Next Article in Journal
An Effective Resistive-Type Alcohol Vapor Sensor Using One-Step Facile Nanoporous Anodic Alumina
Previous Article in Journal
Surface Modification of Silicon Carbide Wafers Using Atmospheric Plasma Etching: Effects of Processing Parameters
Previous Article in Special Issue
Sn-Based Perovskite Solar Cells towards High Stability and Performance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

All-Inorganic CsPbBr3 Perovskite Nanocrystals Synthesized with Olive Oil and Oleylamine at Room Temperature

1
Faculty of Materials Science and Engineering, Jimma Institute of Technology, Jimma University, Jimma P.O. Box 378, Ethiopia
2
Department of Physics, College of Natural and Computational Science, Mizan-Tepi University, Mizan P.O. Box 260, Ethiopia
3
Department of Physics, Baselius College, Kottayam 686001, India
4
School of Energy Materials, Mahatma Gandhi University, Kottayam 686560, India
5
Department of Materials Science and Engineering, Adama Science and Technology University, Adama P.O. Box 1888, Ethiopia
6
Center of Advanced Materials Science and Engineering, Adama Science and Technology University, Adama P.O. Box 1888, Ethiopia
*
Author to whom correspondence should be addressed.
Micromachines 2023, 14(7), 1332; https://doi.org/10.3390/mi14071332
Submission received: 9 June 2023 / Revised: 24 June 2023 / Accepted: 28 June 2023 / Published: 29 June 2023
(This article belongs to the Special Issue Perovskite Semiconductors for Light-Emitting Diodes and Beyond)

Abstract

:
Room temperature (RT) synthesis of the ternary cesium lead bromide CsPbBr3 quantum dots with oleic acid and oleylamine ligands was developed by Zeng and coworkers in 2016. In their works, the supersaturated recrystallization (SR) was adopted as a processing method without requiring inert gas and high-temperature injection. However, the oleic acid ligand for haloplumbate is known to be relatively unstable. Hence, in this work, we employed the eco-friendly olive oil to replace the oleic acid portion for the SR process at RT. Resultantly, we found that the cube-shaped nanocrystal has a size of ~40–42 nm and an optical bandgap of ~2.3 eV independent of the surface ligands, but the photoluminescence lifetime (τav) and crystal packing are dependent on the ligand species, e.g., τav = 3.228 ns (olive oil and oleylamine; here less ordered) vs. 1.167 ns (oleic acid and oleylamine). Importantly, we explain the SR mechanism from the viewpoint of the classical LaMer model combined with the solvent engineering technique in details.

1. Introduction

Halide perovskites (HPs) have the general chemical structure of ABX3, where A (methylammonium (MA) CH3NH3+, formamidinium (FA) CH(NH2)2+, methylhydrazinium CH3(NH2)2+, aziridinium (CH2)2NH2+ and cesium Cs+) and B (lead Pb2+, tin Sn2+, and germanium Ge2+) are cations, and X (chlorine Cl, bromine Br, and iodine I) is an anion [1,2,3,4,5,6,7]. When these HPs serve as a semiconducting layer for the electronic and optoelectronic devices, they share some common properties such as defect tolerance, large absorption coefficient, high dielectric constant, low exciton binding energy, long charge carrier diffusion length, high charge mobility, and tunable bandgap [8,9,10,11,12,13]. Hence, the state-of-the-art devices show the power conversion efficiency of ~26% for photovoltaic (PV) cells, and more than 20% quantum efficiency for light-emitting diodes (LEDs) [4,14,15,16,17]. Furthermore, based on the excellent aforementioned properties, HPs have been applied to lasers, photodetectors, X-ray/γ-ray detectors, medical imaging devices, biosensors, photocatalysis, and luminescence solar concentrators [18,19,20,21,22,23,24,25]. However, the organic–inorganic hybrid HPs have an intrinsic problem in thermal stability originating from the organic cations such as MA and FA in the perovskite structure [26,27]. For example, MABr decomposes at around 250–350 °C [27]. Hence, to overcome this thermal instability, all-inorganic HPs such as CsPbX3 (X = Cl, Br, and I) have been employed for perovskite electronics [27,28,29,30,31,32,33,34].
However, it is notable that all-inorganic Cs-based perovskites with a wide bandgap (Eg > 1.65 eV) [35,36] also have some stability issues. First, the Goldschmidt tolerance factor ( t ) is ( r A + r X ) / { 2 ( r B + r X ) } , in which r A ,   r B ,   and   r X are the radius of cation A, cation B, and anion X, respectively [37]. In the case of CsPbCl3, CsPbBr3, and CsPbI3, each tolerance factor is 0.87, 0.816, and 0.805, respectively [38,39]. Considering that the cubic phase could be stable in the range of 0.813 ≤ t ≤ 1.107 [37], CsPbI3 is structurally unstable among the cesium lead halides. Second, the Cs-based perovskites undergo a multiple polymorph transition, such as orthorhombic, tetragonal, and cubic, depending on the temperature and composition [40]. To be a thermodynamically stable cubic phase with high symmetry, the perovskite should stay at the temperature more than 47 °C for CsPbCl3, 130 °C for CsPbBr3, and 300 °C for CsPbI3, indicating that CsPbI3 is the most challenging compound in spite of its promising energy bandgap ~1.73 eV [41,42,43,44]. Note that CsPbCl3 and CsPbBr3 each have a bandgap of ~3.03 eV and ~2.25 eV, respectively [45]. Furthermore, it is notable that CsPbI3 undergoes a phase transition from the black γ-phase to the yellow δ-phase (i.e., a non-perovskite structure with Eg~2.82 eV), resulting in an undesirable compound for perovskite electronics [46]. Third, according to Li and coworkers, the thermal stability of Cs-based perovskites is in the sequence of CsPbCl3 > CsPbBr3 > CsPbI3. This is because the structural change was observed at 500 °C for CsPbCl3, 400 °C for CsPbBr3, and 200 °C for CsPbI3, respectively, which is based on in situ X-ray diffraction experiments as a function of temperature in ambient conditions [47,48]. Hence, we notice that CsPbCl3 and CsPbBr3 are superior to CsPbI3 from the stability point of view. In this work, we may focus on the green-emitting CsPbBr3 perovskite semiconductor [49,50].
In 1958, Møller reported the structure and photoconductivity of CsPbX3 crystals (X = Cl, Br, and I) [28]. Then, after more than five decades, in 2015, Kovalenko and coworker invented a hot injection (HI) method for synthesizing CsPbX3 (X = halide or its mixture) colloidal quantum dot materials with oleic acid (OA) and oleylamine (OAm) surface ligands [51]. Then, after one year, Zeng and coworker invented another method called supersaturated recrystallization (SR) for synthesizing CsPbX3 nanocrystals (NCs) with OA and OAm at room temperature [52]. In 2019, Shi and coworker tried to replace OA with an edible ‘olive oil (OO)’ using the HI method [53]. Then recently, Jing and coworker tried to replace OA with another ligand called 4,4′-Azobis(4-cyanovalericacid) (CA) at room temperature because OA and OAm can be detached easily from the surface of CsPbX3 [54]. On the other hand, in 2022, Ray and coworkers proposed a new technique to synthesize CsPbX3 (X = I/Br) by modifying the ligand chemistry using olive oil [55].
For fabricating a relatively more stable cubic phase of CsPbX3, there are three methods: (1) the synthesis of NCs with high surface energies, (2) surface ligand engineering, and (2) composition engineering [51,52,53,54,55]. These methods have been commonly utilized for synthesizing CsPbX3 QDs. Hence, the colloidal CsPbX3 NCs have several strong advantages compared to its bulk counterparts. First, based on quantum size and confinement effect, the energy bandgap of QD materials is easily tunable resulting in the full-color spectrum coverage. Second, the NC–ligand complexes can afford versatile solution processes via colloidal chemistry. Third, the perovskite NCs can enhance their optical properties, resulting in a high photoluminescence quantum yield (PLQY) and a narrow full width at half maximum (FWHM) for color purity, which is desirable for lighting and display applications [51,52].
In this study, we tried to replace OA with OO to synthesize CsPbBr3 NCs at room temperature for which we adopted the SR method. To the best of authors’ knowledge, in the past, the OO ligand was tested only by the HI method [53], although OO is promising as a surface ligand for the mass production process in future. Hence, in this work, we employed this OO ligand for synthesis of CsPbBr3 NCs at room temperature. Note that SR is a type of ligand-assisted reprecipitation (LARP) technique, affording a simple, low cost, and a large production capability in ambient conditions. Furthermore, basically, SR is an antisolvent-mediated solution process, resulting in the nucleation and growth of perovskite NCs. Hence, the classical LaMer model [56,57] and solvent engineering (SE) [58,59] should be very important for addressing the SR mechanism in the synthesis of CsPbBr3 NCs at room temperature.

