Previous Article in Journal
Solution-Processable and Eco-Friendly Functionalization of Conductive Silver Nanoparticles Inks for Printable Electronics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Silver-Doped CsPbI2Br Perovskite Semiconductor Thin Films

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, Bonga University, Bonga P.O. Box 334, 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.
Electron. Mater. 2024, 5(2), 56-70; https://doi.org/10.3390/electronicmat5020005
Submission received: 14 April 2024 / Revised: 9 May 2024 / Accepted: 11 May 2024 / Published: 13 May 2024

Abstract

:
All-inorganic perovskite semiconductors have received significant interest for their potential stability over heat and humidity. However, the typical CsPbI3 displays phase instability despite its desirable bandgap of ~1.73 eV. Herein, we studied the mixed halide perovskite CsPbI2Br by varying the silver doping concentration. For this purpose, we examined its bandgap tunability as a function of the silver doping by using density functional theory. Then, we studied the effect of silver on the structural and optical properties of CsPbI2Br. Resultantly, we found that ‘silver doping’ allowed for partial bandgap tunability from 1.91 eV to 2.05 eV, increasing the photoluminescence (PL) lifetime from 0.990 ns to 1.187 ns, and, finally, contributing to the structural stability when examining the aging effect via X-ray diffraction. Then, through the analysis of the intermolecular interactions based on the solubility parameter, we explain the solvent engineering process in relation to the solvent trapping phenomena in CsPbI2Br thin films. However, silver doping may induce a defect morphology (e.g., a pinhole) during the formation of the thin films.

