Liu, Z.; Ye, H.; Sun, B.; Shi, T.; Tang, Z.; Liao, G. Boosting the Efficiency of Carbon-Based Planar CsPbBr3 Perovskite Solar Cells by a Modified Multistep Spin-Coating Technique and Interface Engineering. Nano Energy 2019, 56, 184–195. [Google Scholar] [CrossRef]
  10. Zuo, C.; Ding, L. Drop-Casting to Make Efficient Perovskite Solar Cells under High Humidity. Angew. Chem. Int. Ed. 2021, 60, 11242–11246. [Google Scholar] [CrossRef] [PubMed]
  11. Zuo, C.; Scully, A.D.; Gao, M. Drop-Casting Method to Screen Ruddlesden–Popper Perovskite Formulations for Use in Solar Cells. ACS Appl. Mater. Interfaces 2021, 13, 56217–56225. [Google Scholar] [CrossRef]
  12. Zuo, C.; Scully, A.D.; Tan, W.L.; Zheng, F.; Ghiggino, K.P.; Vak, D.; Weerasinghe, H.; McNeill, C.R.; Angmo, D.; Chesman, A.S.R.; et al. Crystallisation Control of Drop-Cast Quasi-2D/3D Perovskite Layers for Efficient Solar Cells. Commun. Mater. 2020, 1, 33. [Google Scholar] [CrossRef]
  13. Adnan, M.; Lee, J.K. All Sequential Dip-Coating Processed Perovskite Layers from an Aqueous Lead Precursor for High Efficiency Perovskite Solar Cells. Sci. Rep. 2018, 8, 2168. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Adnan, M.; Irshad, Z.; Lee, J.K. Facile All-Dip-Coating Deposition of Highly Efficient (CH3)3NPbI3−xClx Perovskite Materials from Aqueous Non-Halide Lead Precursor. RSC Adv. 2020, 10, 29010–29017. [Google Scholar] [CrossRef] [PubMed]
  15. Irshad, Z.; Adnan, M.; Lee, J.K. Simple Preparation of Highly Efficient MAxFA1−xPbI3 Perovskite Films from an Aqueous Halide-Free Lead Precursor by All Dip-Coating Approach and Application in High-Performance Perovskite Solar Cells. J. Mater. Sci. 2022, 57, 1936–1946. [Google Scholar] [CrossRef]
  16. Barrows, A.T.; Pearson, A.J.; Kwak, C.K.; Dunbar, A.D.F.; Buckley, A.R.; Lidzey, D.G. Efficient Planar Heterojunction Mixed-Halide Perovskite Solar Cells Deposited via Spray-Deposition. Energy Environ. Sci. 2014, 7, 2944–2950. [Google Scholar] [CrossRef]
  17. Bishop, J.E.; Smith, J.A.; Lidzey, D.G. Development of Spray-Coated Perovskite Solar Cells. ACS Appl. Mater. Interfaces 2020, 12, 48237–48245. [Google Scholar] [CrossRef]
  18. Das, S.; Yang, B.; Gu, G.; Joshi, P.C.; Ivanov, I.N.; Rouleau, C.M.; Aytug, T.; Geohegan, D.B.; Xiao, K. High-Performance Flexible Perovskite Solar Cells by Using a Combination of Ultrasonic Spray-Coating and Low Thermal Budget Photonic Curing. ACS Photonics 2015, 2, 680–686. [Google Scholar] [CrossRef]
  19. Schmidt, T.M.; Larsen-Olsen, T.T.; Carlé, J.E.; Angmo, D.; Krebs, F.C. Upscaling of Perovskite Solar Cells: Fully Ambient Roll Processing of Flexible Perovskite Solar Cells with Printed Back Electrodes. Adv. Energy Mater. 2015, 5, 1500569. [Google Scholar] [CrossRef]
  20. Yang, Z.; Chueh, C.-C.; Zuo, F.; Kim, J.H.; Liang, P.-W.; Jen, A.K.-Y. High-Performance Fully Printable Perovskite Solar Cells via Blade-Coating Technique under the Ambient Condition. Adv. Energy Mater. 2015, 5, 1500328. [Google Scholar] [CrossRef]