2. Materials and Methods

2.1. Chemicals

Lead bromide (PbBr2, 99.0%, AR chemicals, Delhi, India), cesium bromide (CsBr, 99.9%, Sigma-Aldrich, Darmstadt, Germany), oleic acid (OA, 98%, Sigma-Aldrich, Darmstadt, Germany), oleylamine (OAm, technical grade 70%, Sigma-Aldrich, Darmstadt, Germany), olive oil (OO, Nice, Kochi, India), hexane (≥97.5%, Sigma-Aldrich, Darmstadt, Germany), N,N-dimethylformamide (DMF, 99.5%, AR chemicals, Delhi, India), and toluene (≥99.5%, AR chemicals, Delhi, India) were used in this experiment. All chemicals were used without further purification.

2.2. Synthesis of CsPbBr3 Nanocrystals

In the SR method [52], the perovskite precursors and ligands, CsBr (0.2 mmol, 43 mg), PbBr2 (0.2 mmol, 73 mg), OA or OO (0.5 mL), and 0.25 mL of oleylamine, were dissolved into 5 mL of a polar DMF solvent and stirred using magnetic stirrers at RT for two hours. Then, 1 mL of the perovskite precursor solution was titrated into 10 mL of a nonpolar toluene antisolvent and stirred for 60 s. Then, the CsPbBr3 NC dispersion exhibited a green-light emission under 365 nm UV illuminations, confirming that the reaction was successful through the SR method. Then, this colloidal dispersion was transferred to a centrifuge tube for centrifugations, and it was centrifuged at 8000 rpm for 10 min. The supernatant was discarded, and the resulting precipitate was re-dispersed in the nonpolar hexane. Then, the final CsPbBr3 sample was dried under a vacuum oven at 60 °C overnight. Finally, the CsPbBr3 samples were kept for further characterization at RT.

2.3. Characterization

The structural properties of CsPbBr3 (OA and OAm) and CsPbBr3 (OO and OAm) were determined by the X-ray diffraction (XRD) measurements measured using the Rigaku mini flex-300/600 diffractometer (Tokyo, Japan). The irradiation used CuKα, λ = 1.5406 Å at 40 kV and 15 mA for the XRD analysis. The CsPbBr3 NC’s size and shape were determined by using the high-resolution transmission electron microscopy (HR-TEM) (Model: JEOL, JEM-2100) with an operating voltage of 200 kV. The optical absorption was characterized by using the Ultraviolet-visible (UV–vis) absorption spectroscopy (SHIMADZU UV-2600, Kyoto, Japan), whereas the photoluminescence (PL) spectra were measured using a spectrophotometer (SHIMADZU RF-6000, Kyoto, Japan) at the excitation wavelength of 375 nm. The Fourier transform infrared spectroscopy (FT-IR) analysis was performed by using the PerkinElmer Spectrum Two FT-IR Spectrometer. For this purpose, the attenuated total reflection (ATR) was employed as a sampling technique to record the FT-IR spectra of the drop-cast CsPbBr3 thin film in the range 4000–400 cm−1 with resolution of 4 cm−1. Here, for the drop-cast thin films, the glass substrate was washed in detergent soap, followed by ultra-sonication using distilled water, acetone, and isopropanol, respectively, for 15 min per each process. Then, the cleaned substrate was dried at 80 °C for 1 h in vacuum oven and subsequently cooled down at room temperature. The prepared sample solution (10 mg perovskite precursors in 1 mL of toluene) was drop cast on the clean glass substrate and dried in ambient conditions. The PL decay curves were recorded by using a time-correlated single photon counting (TCSPC) (Model: Fluorolog 3 TCSPC, Horiba, Irvine, CA, USA).

2.4. Computational Methods

The electronic band structures of the CsPbBr3 unit cell were calculated using the Cambridge Serial Total Energy Package (CASTEP, Materials Studio 2017) software with the density functional theory. The Perdew–Burke–Ernzerhof (PBE) parametrization of the General Gradient Approximation (GGA) was used to describe the exchange correlation functional. The unit cell in the Brillouin zone was applied to evaluate the electronic band structures. Energy of 5 × 10−5 eV/atom, maximum force of 0.01 eV/Å, maximum displacement of 5 × 10−4 Å, and maximum stress of 0.02 GPa were used for all geometry optimization.