1. Introduction

Metal halide perovskites (MHPs) are a next-generation semiconductor for electronics and optoelectronics, such as emerging photovoltaics (PVs), light-emitting diodes (LEDs), field-effect transistors (FETs), photodetectors, and sensors [1,2,3,4,5]. MHPs have a chemical formula of ABX3, in which the A-site = organic, inorganic, or hybrid (MA, FA, Cs, or mixture); the B-site = metal (Pb, Sn, or mixture); the X-site = halide (Cl, Br, I, or mixture). Here, MA and FA stand for methylammonium (CH3NH3+) and formamidinium (HC(NH2)2+), respectively [6,7]. In 2009, the Miyasaka group demonstrated for the first time that MAPbI3 and MAPbBr3 can serve as a visible-light sensitizer for PV cells, resulting in the power conversion efficiency (PCE) of 3.81% and 3.13%, respectively [8]. Then, in 2012 and 2015, MHP solar cells reached ~10% and ~20% PCE, respectively, displaying the dramatic advancement in the PV technologies for a short period of time [9,10]. Then, currently, the certified PCE of MHP solar cells was more than ~26%, indicating that perovskite PVs are on the way to commercialization by competing with the traditional silicon PVs [11,12,13].
In the meantime, in 2014, the Snaith group compared three materials, CsPbI3, MAPbI3, and FAPbI3, which have bandgaps (Eg) of 1.73 eV, 1.57 eV, and 1.48 eV, respectively, and focused on the FAPbIyBr3−y (y = 0 to 1) perovskite system, demonstrating the usefulness of the slightly larger FA cation as well as the tunability of the bandgap [6]. Then, in 2015, researchers paid serious attention to the fact that the organic–inorganic MHP solar cells have intrinsic stability problems, such as vulnerability in heat and humidity, originating from the presence of organic moieties such as MA and FA [14,15]. Hence, the Cahen and Hodes group compared the hybrid organic–inorganic MAPbBr3 and all-inorganic CsPbBr3 to demonstrate that the A-sites in the ABX3 structure do not have to be organic cations for high quality PVs, which paved the way for all-inorganic MHP solar cells with enhanced stability [14]. Then, in the same year, the working all-inorganic CsPbI3 solar cells with a PCE of ~1.7–2.9% were demonstrated for the first time by kinetically overcoming the phase instability [15]. This is because, thermodynamically, the most stable phase of the CsPbI3 is the non-perovskite yellow orthorhombic structure at room temperature (RT) [16]. Then, during the past decade, all-inorganic MHP solar cells were further advanced, which includes the perovskite family of CsPbI3, CsPbBr3, CsPbIBr2, and CsPbI2Br [17,18]. Among these, the two extreme materials are CsPbI3 and CsPbBr3, because the former has high visible-light absorption but poor phase stability, whereas the latter has good phase stability but low light absorption. Hence, it is imperative to find an optimum condition between these two extremes through composition engineering, bringing forth the research interests on the CsPbI2Br with the bandgap of ~1.9 eV. Here, CsPbI2Br shares the promising merits of MHPs, such as the high absorption coefficient, tunable bandgap, long exciton diffusion length, low exciton binding energy, and ambipolar transport with a high mobility, which currently results in the state-of-the-art PCE > 17% [19,20,21]. In addition, based on its specific wide bandgap, it can serve as a photoactive component for the top cell in the multi-junctional tandem solar cells [21].
However, CsPbI2Br also has some disadvantageous properties, such as morphological defects from fast crystallization and an un-matched energetic alignment with the charge transport layer [17,18]. Hence, many research groups have an interest in doping engineering, especially for the B-sites in the CsPbI2Br structure [22,23,24,25,26,27,28,29,30]. For example, Zhu et al. reported antimony (Sb3+) doping in the B-site of CsPbI2Br, resulting in the decrease in surface defects, the suppression of the charge recombination, and the improvement in the phase stability [22]. Ma et al. demonstrated the triple improvement in the film quality, radiative recombination, and energy-level alignment through zirconium (Zr4+) doping in the B-site of CsPbI2Br [23]. Han et al. reported the improved passivation, Fermi-level adjustment, and crystallinity by adding calcium (Ca2+) into CsPbI2Br [24]. In the case of Zhang et al., they doped both A- and B-sites for the air-stable Cs1−xRbxPbI2Br with guanidinium (GA+), leading to high crystallinity, an appropriate surface morphology, favorable electronic properties, and a reduced trap density [25]. Guo et al. proved that niobium (Nb5+) doping can increase the perovskite tolerance factor and stability of the α–CsPbI2Br phase [26]. Yang et al. reported the air-stable CsPb1−xGexI2Br (x = 0.1, 0.2, and 0.3) perovskites by germanium (Ge4+) doping, generating a better effective recombination lifetime and low trap densities [27]. Liu et al. doped CsPbI2Br by indium (In3+) and chlorine (Cl), causing inhibition to the yellow photo-inactive phase and simultaneously improving the long-term ambient stability [28]. Duan et al. studied lanthanide (La3+, Ce3+, Nd3+, Sm3+, Eu3+, Gd3+, Tb3+, Ho3+, Er3+, Yb3+, and Lu3+) doping into CsPbI2Br, producing increased grain sizes and carrier lifetimes [29]. Lau et al. carried out the less toxic strontium (Sr2+) doping into CsPbI2Br, creating the passivation effect [30]. All of these clearly demonstrate that CsPbI2Br could be an improved semiconductor by partially substituting the Pb2+ sites with other metal ions for the next-generation optoelectronic devices [17,18,31].
Hence, in this study, in line with the aforementioned works [22,23,24,25,26,27,28,29,30], we studied the silver-doped CsPbI2Br semiconductor for the first time. Here, it is noteworthy that Chen et al. studied the silver-doped CsPbBr3 semiconductor already [32]. However, in our case, we studied the more promising mixed-halide CsPbI2Br semiconductor with silver doping (theoretical) and silver bromide doping (experimental). Note that the bromine ion (DN = 33.7) has a higher Gutmann donor number (DN: Lewis basicity) than the iodine ion (DN = 28.9) [33,34], suggesting that ‘Br’ can replace ‘I’ in CsPbI2Br with a high probability. In this work, we first calculated the silver-doping effect on the electronic structure of CsPbI2Br using the density functional theory (DFT). Then, by adding silver bromide into the perovskite precursor solutions with a PbI2/AgBr = 1:0.01–0.03 molar ratio, we studied the doping effects on the optical, structural, and morphological properties of semiconductor thin films. In addition, we explained the solvent-trapping phenomenon during the solvent engineering process [2,35], because this trapping was detected by the infrared (IR) spectroscopy. Herein, we employed the Hildebrand and Hansen solubility parameter [36,37] to explain the intermolecular interactions between the solvent and antisolvent.