  21. 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]
  22. Yang, J.; Lim, E.L.; Tan, L.; Wei, Z. Ink Engineering in Blade-Coating Large-Area Perovskite Solar Cells. Adv. Energy Mater. 2022, 12, 2200975. [Google Scholar] [CrossRef]
  23. Jeong, D.-N.; Lee, D.-K.; Seo, S.; Lim, S.Y.; Zhang, Y.; Shin, H.; Cheong, H.; Park, N.-G. Perovskite Cluster-Containing Solution for Scalable D-Bar Coating toward High-Throughput Perovskite Solar Cells. ACS Energy Lett. 2019, 4, 1189–1195. [Google Scholar] [CrossRef]
  24. Lim, K.-S.; Lee, D.-K.; Lee, J.-W.; Park, N.-G. 17% Efficient Perovskite Solar Mini-Module via Hexamethylphosphoramide (HMPA)-Adduct-Based Large-Area D-Bar Coating. J. Mater. Chem. A 2020, 8, 9345–9354. [Google Scholar] [CrossRef]
  25. Troughton, J.; Bryant, D.; Wojciechowski, K.; Carnie, M.J.; Snaith, H.; Worsley, D.A.; Watson, T.M. Highly Efficient, Flexible, Indium-Free Perovskite Solar Cells Employing Metallic Substrates. J. Mater. Chem. A 2015, 3, 9141–9145. [Google Scholar] [CrossRef]
  26. Liu, C.; Cheng, Y.-B.; Ge, Z. Understanding of Perovskite Crystal Growth and Film Formation in Scalable Deposition Processes. Chem. Soc. Rev. 2020, 49, 1653–1687. [Google Scholar] [CrossRef]
  27. Chen, W.; Li, X.; Li, Y.; Li, Y. A Review: Crystal Growth for High-Performance All-Inorganic Perovskite Solar Cells. Energy Environ. Sci. 2020, 13, 1971–1996. [Google Scholar] [CrossRef]
  28. Xiao, Z.; Dong, Q.; Bi, C.; Shao, Y.; Yuan, Y.; Huang, J. Solvent Annealing of Perovskite-Induced Crystal Growth for Photovoltaic-Device Efficiency Enhancement. Adv. Mater. 2014, 26, 6503–6509. [Google Scholar] [CrossRef]
  29. Du, P.; Wang, L.; Li, J.; Luo, J.; Ma, Y.; Tang, J.; Zhai, T. Thermal Evaporation for Halide Perovskite Optoelectronics: Fundamentals, Progress, and Outlook. Adv. Opt. Mater. 2022, 10, 2101770. [Google Scholar] [CrossRef]
  30. Sánchez, S.; Vallés-Pelarda, M.; Alberola-Borràs, J.-A.; Vidal, R.; Jerónimo-Rendón, J.J.; Saliba, M.; Boix, P.P.; Mora-Seró, I. Flash Infrared Annealing as a Cost-Effective and Low Environmental Impact Processing Method for Planar Perovskite Solar Cells. Mater. Today 2019, 31, 39–46. [Google Scholar] [CrossRef]
  31. Li, X.; Bi, D.; Yi, C.; Décoppet, J.-D.; Luo, J.; Zakeeruddin, S.M.; Hagfeldt, A.; Grätzel, M. A Vacuum Flash–Assisted Solution Process for High-Efficiency Large-Area Perovskite Solar Cells. Science 2016, 353, 58–62. [Google Scholar] [CrossRef] [PubMed]
  32. Wu, G.; Li, X.; Zhou, J.; Zhang, J.; Zhang, X.; Leng, X.; Wang, P.; Chen, M.; Zhang, D.; Zhao, K.; et al. Fine Multi-Phase Alignments in 2D Perovskite Solar Cells with Efficiency over 17% via Slow Post-Annealing. Adv. Mater. 2019, 31, 1903889. [Google Scholar] [CrossRef] [PubMed]
