Next Article in Journal
The Nano4XX Nanotechnology Platform: The Triumph of Nanotechnology
Previous Article in Journal
Progress and Challenges of Chloride–Iodide Perovskite Solar Cells: A Critical Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Electronic, Structural, Optical, and Electrical Properties of CsPbX3 Powders (X = Cl, Br, and I) Prepared Using a Surfactant-Free Hydrothermal Approach

by
Carlos Echeverría-Arrondo
1,
Agustin O. Alvarez
1,
Sofia Masi
1,
Francisco Fabregat-Santiago
1 and
Felipe A. La Porta
1,2,3,*
1
Institute of Advanced Materials (INAM), Universitat Jaume I, Av. Sos Baynat, s/n, 12006 Castellón, Spain
2
Nanotechnology and Computational Chemistry Laboratory (NANOQC), Federal Technological University of Paraná, Avenida dos Pioneiros 3131, Londrina 86036-370, PR, Brazil
3
Post-Graduation Program in Chemistry, State University of Londrina, Londrina 86057-970, PR, Brazil
*
Author to whom correspondence should be addressed.
Nanomanufacturing 2023, 3(2), 217-227; https://doi.org/10.3390/nanomanufacturing3020013
Submission received: 19 January 2023 / Revised: 3 May 2023 / Accepted: 5 May 2023 / Published: 19 May 2023

Abstract

:
Recently, several strategies have been adopted for the cesium lead halide, CsPbX3 (X = Cl, Br, and/or I), crystal growth with a perovskite-type structure, paving the way for the further development of innovative optoelectronic and photovoltaic applications. The optoelectronic properties of advanced materials are controlled, in principle, by effects of morphology, particle size, structure, and composition, as well as imperfections in these parameters. Herein, we report a detailed investigation, using theoretical and experimental approaches to evaluate the structural, electronic, optical, and electrical properties of CsPbX3 microcrystals. The microcrystals are synthesized successfully using the hydrothermal method without surfactants. This synthetic approach also offers an easy upscaling for perovskite-related material synthesis from low-cost precursors. Lastly, in this direction, we believe that deeper mechanistic studies, based on the synergy between theory and practice, can guide the discovery and development of new advanced materials with highly tailored properties for applications in optoelectronic devices, as well as other emergent technologies.

1. Introduction

Due to their extraordinary physical properties, lead halide perovskites have attracted noteworthy interest in emerging technologies (photovoltaic, sensor, optoelectronic, etc.) [1,2,3,4]. Notably, lead halide perovskite structures are found in related compounds with the general formula APbX3, usually consisting of a ternary combination of corner-sharing [PbX6] polyhedral clusters in the lattice with an organic or inorganic monovalent cation A, such as methylammonium (MA), formamidinium (FA), or cesium (Cs), with the X site being primarily a monovalent halide (i.e., Cl, Br, I, or a combination thereof) anion [3]. It is well known that APbX3-related compounds can adopt four phases (a-cubic, b-tetragonal, and g- and d-orthorhombic) depending on the tilt and rotation of the [PbX6] polyhedral clusters in the lattice [4,5,6,7]. For this reason, the formation mechanisms of the most likely crystal structure of lead halide perovskites have been widely studied in the literature from the perspective of the Goldschmidt tolerance factor, t, defined as t = (rA + rX)/[sqrt(2)(rB + rX)], where rA, rB, and rX are the ionic radii of the individual A, B, and X species, respectively [7]. Hence, according to this empirical relation, when t ranges from 0.7 to 1, stable lead halide perovskites can generally be formed [3,8,9].
Such materials have a high power conversion efficiency (PCE) with a record of 25.8% (certified 25.5%), which has been achieved over the last decade [10]. Despite this significant PCE result, however, the stability of APbX3-based devices is still weak with regard to well-established photovoltaic technologies [9,10,11,12]. It has also been reported that the metallization method of perovskite solar cells significantly impacts device efficiency (i.e., mainly due to the well-known plasmonic effect) [13,14]. Overall, the addition of nanoscale additives in the form of metallic nanoparticles is an important strategy for enhancing the optoelectronic properties and performance/durability of perovskite-based solar cells, as they can act as nucleation sites for the growth of larger perovskite crystals [13,14,15,16]. On the other hand, recent studies have reported that, among these materials, cesium lead halide (e.g., CsPbCl3, CsPbBr3, and CsPbI3) crystals with a perovskite-type structure have higher chemical stability and are far less soluble (in the significant solvents) compared with organic/inorganic (hybrid) compounds [2,17,18,19]. Indeed, diverse synthetic strategies have widely been adopted to obtain ternary Cs–Pb–X compounds, including solid-state reaction [20], coprecipitation [21], hot injection [2], Bridgman growth [22], sol–gel [23], and hydrothermal [24,25,26,27,28,29]. Among them, hydrothermal methods (Scheme 1) are a promising approach to the large-scale production of these materials with a low-cost and low-temperature process [24,25,26,27,28]. Although this synthetic approach can be considered highly innovative, mainly thanks to the leading role in obtaining a considerable variety of metastable phases with attractive and entirely new properties, interesting in the technological field [24,25,26,27,28,29], it should be noted that the use of hydrothermal methods for the preparation of lead halide perovskites has been neglected in the literature, when compared to the other methods mentioned. However, because of the current need to obtain perovskite powders from low-cost precursors for large-scale device manufacturing, applied in numerous industrial and scientific research possibilities in the field of energy, we believe that using this synthetic approach can be quite enjoyable for this purpose in the future [30,31].
For instance, Kayalvizhi et al. [27] reported the stoichiometric control of ternary Cs–Pb–Br crystals based on the Cs/Pb ratios used in hydrothermal synthesis. It is important to emphasize that the solvent usually used in and after perovskite synthesis can change the stoichiometry, as previously discussed in [29]. For this reason, many recent studies have used mechanochemical methods for the production of large amounts of perovskite crystals, which is a well-known synthetic strategy for the preparation of solvent-free materials [32,33,34]. Despite this, it is important to highlight here that the hydrothermal method makes it possible to easy obtain of huge variety of materials with a high level of purity and unique properties, which most of the time cannot be achieved via other synthetic methods [35,36]. Hence, the hydrothermal method is an interesting choice to meet the development of a surfactant-free lead halide perovskites crystals synthesis with practical applications, with the advantage of easy tailoring desired parameters based on the tuning of the synthesis conditions and, in turn, the final material properties, thus enlarging the window of applications [24,25,26,27,28,29,31].
In this paper, we report a combined experimental/computational investigation on the electronic, structural, optical, and electrical features of CsPbX3 microcrystals (X = Cl, Br, and I) synthesized successfully using the surfactant-free hydrothermal method. To complement the experiments, we performed periodic density functional theory (DFT) calculations for a better understanding of the experimental data. Of course, computational design methods are at the forefront of knowledge and have a pivotal role in elucidating the electronic structure of several advanced materials, i.e., contributing to accelerating the development of better materials. Our work highlights the importance of controlling and understanding the physical properties, and our approach is relevant to address the issue of the structural and electronic disorder degree in crystalline materials [37].