3. Results and Discussion

Figure 1a shows the orthorhombic structure (space group, P b n m ) of CsPbBr3 crystal. However, when temperature increased, CsPbBr3 undergoes a phase transition from orthorhombic to tetragonal ( P 4 / m b m ) at 88 °C and from tetragonal to cubic ( P m 3 ¯ m ) at 130 °C through a tiny rearrangement of the unit cell atoms [60,61]. On the other hand, in the case of NCs, the unit cell structure could be affected not only by crystal size/shape, but also by the Pb2+/ligand ratio [34]. For example, the cubic phase was observed at RT if the crystal size is in nanoscale, indicating the metastable state of NCs. In the case of the electronic structures of CsPbBr3 unit cell, the information could be found in the Supplementary Materials (SM), Figure S1, displaying that the bandgap (Eg) is reduced with from 2.40 eV (cubic) to 2.31 eV (orthorhombic) via 2.37 eV (tetragonal).
Figure 2 shows the chemical structure of (a) OO, (b) OAm, and (c) DMF and toluene, among which both organic OO and OAm surface ligands were employed as a mixed ligand for the synthesis of CsPbBr3 at RT using the SR process [52]. As shown in Figure 2a, the OO’s glycerol fractions (~98% of total OO) are composed of oleic acid, linoleic acid, palmitic acid, and palmitoleic acid, but it also contains small amounts of non-glycerol fraction (~1–2% of the total OO by weight), such as biophenols and triterpenic acids [62]. Here, because of the conformational freedom (e.g., a chain folding) of the C-C single-bond rotation, the high carbon ligand may undergo a chain folding in a typical solvent medium, DMF. Importantly, it is noteworthy that the compositional engineering (or mixing strategy) has been commonly used for perovskite electronics for enhancing stability, e.g., cation mixtures (MA+/FA+), metal ion mixtures (Pb2+/Sn2+ or Ge2+/Sn2+), halogen mixtures (Cl/Br/I), perovskite dimensional mixtures (2D/3D), etc., which resulting in the enhancement of stability of materials and devices [62,63,64,65,66,67]. In the same vein, when we use the edible OO as a surface ligand, it might be useful in stability with a guaranteed benefit of eco-friendliness, which is a desirable characteristic for a mass production process in future.
Figure 3a shows the free energy change ( Δ G ) as a function of the radius ( r ) when the nucleus is assumed to be a spherical according to the LaMer model, a classical nucleation theory [56,57]. Here, the free energy change ( Δ G h o m ) for homogenous nucleation and that ( Δ G h e t ) for heterogeneous nucleation could be defined as follows,
Δ G h o m = 4 3 π r 3 Δ G v + 4 π r 2 γ S L
Δ G h e t = { 4 3 π r 3 Δ G v + 4 π r 2 γ S L } ( 2 + cos θ ) ( 1 cos θ ) 2 4
where Δ G v is the free energy change associated with the formation of a small volume of solid, whereas γ S L is the solid/liquid interfacial free energy. Furthermore, Δ G h e t = Δ G h o m S ( θ ) , where S ( θ ) = ( 2 + cos θ ) ( 1 cos θ ) 2 / 4 is a shape factor at a wetting angle θ . According to the LaMer model [57], when a solution is supersaturated by undercooling, a nucleus could be formed. If this nucleus is larger than the critical size ( r * ), the spontaneous growth is available because Δ G is decreasing as shown in Figure 3a (see the blue solid line). In the case of the SR process, a supersaturated condition is directly available when the perovskite precursor solution (a colloidal dispersion containing bromide plumbate, PbBr n 2 - n with n = 2–6) is dropped into the antisolvent medium, e.g., toluene with Gutmann’s number (DN) ~0.1 kcal/mol, i.e., Lewis basicity (see Table 1 for details) [68]. Here, it is noteworthy that a nonpolar toluene with solubility parameter, δ = 8.9 (cal/cm3)1/2 is antisolvent for the bromide plumbate but it is still miscible with the polar DMF solvent (δ = 12.1 (cal/cm3)1/2) [69], indicating that the perovskite precursor ions (Cs+, Pb2+, Br) will be directly exposed to the antisolvent toluene molecules, resulting in the nucleation and growth of perovskite NCs as explained through the critical point (r*–G*) in the LaMer model. However, it is notable that in the case of SR process, the supersaturation was reached simply by adding the precursor solution into the antisolvent (not by undercooling like in the LaMer model). Interestingly, the SR process is similar to the well-known Solvent Engineering (SE) (i.e., one-step antisolvent method) in the sense that the wet precursor solution is directly exposed to the antisolvent molecules [58,59]. However, the difference is that the SE method is processed during the spin-coating step by dispensing the antisolvent on the top of a wet precursor solution. Furthermore, SE allows for the formation of intermediate phase when a polar solvent DMSO is employed, whereas SR may not allow for it because the colloidal dispersion may undergo a direct crystallization (at least without a co-solvent such as DMSO). Importantly, the SR process corresponds to the heterogeneous nucleation and growth of CsPbBr3 perovskite because the nanoscale crystallization was processed through the surface of haloplumbate dispersed in the DMF medium. Recall that the perovskite precursor solution is a typical colloidal dispersion [70], i.e., haloplumbate is dispersed in a solvent medium, herein DMF with DN = 26.6 kcal/mol. In the case of Br-, its DN is 33.7 kcal/mol [71], indicating that Pb2+ (Lewis acid) may have a stronger coordination bonding with Br ions (Lewis base) than DMF to form the acid–base adducts in spite that DMF is a good solvent for the perovskite precursor materials. See Figure S2 in the Supplementary Materials for the vials containing CsPbBr3 NCs in the supersaturated solution/dispersion.
Figure 4a,b show the experimental XRD patterns for the drop-cast CsPbBr3 thin film on the glass substrate depending on the surface ligand species: (Figure 4a) OA and OAm and (Figure 4b) OO and OAm, whereas Figure 4c,d shows the XRD patterns for (Figure 4c) cubic and (Figure 4d) orthorhombic based on Joint Committee on Powder Diffraction Standards (JCPDS). As shown in Figure 4a–c, the location of XRD peaks such as (100), (110), and (200) are more corresponding to that of cubic (JCPDS: PDF#96-153-0682: space group P m 3 ¯ m ) instead of orthorhombic in spite that it was synthesized at RT [29,55,72,73]. Note that the lattice constant is 5.6050 Å based on the above JCPDS data, which was confirmed by the d-spacing of ~5.478 Å or 5.418 Å = λ / ( 2 sin θ ) at θ = 8.08°(OA) or 8.17° (OO), respectively, when the X-ray wavelength (λ) is 1.5406 Å. Probably, the CsPbBr3 NC may prefer to exist in cubic phase due to its high surface energy and capping ligand effects. Furthermore, all the drop-cast CsPbBr3 nanocrystalline thin films on the glass substrate display two intense peaks at 2θ = 16° and 32°, corresponding to (100) and (200) crystallographic planes, respectively, indicating that there is orientational order to the (100) direction. However, when the OO and OAm ligands were incorporated as a ligand system for CsPbBr3, the additional (110) peak is clearly observed, indicating that the mixed ligand system by OO’s glycerol fraction makes the orientational order be reduced, resulting in a partial amorphous halo (see the baseline shape centered around (110) peak). However, when we estimated the crystallite size (t) by Scherrer equation (Equation (3)), it was in the range of ~40–42 nm for all the samples (Table 2).
t = 0.9 λ B cos θ
where λ (= 0.154 nm) is the X-ray wavelength, and B is the full width at half maximum (FWHM) at the angle θ. Here, it is noteworthy is that the estimated crystallite size is a direct particle size of this single crystal in nanoscale, which will be further addressed in the below section.
Figure 5 shows HR-TEM images for CsPbBr3 depending on the surface ligand system. Figure 5a,b show the case of OA and OAm, whereas Figure 5c,d show the case of OO and OAm. As shown in Figure 5, the TEM images show roughly a nano cubic, which is in line with the XRD data exhibiting the cubic structure.
Figure 6 shows (a) the UV-vis absorption and (b) the PL emission spectra for the two different colloidal dispersions containing CsPbBr3 NCs depending on the ligand species. Here, it is noteworthy that the CsPbBr3 perovskite is a direct semiconductor. In addition, the synthesized CsPbBr3 NCs do not display any grain boundary in Figure 5, indicating a single crystalline nature in nanoscale although there might be some surface defects. Therefore, the CsPbBr3 NCs are different from the disordered conjugated polymers showing a localized state. For disordered system, Mott defined a mobility edge between localized and delocalized states [74]. However, in the case of CsPbBr3 NCs, the sharp absorption edge is expected, affording the decision of optical bandgap by using a simple tangential slope based on the equation, E = h c / λ , where E is energy, h is Planck’s constant, c is the speed of light, and λ is the wavelength of light. Therefore, according to the UV-Vis absorption edge (530 nm vs. 533 nm) in Figure 6a, the corresponding optical bandgap (Eg) is 2.34 eV = 1240/530 (for OA and OAm) and 2.33 eV = 1240/533 (for OO and OAm), respectively. See Equation (4) for the conversion factor, 1240, between the energy (bandgap) and the wavelength (absorption edge).
λ [ nm ] = h c E = ( 6.62607 × 10 34   J × s ) × ( 3 × 10 8   m / s ) ( 10 9   nm / 1   m ) E g ( eV ) × ( 1.60218 × 10 19   J / 1   eV ) = 1240 1 E [ eV ]
This result indicates that the nanoparticle size might be slightly larger in CsPbBr3 synthesized with OO and OAm. This result is in line with the estimated crystallite size trend according to the sharpest peak at the (200) crystallographic plane in Figure 4a and Table 2. On the other hand, the PL spectra exhibit the PL peak at 517 nm for CsPbBr3 with oleic acid and oleylamine and at 526 nm for CsPbBr3 with OO and OAm. Here, the shift of PL peak from 517 nm to 526 nm (a red shift) indicates that the effective crystal size (or aggregation) of CsPbBr3 NCs is slightly larger in the colloidal dispersion state when OO and OAm were employed as the surface ligands, although the apparent size (Table 2) is similar for both ligand systems in the solid state. Furthermore, it is believed that the surface state of NCs is different depending on the NC–ligand complexations. Hence, when OO and OAm were complexed with CsPbBr3 NCs, there might be more surface defects, resulting in more non-radiative recombination (i.e., smaller PL intensity) in Figure 6b. Furthermore, FWHM is 21 nm for the former, but 19 nm for the latter. According to our finding, the narrow emission spectra were obtained below 25 nm, which falls in the required range of 12–42 nm [51] for the optoelectronic applications such as lighting and display technology.
Figure 7a shows PL decay curve for the colloidal CsPbBr3 NC dispersion as a function of ligand species. The average lifetime of electron–hole pairs (Wannier–Mott excitons) is 1.167 ns for CsPbBr3 with OA and OAm, whereas it is 3.228 ns for CsPbBr3 with OO and OAm, indicating that a slightly larger CsPbBr3 with OO and OAm displays a longer life time of excitons. Figure 7b shows the FT-IR spectra for the drop-cast CsPbBr3 nanocrystalline thin film. First of all, the IR spectrum for CsPbBr3 with OA and OAm is relatively simpler than that for CsPbBr3 with OO and OAm. Second, the FT-IR spectrum for OO and OAm shows a high absorption at 3020–2850 cm−1, whereas the corresponding spectrum for OA and OAm is very weak at those wavenumbers, which correspond to the symmetric and asymmetric C-H stretching modes [53,54]. The functional groups of other organic components can be attributed to the absorption peaks at 3781 cm−1 [54] and 3510 cm−1 [53], which originate from the N-H characteristic absorption. The IR peaks at around 1800–1600 cm−1 are related to carbonyl (C=O) mode of stretching vibration which can be shift to lower wavenumber due to any perturbation. The shift of wavenumber may be ascribed to the presence of hydrogen bonding. The peaks at 1538 cm−1 and 1462 cm−1 are related to C-CH3 asymmetric bending. On the other hand, the peaks at 1377 cm−1 and 1262 cm−1 correspond to C-CH3 symmetric bending vibration and C-N stretching vibration, respectively. Absorption peaks in the range of 1000–500 cm−1 indicate that the existence of aromatic bending vibrations [75], which might originate from OO components, trapped toluene (C6H5CH3), and/or aromatic impurities in samples. The peak at 477–447 cm−1 is associated with CsPbBr3–OAm complexes [76].