  33. Cheng, X.; Yang, S.; Cao, B.; Tao, X.; Chen, Z. Single Crystal Perovskite Solar Cells: Development and Perspectives. Adv. Funct. Mater. 2020, 30, 1905021. [Google Scholar] [CrossRef]
  34. Zuo, T.; He, X.; Hu, P.; Jiang, H. Organic-Inorganic Hybrid Perovskite Single Crystals: Crystallization, Molecular Structures, and Bandgap Engineering. ChemNanoMat 2019, 5, 278–289. [Google Scholar] [CrossRef]
  35. Singh, R.; Parashar, M.; Sandhu, S.; Yoo, K.; Lee, J.-J. The Effects of Crystal Structure on the Photovoltaic Performance of Perovskite Solar Cells under Ambient Indoor Illumination. Sol. Energy 2021, 220, 43–50. [Google Scholar] [CrossRef]
  36. Cho, Y.; Jung, H.R.; Jo, W. Halide Perovskite Single Crystals: Growth, Characterization, and Stability for Optoelectronic Applications. Nanoscale 2022, 14, 9248–9277. [Google Scholar] [CrossRef]
  37. Xu, J.; Ma, J.; Gu, Y.; Li, Y.; Li, Y.; Shen, H.; Zhang, Z.; Ma, Y. Progress of Metal Halide Perovskite Crystals From a Crystal Growth Point of View. Cryst. Res. Technol. 2023, 58, 2200128. [Google Scholar] [CrossRef]
  38. Haque, M.A.; Troughton, J.; Baran, D. Processing-Performance Evolution of Perovskite Solar Cells: From Large Grain Polycrystalline Films to Single Crystals. Adv. Energy Mater. 2020, 10, 1902762. [Google Scholar] [CrossRef]
  39. Gao, Q.; Qi, J.; Chen, K.; Xia, M.; Hu, Y.; Mei, A.; Han, H. Halide Perovskite Crystallization Processes and Methods in Nanocrystals, Single Crystals, and Thin Films. Adv. Mater. 2022, 34, 2200720. [Google Scholar] [CrossRef]
  40. Peng, W.; Wang, L.; Murali, B.; Ho, K.-T.; Bera, A.; Cho, N.; Kang, C.-F.; Burlakov, V.M.; Pan, J.; Sinatra, L.; et al. Solution-Grown Monocrystalline Hybrid Perovskite Films for Hole-Transporter-Free Solar Cells. Adv. Mater. 2016, 28, 3383–3390. [Google Scholar] [CrossRef] [Green Version]
  41. Liu, Y.; Zhang, Y.; Yang, Z.; Yang, D.; Ren, X.; Pang, L.; Liu, S. Thinness- and Shape-Controlled Growth for Ultrathin Single-Crystalline Perovskite Wafers for Mass Production of Superior Photoelectronic Devices. Adv. Mater. 2016, 28, 9204–9209. [Google Scholar] [CrossRef] [PubMed]
  42. Chen, J.; Morrow, D.J.; Fu, Y.; Zheng, W.; Zhao, Y.; Dang, L.; Stolt, M.J.; Kohler, D.D.; Wang, X.; Czech, K.J.; et al. Single-Crystal Thin Films of Cesium Lead Bromide Perovskite Epitaxially Grown on Metal Oxide Perovskite (SrTiO3). J. Am. Chem. Soc. 2017, 139, 13525–13532. [Google Scholar] [CrossRef] [Green Version]
  43. Zhumekenov, A.A.; Burlakov, V.M.; Saidaminov, M.I.; Alofi, A.; Haque, M.A.; Turedi, B.; Davaasuren, B.; Dursun, I.; Cho, N.; El-Zohry, A.M.; et al. The Role of Surface Tension in the Crystallization of Metal Halide Perovskites. ACS Energy Lett. 2017, 2, 1782–1788. [Google Scholar] [CrossRef] [Green Version]
  44. Poglitsch, A.; Weber, D. Dynamic Disorder in Methylammoniumtrihalogenoplumbates (II) Observed by Millimeter-wave Spectroscopy. J. Chem. Phys. 1987, 87, 6373–6378. [Google Scholar] [CrossRef]