2. Materials and Methods

2.1. Materials and Hydrothermal Synthesis of CsPbX3 Powders

All reagents were purchased from Sigma-Aldrich, including lead carbonate (PbCO3, 98%), cesium carbonate (Cs2CO3, 99%), hydrogen chloride (HCl, 36 wt.%), hydrogen bromide (HBr, 48 wt.%), and hydrogen iodide (HI, 47 wt.%), and they were used without further purification. CsPbX3 powders were synthesized according to the hydrothermal method using stoichiometric amounts of PbCO3 and Cs2CO3 in a 1:1 molar ratio that were dissolved in 15 mL of 6 M haloid acid solution (HCl, HBr, or HI, respectively) under constant stirring for approximately 10 min. In sequence, the mixture reaction was transferred into a Teflon autoclave that was properly sealed and placed inside a muffle furnace. For all cases, the hydrothermal process was performed free of any template and/or surfactants using a heating rate of 10 °C·min−1 at 140 °C for 2 h.

2.2. Characterization

The as-prepared powders were characterized using a Raman spectrometer model NRS-3100 (JASCO) with He–Ne and Argon ion lasers used for excitation at 633 nm. X-ray powder diffraction measurements in a 2θ range from 10° to 70° at 0.01°·min1 using an X-ray diffractometer (D8 Advance, Bruker-AXS) with Cu Kα radiation (λ = 1.5406 Å) are shown in the Figure S1 (in Supplementary Materials). The morphology of the prepared CsPbX3 powders was analyzed using field-emission scanning electron microscopy (SEM, JEOL 7001F). In this case, the optical properties were then analyzed using UV/Vis absorption spectroscopy with a Cary 300 Bio spectrophotometer operating in diffuse reflection mode. For measurement of the electrical response, polycrystalline pellets (with a diameter of 13.0 cm and a thickness of 4.0 mm) were prepared from the powders by cold-pressing at approximately 5 tons for 15 min, and then dried at 80 °C for 180 min. Subsequently, both sides of the pelletized samples were coated with a gold (Au) layer (approximately 100 nm thick), which was thermally evaporated at a rate of 0.8 A·s1 in an ultrahigh vacuum chamber. The capacitance and impedance spectroscopy measurements of the CsPbX3 pellet coated with Au were acquired using a Solartron 1260A Impedance/Gain-Phase Analyzer.

2.3. Computational Method and Models

The theoretical description of CsPbCl3, CsPbBr3, and CsPbI3 materials is based on computations performed with Quantum ESPRESSO, a density functional theory (DFT) code [38]. The ground-state electronic properties were obtained in two steps. First, we relaxed the atomic positions in the supercell until the forces on the individual nuclei became smaller than 0.001 Ry/Bohr (0.026 eV/A). We used spin polarization, the Optimized Norm-Conserving Vanderbilt pseudopotentials (SG15 ONCV database [39,40]), a plane-wave cutoff energy of 60 Ry, and a Monkhorst–Pack k-point grid over the first Brillouin zone (13 × 13 × 13 for CsPbCl3, 6 × 4 × 6 for CsPbBr3, and 4 × 8 × 3 for CsPbI3). Second, we computed the electronic parameters of the relaxed structures including spin–orbit interactions for a coarser k-point grid (9 × 9 × 9 for CsPbCl3, 4 × 3 × 4 for CsPbBr3, and 3 × 6 × 2 for CsPbI3).
We also investigated CsPbX3 crystals with halide vacancy defects. The computed defective supercells were composed of 134 atoms in the case of CsPbCl3 (3 × 3 × 3 unit cells), 159 atoms in the case of CsPbBr3 (2 × 2 × 2 unit cells), and 159 atoms in the case of CsPbI3 (2 × 4 × 1 unit cells). Since one anion was missing from the studied crystal (either Cl, Br, or I), the considered vacancy concentration was approximately 1%. This anion yielded a color center at the vacancy site and one Bohr magneton per unit cell. All these spin-polarized DFT calculations were run on the Γ point without including spin–orbit interactions.
We also calculated the activation energy for halide migration inside CsPbCl3, CsPbBr3, and CsPbI3 materials. To simulate the diffusion process, the referred ion was displaced in a series of small regular steps within the crystal. In the case of CsPbCl3, the supercell contained 39 atoms in eight unit cells, and its dimensions were 11.46/11.46/11.46 Å; in the case of CsPbBr3, 19 atoms were in a single unit cell with size 8.54/11.93/8.23 Å; in the case of CsPbI3, 39 atoms were in two unit cells with size 10.75/9.77/18.18 Å. The total energies were calculated for every position of the migrating ion along the diffusion path using the Quantum ESPRESSO code [38], Optimized Norm-Conserving Vanderbilt pseudopotentials, a 2 × 2 × 2 k-point grid, and a kinetic energy cutoff of 70 Ry. For initial and final steps, we relaxed the positions of all the atoms in the cells; however, for the intermediate steps, we relaxed all but the position of the migrating ion, which was kept fixed.