4. Conclusions

In this work, we demonstrated the usefulness of olive oil and oleylamine as the surface ligand system (vs. the typical oleic acid and oleylamine mixtures) for the synthesis of CsPbBr3 nanocrystals at room temperature. For this purpose, the supersaturated recrystallization process was employed, enabling a direct synthesis of perovskite nanocrystals. First, according to XRD, the crystallite size is ~40–42 nm in both conditions (oleic acid and oleylamine vs. olive oil and oleylamine), whereas the latter system allows for a less orientational order than the former by showing a partial amorphous halo. The TEM images confirmed that the CsPbBr3 nanocrystals have a cuboid shape in line with XRD data exhibiting a cubic phase. The UV-Vis absorption data exhibited that the optical bandgap of CsPbBr3 is about 2.3 eV in both ligand systems. The PL emission spectra indicated that FWHM is about 19 nm (olive oil and oleylamine) vs. 21 nm (oleic acid and oleylamine), which falls into the desirable FWHM values, i.e., ~12–42 nm for technical applications. Finally, the average PL lifetime is 3.228 ns for olive oil and oleylamine and 1.167 ns for oleic acid and oleylamine. Future work may include the application of CsPbBr3 nanocrystals to the biosensors and optoelectronic devices.

Supplementary Materials

The following are available online at: https://www.mdpi.com/article/10.3390/mi14071332/s1, Figure S1: The electronic structures of CsPbBr3 unit cells displaying polymorphism: (a) orthorhombic, (b) tetragonal, and (c) cubic phase. Here, the bandgap energy (Eg) of each unit cell is 2.31 eV for orthorhombic, 2.37 eV for tetragonal, and 2.40 eV for cubic phase, respectively. Figure S2: Vials containing a CsPbBr3 colloidal dispersion depending on surface ligands: (a) oleic acid and oleylamine—a normal photo, (b) oleic acid and oleylamine—under 365 nm UV light, (c) olive oil and oleylamine—a normal photo, and (d) olive oil and oleylamine—under 365 nm UV light. Hence, CsPbBr3 nanocrystals can serve as a green-light emitter.

Author Contributions

Writing—original draft preparation, G.W.; writing—review and editing, J.Y.K.; conceptualization, G.W., M.A., D.M., A.T., S.T. and J.Y.K.; methodology, G.W. and J.Y.K.; software, G.W. and F.G.A.; formal analysis and investigation, G.W. and J.Y.K.; resources, S.T.; data curation, G.W.; supervision and project administration, M.A., D.M., A.T., S.T. and J.Y.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The datasets used and/or analyzed during the current study are available from the corresponding author upon reasonable request.