  45. Liu, Y.; Yang, Z.; Cui, D.; Ren, X.; Sun, J.; Liu, X.; Zhang, J.; Wei, Q.; Fan, H.; Yu, F.; et al. Two-Inch-Sized Perovskite CH3NH3PbX3 (X = Cl, Br, I) Crystals: Growth and Characterization. Adv. Mater. 2015, 27, 5176–5183. [Google Scholar] [CrossRef] [PubMed]
  46. Saidaminov, M.I.; Abdelhady, A.L.; Murali, B.; Alarousu, E.; Burlakov, V.M.; Peng, W.; Dursun, I.; Wang, L.; He, Y.; Maculan, G.; et al. High-Quality Bulk Hybrid Perovskite Single Crystals within Minutes by Inverse Temperature Crystallization. Nat. Commun. 2015, 6, 7586. [Google Scholar] [CrossRef] [Green Version]
  47. Saidaminov, M.I.; Abdelhady, A.L.; Maculan, G.; Bakr, O.M. Retrograde Solubility of Formamidinium and Methylammonium Lead Halide Perovskites Enabling Rapid Single Crystal Growth. Chem. Commun. 2015, 51, 17658–17661. [Google Scholar] [CrossRef] [Green Version]
  48. Gupta, R.; Korukonda, T.B.; Gupta, S.K.; Dhamaniya, B.P.; Chhillar, P.; Datt, R.; Vashishtha, P.; Gupta, G.; Gupta, V.; Srivastava, R.; et al. Room Temperature Synthesis of Perovskite (MAPbI3) Single Crystal by Anti-Solvent Assisted Inverse Temperature Crystallization Method. J. Cryst. Growth 2020, 537, 125598. [Google Scholar] [CrossRef]
  49. He, M.; Li, B.; Cui, X.; Jiang, B.; He, Y.; Chen, Y.; O’Neil, D.; Szymanski, P.; EI-Sayed, M.A.; Huang, J.; et al. Meniscus-Assisted Solution Printing of Large-Grained Perovskite Films for High-Efficiency Solar Cells. Nat. Commun. 2017, 8, 16045. [Google Scholar] [CrossRef]
  50. Dang, Y.; Ju, D.; Wang, L.; Tao, X. Recent Progress in the Synthesis of Hybrid Halide Perovskite Single Crystals. CrystEngComm 2016, 18, 4476–4484. [Google Scholar] [CrossRef]
  51. Wei, H.; Fang, Y.; Mulligan, P.; Chuirazzi, W.; Fang, H.-H.; Wang, C.; Ecker, B.R.; Gao, Y.; Loi, M.A.; Cao, L.; et al. Sensitive X-Ray Detectors Made of Methylammonium Lead Tribromide Perovskite Single Crystals. Nat. Photonics 2016, 10, 333–339. [Google Scholar] [CrossRef]
  52. Sun, J.; Li, F.; Yuan, J.; Ma, W. Advances in Metal Halide Perovskite Film Preparation: The Role of Anti-Solvent Treatment. Small Methods 2021, 5, 2100046. [Google Scholar] [CrossRef] [PubMed]
  53. Ray, A.; Martín-García, B.; Moliterni, A.; Casati, N.; Boopathi, K.M.; Spirito, D.; Goldoni, L.; Prato, M.; Giacobbe, C.; Giannini, C.; et al. Mixed Dimethylammonium/Methylammonium Lead Halide Perovskite Crystals for Improved Structural Stability and Enhanced Photodetection. Adv. Mater. 2022, 34, 2106160. [Google Scholar] [CrossRef]
  54. Dang, Y.; Zhou, Y.; Liu, X.; Ju, D.; Xia, S.; Xia, H.; Tao, X. Formation of Hybrid Perovskite Tin Iodide Single Crystals by Top-Seeded Solution Growth. Angew. Chem. Int. Ed. 2016, 55, 3447–3450. [Google Scholar] [CrossRef] [PubMed]