3. Results and Discussion

Figure 1A–C show the micro-Raman spectra obtained at room temperature for the CsPbX3 crystals synthesized using the hydrothermal method at 140 °C. We experimentally observed four Raman-active modes for both CsPbCl3 (at approximately 78, 160, 300, and 490 cm−1) and CsPbBr3 (at approximately 76, 128, 152, and 308 cm−1) crystals, as shown in Figure 1A,B, while the pure CsPbI3 crystals exhibited Raman-active modes at 85 and 375 cm−1. These Raman data are in good agreement with those reported in the literature [41,42]. The insets of Figure 1A–C show a schematic representation of the unit cells, cubic phase for CsPbCl3, and orthorhombic phase in both CsPbBr3 and δ-CsPbI3. The structural parameters of CsPbX3 crystals were obtained from DFT calculations (for CsPbCl3, a = 5.73 Å; for CsPbBr3 a = 8.54 Å, b = 11.93 Å, and c = 8.23 Å; for δ-CsPbI3, a = 10.75 Å, b = 4.89 Å, and c = 18.18 Å), which are in accordance with the literature values related to the same compounds but synthesized using different methods [22,43,44,45]. The calculated Pb–anion bond lengths and bond angles (Pb–X–Pb) in CsPbX3 are also in accordance with the literature [22,43,44,45].
The SEM images in Figure 1D–F show the synthesis route produces for both CsPbCl3 and CsPbBr3, with two different morphologies (ranging from cube to octahedral-like), while CsPbI3 exhibited a uniform morphology, only microrod-like. Energy-dispersive X-ray spectroscopy (EDXS) mapping revealed a similar profile for all samples, indicating that the products contained only Cs, Pb, and X (X = Cl, Br, and I) and were uniformly distributed in the desired compositional atomic percentage as synthesized (Figure 1D–F). However, a comparison between the micro-Raman spectra obtained for these as-prepared samples yielded a distinct characteristic for the CsPbI3 crystals. The peculiar CsPbI3 morphology and the different Raman spectrum compared to CsPbCl3 and CsPbBr3 highlights a higher density of structural defects.
To investigate the optical properties of the crystalline perovskite, a comparison between the projected density of states (DOS) and Tauc plot of the CsPbX3 powders are presented (Figure 2A–C). From the combination of experimental and theoretical approaches, we elucidated the bandgap nature of these materials. Our electronic structure calculations showed that the cubic CsPbCl3 and orthorhombic CsPbBr3 phases have a direct bandgap, while the orthorhombic CsPbI3 structure has an indirect transition. Thus, the experimental bandgap values for the as-synthesized CsPbX3 crystals were calculated from the Tauc plot as 2.60 eV, 2.14 eV, and 2.56 eV for CsPbCl3, CsPbBr3 and CsPbI3, respectively. Hence, this trend is in good agreement with the literature and confirms the formation of δ-CsPbI3 [18,45,46,47,48,49]. These bandgap energy values are strongly dependent on the crystalline perovskite structure. It is well known that, among all-inorganic perovskite materials, the cubic CsPbI3 structure (black phase) as an n-type semiconductor has a suitable bandgap of about 1.70 eV for photovoltaic applications [12,50,51,52,53]. However, due to small Cs+ cation, the cubic CsPbI3 structure is metastable at temperatures below 320 °C [12,51,52,53]. As a general strategy, researchers have partially replaced Cs+ with some larger cations, such as methylammonium (abbreviated as MA+) [54,55] and formamidinium (abbreviated as FA+) [53,54,56,57]. Hence, stabilizing the cubic CsPbI3 structure in hydrothermal treatment conditions is still challenging and represents an open field of investigation.
In addition, in order to investigate the influence of defects on the fundamental properties of these crystals, DFT calculations were performed. For this purpose, we removed one anion from the DFT supercell, either Cl, Br, or I, leading to a vacancy concentration of approximately 1%. As expected, the projected DOS (Figure 2D–I) exhibited a vacancy state within the bandgap. This localized state has a p-like character provided that the released electron is occupying one of the partially filled Pb 6p orbitals around the vacancy.
The effective masses of the bulk compounds were obtained as the following mean values: me = Σime,i/3 and mh = Σimh,i/3, where I = x, y, and z are correspond to the cartesian directions in the reciprocal space [58]. To estimate me,i and mh,i, we calculated the electronic energies at 13 k points in the vicinity of the bandgap, and fit them to the parabola E(k) = Eg + ħ2k2(m0/m)/(2m0), where Eg is the bandgap, m0 is the electron mass, and m is the effective mass. From the dispersion band structure, we also estimated the effective masses of carriers and the resulting reduced masses μ = me × mh/(me + mh) (see Table 1), which can be accessed experimentally [1], thereby serving as a reference to positively evaluate the theoretical level employed here [59,60]. In addition, me and mh for mixed-halide CsPb(I1−xBrx)3 and CsPb(Br1−xClx)3 perovskites in the orthorhombic phase are also reported in the literature [61].
Lastly, the electrical response of the prepared CsPbX3 pellets as n-type semiconductors was investigated in different humidity conditions at room temperature, as summarized in Figure 3. For this analysis, an impedance spectroscopy (IS) analysis was carried out directly on the CsPbX3 pellets, fitted with an equivalent circuit (Figure 3B,C). The proposed equivalent circuit consists of a contact resistance (denoted as RS), connected in series with the elements representing the processes in the polycrystalline perovskite: the parallel combination of the charge transport resistance (denoted as Rtr) and dielectric capacitance (denoted as Cd). An interesting feature is that, at room temperature, a decrease in the transport resistance could be observed when the samples were exposed to humidity (see Figure 3B). These results are consistent with the literature [62]. However, a more significant variation could be observed for the CsPbI3 pellet. We believe that this is associated with its different morphology observed in SEM analysis. In other words, it suggests that the prepared CsPbI3 pellets had a higher porosity, when compared to CsPbCl3 and CsPbBr3 samples, which could in principle allow a better mobility for water ions or ease the formation of hydrated phases. Therefore, this contributed to higher resistance in the CsPbI3 pellet. Our analysis also demonstrated that the dielectric capacities of these pellets remained practically constant between the different humidity conditions (dry and wet, see Figure 3C), which may support intrinsic electronic polarization effects for these samples [62,63].
From the cyclic voltammetry (see Figure 3D–F), we found an increase in hysteresis with the decrease in monovalent halide radius (i.e., X = Cl, Br and I). For a full description of this behavior, we performed DFT calculations upon diffusion barriers as present in CsPbX3 bulk structures. We considered the vacancy-mediated migration of halide ions along a direction relevant to the overall diffusion flow, and generated the potential barriers as shown in Figure 3G,H.
On the basis of our DFT calculations, the diffusion barriers for CsPbCl3, CsPbBr3, and CsPbI3 were 0.45, 0.44, and 0.39 eV, respectively [64,65,66]. The decrease in barrier height for ion migration was likely due to a progressive reduction in the orbital overlap, strongly dependent on the crystal structure. Due to the complex nature of inhomogeneity in CsPbX3 structures, the observed experimental trend can be explained, in principle, by the mechanism of vacancy-mediated anion diffusion, which has been studied for lead halide perovskite materials [64].