Acknowledgments

Jimma University and Mizan-Tepi University are acknowledged for supporting G.W.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kojima, A.; Teshima, K.; Shirai, Y.; Miyasaka, T. Organometal halide perovskites as visible-light sensitizers for photovoltaic cells. J. Am. Chem. Soc. 2009, 131, 6050–6051. [Google Scholar] [CrossRef] [PubMed]
  2. 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] [PubMed] [Green Version]
  3. Stranks, S.D.; Eperon, G.E.; Grancini, G.; Menelaou, C.; Alcocer, M.J.P.; Leijtens, T.; Herz, L.M.; Petrozza, A.; Snaith, H.J. Electron-Hole Diffusion Lengths Exceeding 1 Micrometer in an Organometal Trihalide Perovskite Absorber. Science 2013, 342, 341–344. [Google Scholar] [CrossRef] [Green Version]
  4. Min, H.; Kim, M.; Lee, S.-U.; Kim, H.; Kim, G.; Choi, K.; Lee, J.H.; Seok, S.I. Efficient, stable solar cells by using inherent bandgap of a-phase formamidinium lead iodide. Science 2019, 366, 749–753. [Google Scholar] [CrossRef]
  5. Zhou, Y.; Herz, L.M.; Jen, A.K.-Y.; Saliba, M. Advances and challenges in understanding the microscopic structure–property–performance relationship in perovskite solar cells. Nat. Energy 2021, 7, 794–807. [Google Scholar] [CrossRef]
  6. Mączka, M.; Ptak, M.; Ga̧gor, A.; Stefańska, D.; Zarȩba, J.K.; Sieradzki, A. Methylhydrazinium Lead Bromide: Noncentrosymmetric Three-Dimensional Perovskite with Exceptionally Large Framework Distortion and Green Photoluminescence. Chem. Mater. 2020, 32, 1667–1673. [Google Scholar] [CrossRef]
  7. Stefańska, D.; Ptak, M.; Ma̧czka, M. Synthesis, Photoluminescence and Vibrational Properties of Aziridinium Lead Halide Perovskites. Molecules 2022, 27, 7949. [Google Scholar] [CrossRef]
  8. Brivio, F.; Walker, A.B.; Walsh, A. Structural and electronic properties of hybrid perovskites for high-efficiency thin-film photovoltaics from first-principles. APL Mater. 2013, 1, 042111. [Google Scholar] [CrossRef] [Green Version]
  9. Stoumpos, C.C.; Malliakas, C.D.; Kanatzidis, M.G. Semiconducting Tin and Lead Iodide Perovskites with Organic Cations: Phase Transitions, High Mobilities, and Near-Infrared Photoluminescent Properties. Inorg. Chem. 2013, 52, 9019–9038. [Google Scholar] [CrossRef]
  10. Wolf, S.D.; Holovsky, J.; Moon, S.-J.; Löper, P.; Niesen, B.; Ledinsky, M.; Haug, F.-J.; Yum, J.-H.; Ballif, C. Organometallic Halide Perovskites: Sharp Optical Absorption Edge and Its Relation to Photovoltaic Performance. J. Phys. Chem. Lett. 2014, 5, 1035–1039. [Google Scholar] [CrossRef]
  11. 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] [PubMed]
  12. Herz, L.M. Charge-Carrier Mobilities in Metal Halide Perovskites: Fundamental Mechanisms and Limits. ACS Energy Lett. 2017, 2, 1539–1548. [Google Scholar] [CrossRef] [Green Version]
  13. Huang, J.; Yuan, Y.; Shao, Y.; Yan, Y. Understanding the physical properties of hybrid perovskites for photovoltaic applications. Nat. Rev. Mater. 2017, 2, 17042. [Google Scholar] [CrossRef]
  14. Zhao, B.; Bai, S.; Kim, V.; Lamboll, R.; Shivanna, R.; Auras, F.; Richter, J.M.; Yang, L.; Dai, L.; Alsari, M.; et al. High-efficiency perovskite-polymer bulk heterostructure light-emitting diodes. Nat. Photon. 2018, 12, 783. [Google Scholar] [CrossRef] [Green Version]
  15. Mathews, I.; Sofia, S.; Ma, E.; Jean, J.; Laine, H.S.; Siah, S.C.; Buonassisi, T.; Peters, I.M. Economically Sustainable Growth of Perovskite Photovoltaics Manufacturing. Joule 2020, 4, 822–839. [Google Scholar] [CrossRef]
  16. Sun, C.; Jiang, Y.; Cui, M.; Qiao, L.; Wei, J.; Huang, Y.; Zhang, L.; He, T.; Li, S.; Hsu, H.-Y.; et al. High-performance large-area quasi-2D perovskite light-emitting diodes. Nat. Commun. 2021, 12, 2207. [Google Scholar] [CrossRef]
  17. Park, J.; Kim, J.; Yun, H.-S.; Paik, M.J.; Noh, E.; Mun, H.J.; Kim, M.G.; Shin, T.J.; Seok, S.I. Controlled growth of perovskite layers with volatile alkylammonium chlorides. Nature 2023, 616, 724–730. [Google Scholar] [CrossRef]
  18. Zhang, C.; Kuang, D.-B.; Wu, W.-Q. A Review of Diverse Halide Perovskite Morphologies for Efficient Optoelectronic Applications. Small Methods 2020, 4, 1900662. [Google Scholar] [CrossRef]
  19. Pan, L.; Pandey, I.R.; Miceli, A.; Klepov, V.V.; Chung, D.Y.; Kanatzidis, M.G. Perovskite CsPbBr3 Single-Crystal Detector Operating at 1010 Photons s−1 mm−2 for Ultra-High Flux X-ray Detection. Adv. Optical. Mater. 2023, 11, 2202946. [Google Scholar] [CrossRef]
  20. Jeong, B.; Veith, L.; Smolders, T.J.A.M.; Wolf, M.J.; Asadi, K. Room-Temperature Halide Perovskite Field-Effect Transistors by Ion Transport Mitigation. Adv. Mater. 2021, 33, 2100486. [Google Scholar] [CrossRef]
  21. Zhang, Q.; Shang, Q.; Su, R.; Do, T.T.H.; Xiong, Q. Halide Perovskite Semiconductor Lasers: Materials, Cavity Design, and Low Threshold. Nano Lett. 2021, 21, 1903–1914. [Google Scholar] [CrossRef] [PubMed]
  22. Zhao, L.; Zhou, Y.; Shi, Z.; Ni, Z.; Wang, M.; Liu, Y.; Huang, J. High-yield growth of FACsPbBr3 single crystals with low defect density from mixed solvents for gamma-ray spectroscopy. Nat. Photon. 2023, 17, 315–323. [Google Scholar] [CrossRef]
  23. He, Y.; Petryk, M.; Liu, Z.; Chica, D.G.; Hadar, I.; Leak, C.; Ke, W.; Spanopoulos, I.; Lin, W.; Chung, D.Y.; et al. CsPbBr3 perovskite detectors with 1.4% energy resolution for high-energy γ-rays. Nat. Photon. 2021, 15, 36–42. [Google Scholar] [CrossRef]