  55. Ma, L.; Yan, Z.; Zhou, X.; Pi, Y.; Du, Y.; Huang, J.; Wang, K.; Wu, K.; Zhuang, C.; Han, X. A Polymer Controlled Nucleation Route towards the Generalized Growth of Organic-Inorganic Perovskite Single Crystals. Nat. Commun. 2021, 12, 2023. [Google Scholar] [CrossRef] [PubMed]
  56. Seo, S.; Jeon, I.; Xiang, R.; Lee, C.; Zhang, H.; Tanaka, T.; Lee, J.-W.; Suh, D.; Ogamoto, T.; Nishikubo, R.; et al. Semiconducting Carbon Nanotubes as Crystal Growth Templates and Grain Bridges in Perovskite Solar Cells. J. Mater. Chem. A 2019, 7, 12987–12992. [Google Scholar] [CrossRef]
  57. Jin, B.; Ming, Y.; Wu, Z.; Cao, J.; Liu, Y.; Zhu, Y.; Wang, S.; Liang, Z.; Wu, C. Silk Fibroin Induced Homeotropic Alignment of Perovskite Crystals toward High Efficiency and Stability. Nano Energy 2022, 94, 106936. [Google Scholar] [CrossRef]
  58. Wang, S.; Gu, Z.; Zhao, R.; Zhang, T.; Lou, Y.; Guo, L.; Su, M.; Li, L.; Zhang, Y.; Song, Y. A General Method for Growth of Perovskite Single-Crystal Arrays for High Performance Photodetectors. Nano Res. 2022, 15, 6568–6573. [Google Scholar] [CrossRef]
  59. Lee, M.M.; Teuscher, J.; Miyasaka, T.; Murakami, T.N.; Snaith, H.J. Efficient Hybrid Solar Cells Based on Meso-Superstructured Organometal Halide Perovskites. Science 2012, 338, 643–647. [Google Scholar] [CrossRef] [Green Version]
  60. Liang, P.-W.; Liao, C.-Y.; Chueh, C.-C.; Zuo, F.; Williams, S.T.; Xin, X.-K.; Lin, J.; Jen, A.K.-Y. Additive Enhanced Crystallization of Solution-Processed Perovskite for Highly Efficient Planar-Heterojunction Solar Cells. Adv. Mater. 2014, 26, 3748–3754. [Google Scholar] [CrossRef]
  61. Eperon, G.E.; Burlakov, V.M.; Docampo, P.; Goriely, A.; Snaith, H.J. Morphological Control for High Performance, Solution-Processed Planar Heterojunction Perovskite Solar Cells. Adv. Funct. Mater. 2014, 24, 151–157. [Google Scholar] [CrossRef]
  62. Tan, Z.-K.; Moghaddam, R.S.; Lai, M.L.; Docampo, P.; Higler, R.; Deschler, F.; Price, M.; Sadhanala, A.; Pazos, L.M.; Credgington, D.; et al. Bright Light-Emitting Diodes Based on Organometal Halide Perovskite. Nat. Nanotechnol. 2014, 9, 687–692. [Google Scholar] [CrossRef]
  63. Wang, J.; Wang, N.; Jin, Y.; Si, J.; Tan, Z.-K.; Du, H.; Cheng, L.; Dai, X.; Bai, S.; He, H.; et al. Interfacial Control Toward Efficient and Low-Voltage Perovskite Light-Emitting Diodes. Adv. Mater. 2015, 27, 2311–2316. [Google Scholar] [CrossRef] [PubMed]
  64. Deschler, F.; Price, M.; Pathak, S.; Klintberg, L.E.; Jarausch, D.-D.; Higler, R.; Hüttner, S.; Leijtens, T.; Stranks, S.D.; Snaith, H.J.; et al. High Photoluminescence Efficiency and Optically Pumped Lasing in Solution-Processed Mixed Halide Perovskite Semiconductors. J. Phys. Chem. Lett. 2014, 5, 1421–1426. [Google Scholar] [CrossRef]
  65. Li, J.; Han, H.; Li, B.; Zhao, C.; Xu, J.; Yao, J. Solvent Evaporation Induced Preferential Crystal Orientation BiI3 Films for the High Efficiency MA3Bi2I9 Perovskite Solar Cells. J. Alloys Compd. 2022, 909, 164725. [Google Scholar] [CrossRef]