2. Materials and Methods

2.1. Materials

Lead iodide (PbI2, 99.99%, Sigma-Aldrich, Darmstadt, Germany), cesium bromide (CsBr, 99.9%, Sigma-Aldrich, Darmstadt, Germany), silver bromide (AgBr, 99.99%, AR Chemicals, Delhi, India), dimethyl sulfoxide (DMSO; 99.0%, Sigma-Aldrich), and chlorobenzene (CB, ≥99.5%, AR Chemicals, Delhi, India) were purchased and used without further purification.

2.2. Methods

The perovskite precursor solutions were prepared to synthesize the CsPbI2Br perovskite semiconductor without or with AgBr, for which the composition was controlled as follows: (CsBr)1(PbI2)1−x(AgBr)x with x = 0.00, 0.01, 0.02, and 0.03. Here, the control sample was composed of CsBr (1M) and PbI2 (1M) in 1 mL DMSO. Then, after stirring at 500 rpm using a magnetic stirrer for 12 h at 70 °C, the solutions were filtered by a polytetrafluoroethylene (PTFE) syringe filter with a 0.22 μm pore size. Then, 70 μL of solution was dropped on top of the glass substrate (Microscope Slides; Grade-P.2; model number: ISO8037/IS3099 [38]; size: 76 mm × 26 mm × 1.35 mm; Rohem Instruments, Maharashtra, India) and spin-coated at 1500 rpm for 45 s using a spin coater (model: spinNXG-P2; Apex Instruments, Kolkata, India). During spinning, after ~25 s, 200 μL of CB was dispensed on top of the wet perovskite precursor film. After spin-coating, the sample was annealed thermally at 70 °C for 2 min and subsequently at 280 °C for 10 min. Then, all the characterizations were carried out under ambient conditions.

2.3. Thin Film Characterizations

The X-ray diffraction (XRD) patterns were obtained using the Drawell XRD-7000 diffractometer (Shanghai, China) with Cu Kα (3 KW) X-ray radiation (λ = 1.5406 Å), having a source potential of 30 kV and source current of 25 mA (here, with a 2θ range of 10° to 60°; a scan rate of 1° per minute; a step size of 0.01°). Scanning electron microscope (SEM) images were obtained using the benchtop SEM (JCM-6000 Plus, JEOL, Tokyo, Japan) at 5 kV. On the other hand, the high-resolution transmission electron microscopy (HR-TEM) images were investigated by using the model JEM-2100 (Peabody, MA, USA) at 200 kV. The absorption spectra of the film were investigated by ultraviolet–visible (UV–Vis) spectroscopy (PerkinElmer Lambda 25, Kyoto, Japan) in the wavelength range of 350 nm to 750 nm at the scanning speed of 240 nm per minute and the step size of 1 nm. The photoluminescence (PL) lifetime curves were recorded by using the time-correlated single-photon counting (TCSPC) (model: Fluor log 3 TCSPC, Horiba, and Houston, TX, USA) with an excitation wavelength of 570 nm. The Fourier transform infrared (FT-IR) spectroscopy data were obtained using the PerkinElmer spectrum two FT-IR spectrometer (Waltham, MA, USA). Here, the attenuated total reflection (ATR) was employed to record the transmittance data from 4000 to 400 cm−1 with a resolution of 0.5 cm−1. All measurements were taken at room temperature in ambient conditions.

2.4. Computational Method

The electronic band structures of the compounds (CsPbI2Br and Ag-doped CsPbI2Br) were calculated based on the DFT using the Vienna Ab initio Simulation Package (VASP) software (version 4.3.3) in the supercomputing resources provided by the Indian Institute of Science (Bengaluru, India).