  24. Zhang, X.; Peng, H.; Liu, J.; Yuan, Y. Highly sensitive plasmonic biosensor enhanced by perovskite-graphene hybrid configuration. J. Opt. 2023, 25, 075002. [Google Scholar] [CrossRef]
  25. Huang, H.; Pradhan, B.; Hofkens, J.; Roeffaers, M.B.J.; Steele, J.A. Solar-Driven Metal Halide Perovskite Photocatalysis: Design, Stability, and Performance. ACS Energy Lett. 2020, 5, 1107–1123. [Google Scholar] [CrossRef]
  26. Zhang, D.; Li, D.; Hu, Y.; Mei, A.; Han, H. Degradation pathways in perovskite solar cells and how to meet international standards. Commun. Mater. 2022, 3, 58. [Google Scholar] [CrossRef]
  27. Kulbak, M.; Gupta, S.; Kedem, N.; Levine, I.; Bendikov, T.; Hodes, G.; Cahen, D. Cesium Enhances Long-Term Stability of Lead Bromide Perovskite-Based Solar Cells. J. Phys. Chem. Lett. 2016, 7, 167–172. [Google Scholar] [CrossRef] [Green Version]
  28. Møller, C. Crystal Structure and Photoconductivity of Cæsium Plumbohalides. Nature 1958, 182, 1436. [Google Scholar] [CrossRef]
  29. Brinck, S.T.; Infante, I. Surface Termination, Morphology, and Bright Photoluminescence of Cesium Lead Halide Perovskite Nanocrystals. ACS Energy Lett. 2016, 1, 1266–1272. [Google Scholar] [CrossRef]
  30. Rakita, Y.; Kedem, N.; Gupta, S.; Sadhanala, A.; Kalchenko, V.; Böhm, M.L.; Kulbak, M.; Friend, R.H.; Cahen, D.; Hodes, G. Low-Temperature Solution-Grown CsPbBr3 Single Crystals and Their Characterization. Cryst. Growth Des. 2016, 16, 5717–5725. [Google Scholar] [CrossRef] [Green Version]
  31. He, X.; Qiu, Y.; Yang, S. Fully-Inorganic Trihalide Perovskite Nanocrystals: A New Research Frontier of Optoelectronic Materials. Adv. Mater. 2017, 29, 1700775. [Google Scholar] [CrossRef] [PubMed]
  32. Fang, F.; Chen, W.; Li, Y.; Liu, H.; Mei, M.; Zhang, R.; Hao, J.; Mikita, M.; Cao, W.; Pan, R.; et al. Employing Polar Solvent Controlled Ionization in Precursors for Synthesis of High-Quality Inorganic Perovskite Nanocrystals at Room Temperature. Adv. Funct. Mater. 2018, 28, 1706000. [Google Scholar] [CrossRef]
  33. Tenailleau, C.; Aharon, S.; Cohen, B.-E.; Etgar, L. Cell refinement ofCsPbBr3 perovskite nanoparticles and thin films. Nanoscale Adv. 2019, 1, 147–153. [Google Scholar] [CrossRef] [Green Version]
  34. Grisorio, R.; Fanizza, E.; Allegretta, I.; Altamura, D.; Striccoli, M.; Terzano, R.; Giannini, C.; Vergaro, V.; Ciccarella, G.; Margiotta, N.; et al. Insights into the role of the lead/surfactant ratio in the formation and passivation of cesium lead bromide perovskite nanocrystals. Nanoscale 2020, 12, 623–637. [Google Scholar] [CrossRef] [PubMed]
  35. Xu, F.; Zhang, M.; Li, Z.; Yang, X.; Zhu, R. Challenges and Perspectives toward Future Wide-Bandgap Mixed-Halide Perovskite Photovoltaics. Adv. Energy Mater. 2023, 13, 2203911. [Google Scholar] [CrossRef]
  36. Xu, F.; Demir, H.V. LEDs using halide perovskite nanocrystal emitters. Nanoscale 2019, 11, 11402–11412. [Google Scholar]
  37. Zhao, Y.; Zhu, K. Organic–inorganic hybrid lead halide perovskites for optoelectronic and electronic applications. Chem. Soc. Rev. 2016, 45, 655–689. [Google Scholar] [CrossRef] [PubMed]
  38. Zhu, X.; Ge, L.; Wang, Y.; Li, M.; Zhang, R.; Xu, M.; Zhao, Z.; Lv, W.; Chen, R. Recent Advances in Enhancing and Enriching the Optical Properties of Cl-Based CsPbX3 Nanocrystals. Adv. Opt. Mater. 2021, 9, 2100058. [Google Scholar] [CrossRef]
  39. Xiang, W.; Liu, S.F.; Tress, W. A review on the stability of inorganic metal halide perovskites: Challenges and opportunities for stable solar cells. Energy Environ. Sci. 2021, 14, 2090–2113. [Google Scholar] [CrossRef]
  40. Alaei, A.; Circelli, A.; Yuan, Y.; Yang, Y.; Lee, S.S. Polymorphism in metal halide perovskites. Mater. Adv. 2021, 2, 47–63. [Google Scholar] [CrossRef]
  41. Calistru, D.M.; Mihut, L.; Lefrant, S.; Baltog, I. Identification of the symmetry of phonon modes in CsPbCl3 in phase IV by Raman and resonance-Raman scattering. J. Appl. Phys. 1997, 82, 5391–5395. [Google Scholar] [CrossRef]
  42. 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] [Green Version]
  43. 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] [Green Version]
  44. Ahmad, M.; Rehman, G.; Ali, L.; Shafiq, M.; Iqbal, R.; Ahmad, R.; Khan, T.; Jalali-Asadabadi, S.; Maqbool, M.; Ahmad, I. Structural, electronic and optical properties of CsPbX3 (X = Cl, Br, I) for energy storage and hybrid solar cell applications. J. Alloys Compd. 2017, 705, 828–839. [Google Scholar] [CrossRef]
  45. He, Y.; Stoumpos, C.C.; Hadar, I.; Luo, Z.; McCall, K.M.M.; Liu, Z.; Chung, D.Y.; Wessels, B.W.; Kanatzidis, M.G. Demonstration of Energy-Resolved γ-Ray Detection at Room Temperature by the CsPbCl3 Perovskite Semiconductor. J. Am. Chem. Soc. 2021, 143, 2068–2077. [Google Scholar] [CrossRef]
  46. Eperon, G.G.; Paterno, G.M.; Sutton, R.J.; Zampetti, A.; Haghighirad, A.A.; Cacialli, F.; Snaith, H.J. Inorganic caesium lead iodide perovskite solar cells. J. Mater. Chem. A 2015, 3, 19688–19695. [Google Scholar] [CrossRef]
  47. 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]
  48. Boote, B.W.; Andaraarachchi, H.P.; Rosales, B.A.; Blome-Fernández, R.; Zhu, F.; Reichert, M.D.; Santra, K.; Li, J.; Petrich, J.W.; Vela, J.; et al. Unveiling the Photo- and Thermal-Stability of Cesium Lead Halide Perovskite Nanocrystals. Chem. Phys. Chem. 2019, 20, 2647–2656. [Google Scholar] [CrossRef]
  49. Stoumpos, C.C.; Malliakas, C.D.; Peters, J.A.; Liu, Z.; 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]