  66. Bai, D.; Bian, H.; Jin, Z.; Wang, H.; Meng, L.; Wang, Q.; Liu, S. Temperature-Assisted Crystallization for Inorganic CsPbI2Br Perovskite Solar Cells to Attain High Stabilized Efficiency 14.81%. Nano Energy 2018, 52, 408–415. [Google Scholar] [CrossRef]
  67. Kang, R.; Kim, J.-E.; Yeo, J.-S.; Lee, S.; Jeon, Y.-J.; Kim, D.-Y. Optimized Organometal Halide Perovskite Planar Hybrid Solar Cells via Control of Solvent Evaporation Rate. J. Phys. Chem. C 2014, 118, 26513–26520. [Google Scholar] [CrossRef]
  68. Vaynzof, Y. The Future of Perovskite Photovoltaics—Thermal Evaporation or Solution Processing? Adv. Energy Mater. 2020, 10, 2003073. [Google Scholar] [CrossRef]
  69. Hudait, B.; Dutta, S.K.; Patra, A.; Nasipuri, D.; Pradhan, N. Facets Directed Connecting Perovskite Nanocrystals. J. Am. Chem. Soc. 2020, 142, 7207–7217. [Google Scholar] [CrossRef]
  70. Michell, R.M.; Müller, A.J. Confined Crystallization of Polymeric Materials. Prog. Polym. Sci. 2016, 54–55, 183–213. [Google Scholar] [CrossRef]
  71. Reiter, G.; Botiz, I.; Graveleau, L.; Grozev, N.; Albrecht, K.; Mourran, A.; Möller, M. Morphologies of Polymer Crystals in Thin Films. In Lecture Notes in Physics: Progress in Understanding of Polymer Crystallization; Reiter, G., Strobl, G.R., Eds.; Springer: Berlin/Heidelberg, Germany, 2007; Volume 714, pp. 179–200. [Google Scholar] [CrossRef]
  72. Nguyen-Tran, T.; Truong, T.T.; Nguyen, T.M.; Nguyen, D.T.; Luu, Q.M.; Nguyen, H.H.; Tran, C.T.K.; Bui, H.T.T. Growth and Morphology Control of CH3NH3PbBr3 Crystals. J. Mater. Sci. 2019, 54, 14797–14808. [Google Scholar] [CrossRef]
  73. Demange, G.; Zapolsky, H.; Patte, R.; Brunel, M. A Phase Field Model for Snow Crystal Growth in Three Dimensions. npj Comput. Mater. 2017, 3, 15. [Google Scholar] [CrossRef] [Green Version]
  74. Sekerka, R.F. Role of Instabilities in Determination of the Shapes of Growing Crystals. J. Cryst. Growth 1993, 128, 1–12. [Google Scholar] [CrossRef]
  75. Grozev, N.; Botiz, I.; Reiter, G. Morphological Instabilities of Polymer Crystals. Eur. Phys. J. E 2008, 27, 63–71. [Google Scholar] [CrossRef]
  76. Mehdi, H.; Mhamdi, A.; Bouazizi, A. Effect of Perovskite Precursor Ratios and Solvents Volume on the Efficiency of MAPbI3-XClx Mixed Halide Perovskite Solar Cells. Mater. Sci. Semicond. Process. 2020, 109, 104915. [Google Scholar] [CrossRef]
Figure 1. (Left) Schematic representation of the experimental setup used to monitor the growth of perovskite crystals in real time and direct space. (Right) An optical photograph depicting the equipment used to conduct the work presented herein. A hot stage connected to its corresponding temperature controller is positioned under the OM. The OM camera is linked to a PC. A sample (i.e., a Teflon ring glued on a glass substrate) is placed on the hot stage. The latter is more visible in the zoomed inset shown in the upper right corner.