4. Conclusions

In summary, the hydrothermal method was developed to prepare CsPbX3 powders at 140 °C for 120 min, highlighting its potential to obtain single-phase perovskite materials at low temperatures with a high reaction yield; this can be considered a very attractive strategy for their obtention on a large scale. Thus, given the wide versatility of these halide perovskite-based materials, the understanding of the structural/electronic differences observed for these materials was analyzed from periodic models based on DFT calculations. The analysis of theoretical results indicated that a breaking symmetry process in these structures is, in principle, strongly associated by order-disorder effects; furthermore, this disordered structure leads to the formation of new electronic states in the bandgap. Our findings also suggest that defect states, in turn, play a pivotal role in the optical and electronic responses of CsPbX3 perovskite structures. Therefore, the combined use of such strategies based on the synergy between theory and practice might provide clues to obtain novel advanced materials that will give rise to new technological applications in the future.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nanomanufacturing3020013/s1, Figure S1: XRD patterns of CsPbX3 samples prepared by hydrothermal method: (A) CsPbCl3, (B) CsPbBr3 and (C) CsPbI3.

Author Contributions

Conceptualization, F.A.L.P.; methodology, F.A.L.P., A.O.A., S.M. and C.E.-A.; software, C.E.-A.; validation, F.A.L.P., A.O.A., S.M. and C.E.-A.; formal analysis, F.A.L.P. and S.M.; investigation, F.A.L.P., A.O.A., S.M. and C.E.-A.; resources, F.A.L.P. and F.F.-S.; data curation, F.A.L.P., A.O.A., S.M. and C.E.-A.; writing—original draft preparation, F.A.L.P., C.E.-A. and S.M.; writing—review and editing, F.A.L.P., S.M. and F.F.-S.; visualization, F.A.L.P., A.O.A., S.M. and C.E.-A.; funding acquisition; F.A.L.P. and F.F.-S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Brazilian agencies Fundação Araucária, CAPES, and CNPq (203012/2017-8). The MAESTRO project that resulted in this publication received funding from the European Union’s Horizon 2020 research and innovation program under grant agreement No. 764787.

Data Availability Statement

Not applicable.