  50. Kobayashi, M.; Omata, K.; Sugimoto, S.; Tamagawa, Y.; Kuroiwa, T.; Asada, H.; Takeuchi, H.; Kondo, S. Scintillation characteristics of CsPbCl3 single crystals. Nucl. Instrum. Methods Phys. Res. A 2008, 592, 369–373. [Google Scholar] [CrossRef]
  51. 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] [Green Version]
  52. Li, X.; Wu, Y.; Zhang, S.; Cai, B.; Gu, Y.; Song, J.; Zeng, H. CsPbX3 Quantum Dots for Lighting and Displays: Room-Temperature Synthesis, Photoluminescence Superiorities, Underlying Origins and White Light-Emitting Diodes. Adv. Funct. Mater. 2016, 26, 2435–2445. [Google Scholar] [CrossRef]
  53. Xu, H.; Wang, J.; Xuan, T.; Lv, C.; Hou, J.; Zhang, L.; Dong, Y.; Shi, J. Convenient and large-scale synthesis of high-quality, all-inorganic lead halide perovskite nanocrystals for white light-emitting diodes. Chem. Eng. J. 2019, 364, 20–27. [Google Scholar] [CrossRef]
  54. Zhang, Y.; Li, G.; She, C.; Liu, S.; Yue, F.; Jing, C.; Cheng, Y.; Chu, J. Room temperature preparation of highly stable cesium lead halide perovskite nanocrystals by ligand modification for white light-emitting diodes. Nano Res. 2021, 14, 2770–2775. [Google Scholar] [CrossRef]
  55. Ghorai, A.; Mahato, S.; Srivastava, S.K.; Ray, S.K. Atomic Insights of Stable, Monodispersed CsPbI3−xBrx (x = 0, 1, 2, 3) Nanocrystals Synthesized by Modified Ligand Cell. Adv. Funct. Mater. 2022, 32, 2202087. [Google Scholar] [CrossRef]
  56. La Mer, V.K. Nucleation in Phase Transitions. Ind. Eng. Chem. 1952, 44, 1270–1277. [Google Scholar] [CrossRef]
  57. Jung, M.; Ji, S.-G.; Kim, G.; Seok, S.-I. Perovskite precursor solution chemistry: From fundamentals to photovoltaic applications. Chem. Soc. Rev. 2019, 48, 2011–2038. [Google Scholar] [CrossRef] [PubMed]
  58. Jeon, N.J.; Noh, J.H.; Kim, Y.C.; Yang, W.S.; Ryu, S.; Seok, S.I. Solvent engineering for high-performance inorganic–organic hybrid perovskite solar cells. Nat. Mater. 2014, 13, 897–903. [Google Scholar] [CrossRef]
  59. Xiao, M.A.; Huang, F.; Huang, W.; Dkhissi, Y.; Zhu, Y.; Etheridge, J.; Gray-Weale, A.; Bach, U.; Cheng, Y.-B.; Spiccia, L. 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]
  60. Rodová, M.; Brožek, J.; Knížek, K.; Nitsch, K. Phase transitions in ternary caesium lead bromide. J. Therm. Anal. Calorim. 2003, 71, 667–673. [Google Scholar] [CrossRef]
  61. Malyshkin, D.; Sereda, V.; Ivanov, I.; Mazurin, M.; Sednev-Lugovets, A.; Tsvetkov, D.; Zuev, A. New phase transition in CsPbBr3. Mater. Lett. 2020, 278, 128458. [Google Scholar] [CrossRef]
  62. Nikou, T.; Witt, M.; Stathopoulos, P.; Barsch, A.; Halabalaki, M. Olive Oil Quality and Authenticity Assessment Aspects Employing FIA-MRMS and LC-Orbitrap MS Metabolomic Approaches. Front. Public Health 2020, 8, 558226. [Google Scholar] [CrossRef] [PubMed]
  63. Noh, J.H.; Im, S.H.; Heo, J.H.; Mandal, T.N.; Seok, S.I. Chemical Management for Colorful, Efficient, and Stable Inorganic–Organic Hybrid Nanostructured Solar Cells. Nano Lett. 2013, 13, 1764–1769. [Google Scholar] [CrossRef]
  64. 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] [PubMed]
  65. Jang, Y.-W.; Lee, S.; Yeom, K.M.; Jeong, K.; Choi, K.; Choi, M.; Noh, J.H. Intact 2D/3D halide junction perovskite solar cells via solid-phase in-plane growth. Nat. Energy 2021, 6, 63–71. [Google Scholar] [CrossRef]
  66. Zhou, Y.-Q.; Xu, J.; Liu, J.-B.; Liu, B.-X. Alloy engineering in mixed Sn–Ge perovskites for photovoltaic application. J. Mater. Chem. A 2021, 9, 6955–6961. [Google Scholar] [CrossRef]
  67. Li, J.; Yan, N.; Fang, Z.; Liu, S.F. Alkyl Diamine-Induced (100)-Preferred Crystal Orientation for Efficient Pb–Sn Perovskite Solar Cells. ACS Appl. Energy Mater. 2022, 5, 6936–6942. [Google Scholar] [CrossRef]
  68. Gutmann, V. Solvent effects on the reactivities of organometallic compounds. Coord. Chem. Rev. 1976, 18, 225–255. [Google Scholar] [CrossRef]
  69. Belmares, M.; Blanco, M.; Goddard, W.A., III; Ross, R.B.; Caldwell, G.; Chou, S.-H.; Pham, J.; Olofson, P.M.; Thomas, C. Hildebrand and Hansen solubility parameters from molecular dynamics with applications to electronic nose polymer sensors. J. Comput. Chem. 2004, 25, 1814–1826. [Google Scholar] [CrossRef] [Green Version]
  70. Yan, K.; Long, M.; Zhang, T.; Wei, Z.; Chen, H.; Yang, S.; Xu, J. Hybrid Halide Perovskite Solar Cell Precursors: Colloidal Chemistry and Coordination Engineering behind Device Processing for High Efficiency. J. Am. Chem. Soc. 2015, 137, 4460–4468. [Google Scholar] [CrossRef]
  71. Petrov, A.A.; Ordinartsev, A.A.; Fateev, S.A.; Goodilin, E.A.; Tarasov, A.B. Solubility of Hybrid Halide Perovskites in DMF and DMSO. Molecules 2021, 26, 7541. [Google Scholar] [CrossRef]
  72. Huangfu, C.; Feng, L. High-performance fluorescent sensor based on CsPbBr3 quantum dots for rapid analysis of total polar materials in edible oils. Sens. Actuators B Chem. 2021, 344, 130193. [Google Scholar] [CrossRef]
  73. Chen, H.; Wang, Y.; Wang, J.; Liu, W. Thermal Stability of CsPbBr3 Perovskite Quantum Dots Assembled with SBA-15. Coatings 2021, 11, 953. [Google Scholar] [CrossRef]
  74. Mott, S.N. The mobility edge since 1967. J. Phys. C Solid State Phys. 1987, 20, 3075–3102. [Google Scholar] [CrossRef]
  75. Smith, B.C. Infrared Spectral Interpretation: A Systematic Approach; CRC Press: Boca Raton, FL, USA, 1999. [Google Scholar]