Figure 1. (Left) Schematic representation of the experimental setup used to monitor the growth of perovskite crystals in real time and direct space. (Right) An optical photograph depicting the equipment used to conduct the work presented herein. A hot stage connected to its corresponding temperature controller is positioned under the OM. The OM camera is linked to a PC. A sample (i.e., a Teflon ring glued on a glass substrate) is placed on the hot stage. The latter is more visible in the zoomed inset shown in the upper right corner.
Materials 16 02625 g001
Figure 2. Optical micrographs depicting sixfold symmetrical perovskite crystals grown from a hybrid lead-mixed halide CH3NH3PbI3–xClx perovskite precursor ink at processing temperatures of 110 °C (ad), 80 °C (fi), 60 °C (kn) and 40 °C (ps) and at corresponding growth rates of 9.74 µm/s (e), 3.06 µm/s (j), 0.64 µm/s (o) and 0.19 µm/s (t), respectively. The white dotted arrows indicate the direction along which the growth rates were determined. The dashed lines in (e,j,o,t) are linear fits to the data points. Their corresponding slopes indicate the average crystal growth rates. Note that t0 is different for each temperature.
Figure 2. Optical micrographs depicting sixfold symmetrical perovskite crystals grown from a hybrid lead-mixed halide CH3NH3PbI3–xClx perovskite precursor ink at processing temperatures of 110 °C (ad), 80 °C (fi), 60 °C (kn) and 40 °C (ps) and at corresponding growth rates of 9.74 µm/s (e), 3.06 µm/s (j), 0.64 µm/s (o) and 0.19 µm/s (t), respectively. The white dotted arrows indicate the direction along which the growth rates were determined. The dashed lines in (e,j,o,t) are linear fits to the data points. Their corresponding slopes indicate the average crystal growth rates. Note that t0 is different for each temperature.
Materials 16 02625 g002
Figure 3. Optical micrographs depicting fourfold symmetrical crystals grown from hybrid lead-mixed halide CH3NH3PbI3–xClx perovskite at processing temperatures of 90 °C (ad), 80 °C (eh), 70 °C (il), 60 °C (mp) and 50 °C (qt). Note that t0 is different for each temperature.
Figure 3. Optical micrographs depicting fourfold symmetrical crystals grown from hybrid lead-mixed halide CH3NH3PbI3–xClx perovskite at processing temperatures of 90 °C (ad), 80 °C (eh), 70 °C (il), 60 °C (mp) and 50 °C (qt). Note that t0 is different for each temperature.
Materials 16 02625 g003
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Petrovai, I.; Todor-Boer, O.; David, L.; Botiz, I. Growth of Hybrid Perovskite Crystals from CH3NH3PbI3–xClx Solutions Subjected to Constant Solvent Evaporation Rates. Materials 2023, 16, 2625. https://doi.org/10.3390/ma16072625

AMA Style

Petrovai I, Todor-Boer O, David L, Botiz I. Growth of Hybrid Perovskite Crystals from CH3NH3PbI3–xClx Solutions Subjected to Constant Solvent Evaporation Rates. Materials. 2023; 16(7):2625. https://doi.org/10.3390/ma16072625

Chicago/Turabian Style

Petrovai, Ioan, Otto Todor-Boer, Leontin David, and Ioan Botiz. 2023. "Growth of Hybrid Perovskite Crystals from CH3NH3PbI3–xClx Solutions Subjected to Constant Solvent Evaporation Rates" Materials 16, no. 7: 2625. https://doi.org/10.3390/ma16072625

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