Acknowledgments

F.A.L.P. is thankful to CNPq for the fellowships. The authors are especially grateful to Iván Mora Seró (INAM at University Jaume I) for the support and discussion of these results.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Brenner, T.M.; Egger, D.A.; Kronik, L.; Hodes, G.; Cahen, D. Hybrid organic—Inorganic perovskites: Low-cost semiconductors with intriguing charge-transport properties. Nat. Rev. Mater. 2016, 1, 15007. [Google Scholar] [CrossRef]
  2. Protesescu, L.; Yakunin, S.; Bodnarchuk, M.I.; Krieg, F.; Caputo, R.; Hendon, C.H.; Yang, R.X.; Walsh, A.; Kovalenko, M.V. Nanocrystals of cesium lead halide perovskites (CsPbX3, X = Cl, Br, and I): Novel optoelectronic materials showing bright emission with wide color gamut. Nano Lett. 2015, 15, 3692–3696. [Google Scholar] [CrossRef] [PubMed]
  3. Manser, J.S.; Christians, J.A.; Kamat, P.V. Intriguing Optoelectronic Properties of Metal Halide Perovskites. Chem. Rev. 2016, 116, 12956–13008. [Google Scholar] [CrossRef]
  4. Pinto, F.M.; Dey, S.; Duarte, T.M.; Taft, C.A.; La Porta, F.A. Perovskite-like quantum dots designed for advanced optoelectronic applications. In Functional Properties of Advanced Engineering Materials and Biomolecules; La Porta, F., Taft, C., Eds.; Springer: Cham, Switzerland, 2021. [Google Scholar] [CrossRef]
  5. Glazer, A.M. The classification of tilted octahedra in perovskites. Acta Crystallogr. Sect. B 1972, 28, 3384–3392. [Google Scholar] [CrossRef]
  6. Nakata, M.M.; Mazzo, T.M.; Casali, G.P.; La Porta, F.A.; Longo, E. A large red-shift in the photoluminescence emission of Mg1−xSrxTiO3. Chem. Phys. Lett. 2015, 622, 9–14. [Google Scholar] [CrossRef]
  7. Sutton, R.J.; Filip, M.R.; Haghighirad, A.A.; Sakai, N.; Wenger, B.; Giustino, F.; Snaith, H.J. Cubic or Orthorhombic? Revealing the Crystal Structure of Metastable Black-Phase CsPbI3 by Theory and Experiment. ACS Energy Lett. 2018, 3, 1787–1794. [Google Scholar] [CrossRef]
  8. Goldschmidt, V.M. Die Gesetze der Krystallochemie. Naturwissenschaften 1926, 14, 477–485. [Google Scholar] [CrossRef]
  9. Akkerman, Q.A.; Abdelhady, A.L.; Manna, L. Zero-Dimensional Cesium Lead Halides: History, Properties, and Challenges. J. Phys. Chem. Lett. 2018, 9, 2326–2337. [Google Scholar] [CrossRef]
  10. Best Research-Cell Efficiency Chart. Available online: https://www.nrel.gov/pv/cell-efficiency.html (accessed on 15 January 2022).
  11. Saliba, M.; Matsui, T.; Domanski, K.; Correa-Baena, J.-P.; Nazeeruddin, M.K.; Zakeeruddin, S.M.; Tress, W.; Abate, A.; Hagfeldt, A.; Gratzel, M. Cesium-containing triple cation perovskite solar cells: Improved stability, reproducibility and high efficiency. Energy Environ. Sci. 2016, 9, 1989–1997. [Google Scholar] [CrossRef]
  12. Pinto, F.M.; de Conti, M.C.M.D.; Dey, S.; Velilla, E.; Taft, C.A.; La Porta, F.A. Emerging Metal-Halide Perovskite Materials for Enhanced Solar Cells and Light-Emitting Applications. In Research Topics in Bioactivity, Environment and Energy; Engineering Materials; Taft, C.A., de Lazaro, S.R., Eds.; Springer: Cham, Switzerland, 2022; pp. 45–85. [Google Scholar]
  13. Jacak, J.E.; Jacak, W.A. Routes for Metallization of Perovskite Solar Cells. Materials 2022, 15, 2254–2279. [Google Scholar] [CrossRef]
  14. Laska, M.; Krzemińska, Z.; Kluczk-Korch, K.; Schaadt, D.; Popko, E.; Jacak, W.A.; Jacak, J.E. Metallization of solar cells, exciton channel of plasmon photovoltaic effect in perovskite cells. Nano Energy 2020, 75, 104751. [Google Scholar] [CrossRef]
  15. Gao, Y.; Wu, Y.; Lu, H.; Chen, C.; Liu, Y.; Bai, X.; Yang, L.; Yu, W.W.; Dai, Q.; Zhang, Y. CsPbBr3 perovskite nanoparticles as additive for environmentally stable perovskite solar cells with 20.46% efficiency. Nano Energy 2019, 59, 517–526. [Google Scholar] [CrossRef]
  16. Thanh, N.T.K.; Maclean, N.; Mahiddine, S. Mechanisms of Nucleation and Growth of Nanoparticles in Solution. Chem. Rev. 2014, 114, 7610–7630. [Google Scholar] [CrossRef] [PubMed]
  17. Akkerman, Q.A.; Rainò, G.; Kovalenko, M.V.; Manna, L. Genesis, challenges and opportunities for colloidal lead halide perovskite nanocrystals. Nat. Mater. 2018, 17, 394–405. [Google Scholar] [CrossRef]
  18. Krieg, F.; Ochsenbein, S.; Yakunin, S.; Brinck, S.; Aellen, P.; Süess, A.; Clerc, B.; Guggisberg, D.; Nazarenko, O.; Shynkarenko, Y.; et al. Colloidal CsPbX3 (X = Cl, Br, I) Nanocrystals 2.0: Zwitterionic Capping Ligands for Improved Durability and Stability. ACS Energy Lett. 2018, 3, 641–646. [Google Scholar] [CrossRef]
  19. Wang, P.; Zhang, X.; Zhou, Y.; Jiang, Q.; Ye, Q.; Chu, Z.; Li, X.; Yang, X.; Yin, Z.; You, J. Solvent-controlled growth of inorganic perovskite films in dry environment for efficient and stable solar cells. Nat. Comm. 2018, 9, 2225. [Google Scholar] [CrossRef]
  20. Protesescu, L.; Yakunin, S.; Nazarenko, O.; Dirin, D.N.; Kovalenko, M.V. Low-Cost Synthesis of Highly Luminescent Colloidal Lead Halide Perovskite Nanocrystals by Wet Ball Milling. ACS Appl. Nano Mater. 2018, 1, 1300–1308. [Google Scholar] [CrossRef] [PubMed]
  21. Zhang, J.; Fan, L.; Li, J.; Liu, X.; Wang, R.; Wang, L.; Tu, G. Growth mechanism of CsPbBr3 perovskite nanocrystals by a co-precipitation method in a CSTR system. Nano Res. 2019, 12, 121–127. [Google Scholar] [CrossRef]
  22. Stoumpos, C.C.; Malliakas, C.D.; Peters, J.A.; Liu, Z.F.; Sebastian, M.; Im, J.; Chasapis, T.C.; Wibowo, A.C.; Chung, D.Y.; Freeman, A.J.; et al. Crystal Growth of the Perovskite Semiconductor CsPbBr3: A New Material for High-Energy Radiation Detection. Cryst. Growth Des. 2013, 13, 2722–2727. [Google Scholar] [CrossRef]
  23. Kerner, R.A.; Zhao, L.; Xiao, Z.; Rand, B.P. Ultrasmooth metal halide perovskite thin films via sol–gel processing. J. Mater. Chem. A 2016, 4, 8308–8315. [Google Scholar] [CrossRef]
  24. Liang, J.; Wang, C.; Zhao, P.; Lu, Z.; Xu, Z.; Wang, Y.; Zhu, H.; Hu, Y.; Zhu, G.; Ma, L.; et al. Solution synthesis and phase control of inorganic perovskites for high-performance optoelectronic devices. Nanoscale 2017, 9, 11841. [Google Scholar] [CrossRef] [PubMed]