  76. Akhil, S.; Dutt, V.G.V.; Mishra, N. Surface modification for improving the photoredox activity of CsPbBr3 nanocrystals. Nanoscale Adv. 2021, 3, 2547–2553. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Polyhedral of CsPbBr3 unit cell: (a) orthorhombic, (b) tetragonal, and (c) cubic structure.
Figure 1. Polyhedral of CsPbBr3 unit cell: (a) orthorhombic, (b) tetragonal, and (c) cubic structure.
Micromachines 14 01332 g001
Figure 2. Chemical structures of (a) olive oil: glycerol fractions, ~98%; (b) oleylamine; and (c) DMF and toluene.
Figure 2. Chemical structures of (a) olive oil: glycerol fractions, ~98%; (b) oleylamine; and (c) DMF and toluene.
Micromachines 14 01332 g002
Figure 3. (a) The free energy change ( Δ G ) in the nucleation and growth process as a function of r (a radius of a spherical nucleus by assumption). Here, Δ G * is the critical free energy barrier at the critical radius ( r * ), where a unit cell is formed. (b) Schematic explanation of supersaturated recrystallization (SR) of CsPbBr3 in the antisolvent toluene.
Figure 3. (a) The free energy change ( Δ G ) in the nucleation and growth process as a function of r (a radius of a spherical nucleus by assumption). Here, Δ G * is the critical free energy barrier at the critical radius ( r * ), where a unit cell is formed. (b) Schematic explanation of supersaturated recrystallization (SR) of CsPbBr3 in the antisolvent toluene.
Micromachines 14 01332 g003
Figure 4. XRD patterns of CsPbBr3 nanocrystal drop-cast thin film depending on the ligand species: (a) oleic acid and oleylamine and (b) olive oil and oleylamine. (c) Joint Committee on Powder Diffraction Standards (JCPDS) (PDF#96-153-0682) and (d) JCPDS (PDF#96-153-3063).
Figure 4. XRD patterns of CsPbBr3 nanocrystal drop-cast thin film depending on the ligand species: (a) oleic acid and oleylamine and (b) olive oil and oleylamine. (c) Joint Committee on Powder Diffraction Standards (JCPDS) (PDF#96-153-0682) and (d) JCPDS (PDF#96-153-3063).
Micromachines 14 01332 g004
Figure 5. HR-TEM images of CsPbBr3 nanocrystals (nano cubic) when the ligand was oleic acid and oleylamine (a,b) and olive oil and oleylamine (c,d).
Figure 5. HR-TEM images of CsPbBr3 nanocrystals (nano cubic) when the ligand was oleic acid and oleylamine (a,b) and olive oil and oleylamine (c,d).
Micromachines 14 01332 g005
Figure 6. (a) UV-Vis absorbance spectra of the CsPbBr3 nanocrystal dispersion depending on the ligand species. Here, λ = 530 nm corresponds to Eg ≈ 2.34 eV (oleic acid and oleylamine) and λ = 533 nm corresponds to Eg ≈ 2.33 eV (olive oil and oleylamine). (b) PL spectra of CsPbBr3 nanocrystals depending on the ligand species.
Figure 6. (a) UV-Vis absorbance spectra of the CsPbBr3 nanocrystal dispersion depending on the ligand species. Here, λ = 530 nm corresponds to Eg ≈ 2.34 eV (oleic acid and oleylamine) and λ = 533 nm corresponds to Eg ≈ 2.33 eV (olive oil and oleylamine). (b) PL spectra of CsPbBr3 nanocrystals depending on the ligand species.
Micromachines 14 01332 g006
Figure 7. (a) PL spectra decay of CsPbBr3 nanocrystals depending on the ligand species. (b) FT-IR spectra of CsPbBr3 nanocrystals depending on the ligand species.
Figure 7. (a) PL spectra decay of CsPbBr3 nanocrystals depending on the ligand species. (b) FT-IR spectra of CsPbBr3 nanocrystals depending on the ligand species.
Micromachines 14 01332 g007
Table 1. Gutmann’s solvent donor number (DN), solubility parameter (δ), dielectric constant (ε), molar mass, density, and boiling point (b.p.), and the chemical structures of solvents.
Table 1. Gutmann’s solvent donor number (DN), solubility parameter (δ), dielectric constant (ε), molar mass, density, and boiling point (b.p.), and the chemical structures of solvents.
SolventDN
(kcal/mol)
δ 1
(cal/cm3)1/2
εMolar Mass
(g/mol)
Density
(g/cm3)
b.p.
(°C)
Chemical
Structure
DMF26.612.136.170.0950.948153C3H7NO
Toluene0.18.92.3892.1400.867111C₆H₅CH₃
1 In MKS unit, the δ value is 24. 8 MPa1/2 for DMF and 18.2 MPa1/2 for toluene, respectively.
Table 2. Crystallite size (t) of CsPbBr3 at the (200) crystallographic plane. Here, β denotes a full width at half maximum (FWHM).
Table 2. Crystallite size (t) of CsPbBr3 at the (200) crystallographic plane. Here, β denotes a full width at half maximum (FWHM).
CsPbBr3 Nanocrystals
Crystal Planes2θ (°)θ (°)β (Radian)t (nm)
Oleic Acid
(OA)
(100)16.158.080.00332342
(200)31.6515.830.00364140
Olive Oil
(OO)
(100)16.338.170.00343841
(200)31.8015.900.00343842
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

Welyab, G.; Abebe, M.; Mani, D.; Thankappan, A.; Thomas, S.; Aga, F.G.; Kim, J.Y. All-Inorganic CsPbBr3 Perovskite Nanocrystals Synthesized with Olive Oil and Oleylamine at Room Temperature. Micromachines 2023, 14, 1332. https://doi.org/10.3390/mi14071332

AMA Style

Welyab G, Abebe M, Mani D, Thankappan A, Thomas S, Aga FG, Kim JY. All-Inorganic CsPbBr3 Perovskite Nanocrystals Synthesized with Olive Oil and Oleylamine at Room Temperature. Micromachines. 2023; 14(7):1332. https://doi.org/10.3390/mi14071332

Chicago/Turabian Style

Welyab, Getachew, Mulualem Abebe, Dhakshnamoorthy Mani, Aparna Thankappan, Sabu Thomas, Fekadu Gochole Aga, and Jung Yong Kim. 2023. "All-Inorganic CsPbBr3 Perovskite Nanocrystals Synthesized with Olive Oil and Oleylamine at Room Temperature" Micromachines 14, no. 7: 1332. https://doi.org/10.3390/mi14071332

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