  25. Vidyasagar, C.C.; Muñoz Flores, B.M.; Jiménez Pérez, V.M. Recent Advances in Synthesis and Properties of Hybrid Halide Perovskites for Photovoltaics. Nano-Micro Lett. 2018, 10, 68. [Google Scholar] [CrossRef] [PubMed]
  26. Darr, J.A.; Zhang, J.; Makwana, N.M.; Weng, X. Continuous Hydrothermal Synthesis of Inorganic Nanoparticles: Applications and Future Directions. Chem. Rev. 2017, 117, 11125–11238. [Google Scholar] [CrossRef]
  27. Kayalvizhi, T.; Sathaya, A.; Preethi Meher, K.R.S. Hydrothermal Synthesis of Perovskite CsPbBr3: Spotlight on the Role of Stoichiometry in Phase Formation Mechanism. J. Electron. Mater. 2022, 51, 3466–3475. [Google Scholar] [CrossRef]
  28. Kayalvizhi, T.; Sathaya, A.; Preethi Meher, K.R.S. Facile synthesis of monoclinic CsPbBr3—For promising photovoltaic characteristics. AIP Conf. Proc. 2020, 2265, 030654. [Google Scholar]
  29. La Porta, F.A.; Masi, S. Solvent-Mediated Structural Evolution Mechanism from Cs4PbBr6 to CsPbBr3 Crystals. Nanomanufacturing 2021, 1, 67–74. [Google Scholar] [CrossRef]
  30. Qiu, L.; Ono, L.K.; Qi, Y. Advances and challenges to the commercialization of organic-inorganic halide perovskite solar cell technology. Mater. Today Energy 2018, 7, 169–189. [Google Scholar] [CrossRef]
  31. Bibi, A.; Lee, I.; Nah, Y.; Allam, O.; Kim, H.; Quan, L.N.; Tang, J.; Walsh, A.; Jang, S.S.; Sargent, E.H.; et al. Lead-free halide double perovskites: Toward stable and sustainable optoelectronic devices. Mater. Today 2021, 49, 123–144. [Google Scholar] [CrossRef]
  32. Palazon, F.; El Ajjouri, Y.; Sebastia-Luna, P.; Lauciello, S.; Manna, L.; Bolink, H.J. Mechanochemical synthesis of inorganic halide perovskites: Evolution of phase-purity, morphology, and photoluminescence. J. Mater. Chem. C 2019, 7, 11406. [Google Scholar] [CrossRef]
  33. Rosales, B.A.; Wei, L.; Vela, J. Synthesis and mixing of complex halide perovskites by solvent-free solid-state methods. J. Solid State Chem. 2019, 271, 206–215. [Google Scholar] [CrossRef]
  34. Pal, P.; Saha, S.; Banik, A.; Sarkar, A.; Biswas, K. All-Solid-State Mechanochemical Synthesis and Post-Synthetic Transformation of Inorganic Perovskite-type Halides. Chem. Eur. J. 2018, 24, 1811–1815. [Google Scholar] [CrossRef] [PubMed]
  35. Mertzsch, N. Hydrothermal processes in industry. ChemTexts 2020, 6, 21. [Google Scholar] [CrossRef]
  36. De Conti, M.C.M.D.; Dey, S.; Pottker, W.E.; La Porta, F.A. An overview into advantages and applications of conventional and unconventional hydro(solvo)thermal approaches for novel advanced materials design. In ChemRxiv; Cambridge Open Engage: Cambridge, UK, 2022. [Google Scholar]
  37. Pinto, F.M.; Suzuki, V.Y.; Silva, R.C.; La Porta, F.A. Oxygen defects and surface chemistry of reducible oxides. Front. Mater. 2019, 6, 260. [Google Scholar] [CrossRef]
  38. Giannozzi, P.; Baroni, S.; Bonini, N.; Calandra, M.; Car, R.; Cavazzoni, C.; Ceresoli, D.; Chiarotti, G.L.; Cococcioni, M.; Dabo, I.; et al. QUANTUM ESPRESSO: A modular and open-source software project for quantum simulations of materials. J. Phys. Condens. Matter 2009, 21, 395502. [Google Scholar] [CrossRef] [PubMed]
  39. Web Standards and Reference Data for First-Principles Simulations. Available online: http://www.quantum-simulation.org (accessed on 5 May 2019).
  40. Hamann, D.R. Optimized norm-conserving Vanderbilt pseudopotentials. Phys. Rev. B 2017, 88, 085117. [Google Scholar] [CrossRef]
  41. Thesika, K.; Murugan, A.V. Microwave-Enhanced Chemistry at Solid–Liquid Interfaces: Synthesis of All-Inorganic CsPbX3 Nanocrystals and Unveiling the Anion-Induced Evolution of Structural and Optical Properties. Inorg. Chem. 2020, 59, 6161–6175. [Google Scholar] [CrossRef]
  42. Cha, J.-H.; Han, J.H.; Yin, W.; Park, C.; Park, Y.; Ahn, T.K.; Cho, J.H.; Jung, D.-Y. Photoresponse of CsPbBr3 and Cs4PbBr6 Perovskite Single Crystals. J. Phys. Chem. Lett. 2017, 8, 565–570. [Google Scholar] [CrossRef]
  43. Moller, C.K. Crystal structure and photoconductivity of cesium plumbohalides. Nature 1958, 182, 143. [Google Scholar] [CrossRef]
  44. Trots, D.M.; Myagkota, S.V. High-Temperature Structural Evolution of Caesium and Rubidium Triiodoplumbates. J. Phys. Chem. Solids 2008, 69, 2520–2526. [Google Scholar] [CrossRef]
  45. Zhao, X.; Tang, T.; Xie, Q.; Gao, L.; Lu, L.; Tang, Y. First-principles study on the electronic and optical properties of the orthorhombic CsPbBr3 and CsPbI3 with Cmcm space group. New J. Chem. 2021, 45, 15857–15862. [Google Scholar] [CrossRef]
  46. Liao, M.; Shan, B.; Li, M. In Situ Raman Spectroscopic Studies of Thermal Stability of All-Inorganic Cesium Lead Halide (CsPbX3, X = Cl, Br, I) Perovskite Nanocrystals. J. Phys. Chem. Lett. 2019, 10, 1217–1225. [Google Scholar] [CrossRef] [PubMed]
  47. Lai, M.; Kong, Q.; Bischak, C.G.; Yu, Y.; Dou, L.; Eaton, S.W.; Ginsberg, N.S.; Yang, P. Structural, optical, and electrical properties of phase-controlled cesium lead iodide nanowires. Nano Res. 2017, 10, 1107–1114. [Google Scholar] [CrossRef]
  48. Zhang, H.; Liu, X.; Dong, J.; Yu, H.; Zhou, C.; Zhang, B.; Xu, Y.; Jie, W. Centimeter-sized inorganic lead halide perovskite CsPbBr3 crystals grown by an improved solution method. Cryst. Growth Des. 2017, 17, 6426–6431. [Google Scholar] [CrossRef]
  49. Linaburg, M.R.; McClure, E.T.; Majher, J.D.; Woodward, P.M. Cs1–xRbxPbCl3 and Cs1–xRbxPbBr3 Solid Solutions: Understanding Octahedral Tilting in Lead Halide Perovskites. Chem. Mater. 2017, 29, 3507–3514. [Google Scholar] [CrossRef]
  50. Xu, S.; Libanori, A.; Luo, G.; Chen, J. Engineering bandgap of CsPbI3 over 1.7 eV with enhanced stability and transport properties. IScience 2021, 24, 102235. [Google Scholar] [CrossRef]
  51. Chhillar, P.; Dhamaniya, B.P.; Pathak, S.K.; Karak, S. Stabilization of Photoactive γ-CsPbI3 Perovskite Phase by Incorporation of Mg. ACS Appl. Electron. Mater. 2022, 4, 5368–5378. [Google Scholar] [CrossRef]
  52. Filip, M.R.; Eperon, G.E.; Snaith, H.J.; Giustino, F. Steric Engineering of Metal-Halide Perovskites with Tunable Optical Band Gaps. Nat. Commun. 2014, 5, 5757. [Google Scholar] [CrossRef]
  53. Masi, S.; Gualdrón-Reyes, A.F.; Mora-Seró, I. Stabilization of Black Perovskite Phase in FAPbI3 and CsPbI3. ACS Energy Lett. 2020, 5, 1974–1985. [Google Scholar] [CrossRef]
  54. Jeon, N.J.; Noh, J.H.; Yang, W.S.; Kim, Y.C.; Ryu, S.; Seo, J.; Seok, S.I. Compositional Engineering of Perovskite Materials for High-Performance Solar Cells. Nature 2015, 517, 476–480. [Google Scholar] [CrossRef]
  55. Gong, J.; Guo, P.; Benjamin, S.E.; Gregory Van Patten, P.; Schaller, R.D.; Xu, T. Cation engineering on lead iodide perovskites for stable and high-performance photovoltaic applications. J. Energy Chem. 2018, 27, 1017–1039. [Google Scholar] [CrossRef]
  56. Li, Z.; Yang, M.; Park, J.-S.; Wei, S.-H.; Berry, J.J.; Zhu, K. Stabilizing Perovskite Structures by Tuning Tolerance Factor: Formation of Formamidinium and Cesium Lead Iodide Solid-State Alloys. Chem. Mater. 2016, 28, 284–292. [Google Scholar] [CrossRef]
  57. Hao, M.; Bai, Y.; Zeiske, S.; Ren, L.; Liu, J.; Yuan, Y.; Zarrabi, N.; Cheng, N.; Ghasemi, M.; Chen, P.; et al. Ligand-assisted cation-exchange engineering for high-efficiency colloidal Cs1−xFAxPbI3 quantum dot solar cells with reduced phase segregation. Nat. Energy 2020, 5, 79–88. [Google Scholar] [CrossRef]
  58. Luttinger, J.M.; Kohn, W. Motion of electrons and holes in perturbed periodic fields. Phys. Rev. 1955, 97, 869. [Google Scholar] [CrossRef]
  59. Kohanoff, J. Electronic Structure Calculations for Solids and Molecules: Theory and Computational Methods; Cambridge University Press: Cambridge, UK, 2006. [Google Scholar]
  60. Tanaka, K.; Takahashi, T.; Ban, T.; Kondo, T.; Uchida, K.; Miura, N. Comparative study on the excitons in lead-halide-based perovskite-type crystals CH 3NH 3PbBr 3 CH3NH3PbI3. Solid State Commun. 2003, 127, 619–623. [Google Scholar] [CrossRef]
  61. Ghaithan, H.M.; Alahmed, Z.A.; Qaid, S.M.H.; Aldwayyan, A.S. Density Functional Theory Analysis of Structural, Electronic, and Optical Properties of Mixed-Halide Orthorhombic Inorganic Perovskites. ACS Omega 2021, 6, 30752–30761. [Google Scholar] [CrossRef]
  62. García-Fernández, A.; Moradi, Z.; Bermúdez-García, J.M.; Sánchez-Andújar, M.; Gimeno, V.A.; Castro-García, S.; Señarís-Rodríguez, M.A.; Mas-Marzá, E.; Garcia-Belmonte, G.; Fabregat-Santiago, F. Effect of Environmental Humidity on the Electrical Properties of Lead Halide Perovskites. J. Phys. Chem. C 2019, 123, 2011–2018. [Google Scholar] [CrossRef]
  63. Bisquert, J.; Bertoluzzi, L.; Mora-Sero, I.; Garcia-Belmonte, G. Theory of Impedance and Capacitance Spectroscopy of Solar Cells with Dielectric Relaxation, Drift-Diffusion Transport, and Recombination. J. Phys. Chem. C 2014, 118, 18983–18991. [Google Scholar] [CrossRef]
  64. Walsh, A.; Stranks, S.D. Taking Control of Ion Transport in Halide Perovskite Solar Cells. ACS Energy Lett. 2018, 3, 1983–1990. [Google Scholar] [CrossRef]
  65. Eames, C.; Frost, J.M.; Barnes, P.R.F.; O’Regan, B.C.; Walsh, A.; Islam, M.S. Ionic transport in hybrid lead iodide perovskite solar cells. Nat. Comm. 2015, 6, 7497. [Google Scholar] [CrossRef]
  66. Mizusaki, J.; Arai, K.; Fueki, K. Ionic conduction of the perovskite-type halides. Solid State Ion. 1983, 11, 203–211. [Google Scholar] [CrossRef]
Scheme 1. Hydrothermal synthesis of CsPbX3 crystals from precursor dispersion.
Scheme 1. Hydrothermal synthesis of CsPbX3 crystals from precursor dispersion.
Nanomanufacturing 03 00013 sch001
Figure 1. (AC) Micro-Raman spectra, (DF) SEM images, and EDXS mapping of the CsPbX3 powders synthesized using hydrothermal method at 140 °C: (left) CsPbCl3, (middle) CsPbBr3, and (right) d-CsPbI3. Inset schematic representation of the cubic and orthorhombic unit cells corresponding to the CsPbX3 crystals.
Figure 1. (AC) Micro-Raman spectra, (DF) SEM images, and EDXS mapping of the CsPbX3 powders synthesized using hydrothermal method at 140 °C: (left) CsPbCl3, (middle) CsPbBr3, and (right) d-CsPbI3. Inset schematic representation of the cubic and orthorhombic unit cells corresponding to the CsPbX3 crystals.
Nanomanufacturing 03 00013 g001
Figure 2. Tauc plot from UV/Vis spectra of the prepared CsPbX3 powders: (A) CsPbCl3, (B) CsPbBr3, and (C) δ-CsPbI3. (DI) Projected DOS for the ordered (left) and disordered (right) CsPbX3 models. The position of the defect states is indicated by arrows.
Figure 2. Tauc plot from UV/Vis spectra of the prepared CsPbX3 powders: (A) CsPbCl3, (B) CsPbBr3, and (C) δ-CsPbI3. (DI) Projected DOS for the ordered (left) and disordered (right) CsPbX3 models. The position of the defect states is indicated by arrows.
Nanomanufacturing 03 00013 g002
Figure 3. (B) Resistance and (C) capacitance corresponding to the fitting of the IS measurements with the (A) equivalent circuit, and (DF) cyclic voltammetry at a rate of 20 mV/s, obtained for dry and wet prepared CsPbX3 pellets at room temperature. (G) Migration paths in CsPbCl3, CsPbBr3, and CsPbI3 perovskites and (H) corresponding energy profiles along these paths. Vacancies are indicated as blue dotted circles.
Figure 3. (B) Resistance and (C) capacitance corresponding to the fitting of the IS measurements with the (A) equivalent circuit, and (DF) cyclic voltammetry at a rate of 20 mV/s, obtained for dry and wet prepared CsPbX3 pellets at room temperature. (G) Migration paths in CsPbCl3, CsPbBr3, and CsPbI3 perovskites and (H) corresponding energy profiles along these paths. Vacancies are indicated as blue dotted circles.
Nanomanufacturing 03 00013 g003
Table 1. Effective and reduced masses calculated for CsPbX3 structures.
Table 1. Effective and reduced masses calculated for CsPbX3 structures.
Modelmemhµ
CsPbCl30.280.270.14
CsPbBr30.240.250.12
CsPbI31.321.480.70
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

Echeverría-Arrondo, C.; Alvarez, A.O.; Masi, S.; Fabregat-Santiago, F.; Porta, F.A.L. Electronic, Structural, Optical, and Electrical Properties of CsPbX3 Powders (X = Cl, Br, and I) Prepared Using a Surfactant-Free Hydrothermal Approach. Nanomanufacturing 2023, 3, 217-227. https://doi.org/10.3390/nanomanufacturing3020013

AMA Style

Echeverría-Arrondo C, Alvarez AO, Masi S, Fabregat-Santiago F, Porta FAL. Electronic, Structural, Optical, and Electrical Properties of CsPbX3 Powders (X = Cl, Br, and I) Prepared Using a Surfactant-Free Hydrothermal Approach. Nanomanufacturing. 2023; 3(2):217-227. https://doi.org/10.3390/nanomanufacturing3020013

Chicago/Turabian Style

Echeverría-Arrondo, Carlos, Agustin O. Alvarez, Sofia Masi, Francisco Fabregat-Santiago, and Felipe A. La Porta. 2023. "Electronic, Structural, Optical, and Electrical Properties of CsPbX3 Powders (X = Cl, Br, and I) Prepared Using a Surfactant-Free Hydrothermal Approach" Nanomanufacturing 3, no. 2: 217-227. https://doi.org/10.3390/nanomanufacturing3020013

Article Metrics

Back to TopTop