3. Results and Discussion

The projector-augmented-wave (PAW) method, which is included in the VASP code, was used for all DFT calculations [39]. The Perdew–Burke–Ernzerhof (PBE) exchange–correlation functional was used for all structural relaxations [40]. After optimizing the CsPbI2Br structure based on the tetragonal β-CsPbI3 crystal (space group P4/mbm), the lattice parameters were determined to be a = b = 6.395 Å and c = 5.988 Å. These values were very close to Chen et al.’s reports of a = b = 6.40 Å and c = 5.97 Å for the pseudo-cubic α-phase CsPbI2Br [41]. Next, the 2 × 2 × 1 supercells (Figure 1) were built using the optimized structure as a basis to study the electronic structures of Cs4Pb4−xAgxI8Br4 (x = 0, 1, 2, 3, and 4) perovskites. Here, when x = 0, the structure is equivalent to four times of the pure CsPbI2Br perovskite, as shown in Figure 1a (supercell), whereas, when x = 1, 2, 3, or 4, the structures become the silver-doped perovskite, as shown in Figure 1b. Here, note that the radius of Br is smaller than that of I, which allows the CsPbI2Br to form a stable crystal structure compared to CsPbI3. In this study, Pb-Br and Pb-I have bond lengths of 2.9940 Å and 3.1975 Å, respectively.
According to Shannon [42], the effective ionic radius at the relevant coordination number (CN) is as follows: Cs+ (CN = XII) (1.88 Å), Pb2+ (VI) (1.19 Å), I (VI) (2.20 Å), and Br (VI) (1.96 Å). Hence, if we calculate Goldschmidt’s tolerance factor, t = R A + R X / 2 R B + R X , t is 0.851 for CsPbI3, 0.855 for CsPbI2Br, and 0.862 for CsPbBr3, respectively. Here, R A , R B , and R X are the radii of A, B, and X, respectively. On the other hand, if we calculate the effective octahedral factor, μ = R B / R X , μ is 0.541 for CsPbI3, 0.561 for CsPbI2Br, and 0.607 for CsPbBr3, respectively. Note that, for CsPbI2Br, we used the average R ¯ I 2 B r = R I × 2 + R B r × 1 / 3 = 2.12 Å for the radius of the mixed halogen. Moreover, to be a cubic phase, the perovskite should have 0.81 < t < 1.11 and 0.44 < μ < 0.90 [17,18].
The ionic radius of cesium is dependent on the CN. For example, if cesium has a CN of VI, VIII, IX, X, XI, and XII, the ionic radius is 1.67 Å, 1.74 Å, 1.78 Å, 1.81 Å, 1.85 Å, and 1.88 Å, respectively. Hence, if some studies in the literature adopt CN = VI instead of XII for the cesium ion, the ionic radius will be 1.67 Å. Therefore, t = 0.807 could be reported for CsPbI3 and t = 0.815 for CsPbBr3, respectively [26]. However, the Cs+ cation has a 12-fold coordination site (the A-site has CN = 12 for all the perovskites), not a 6-fold one [43], indicating that Cs+ (CN = XII) (1.88 Å) should be correct. Importantly, the material should have a tolerance factor of 0.9–1.0 to form an ideal cubic structure [44], indicating the aforementioned t values are somewhat away from this ideal range. In other words, they can easily undergo structural deformation for reducing the Gibbs free energy. Furthermore, thermodynamically, the most stable structure of the CsPbI3 is non-perovskite orthorhombic yellow delta-phase, with Eg = ~2.82 eV at the RT, whereas that of the CsPbBr3 is the orthorhombic gamma-phase, with Eg = ~2.3 eV (here, it is notable that CsPbBr3 has no yellow phase) [45,46,47]. Therefore, it should be reasonable to study CsPbI2Br for improving the structural stability of CsPbI3 via composition engineering (e.g., the B- and X-site modification in the ABX3 structure) for PV applications.
Figure 2 shows (a–e) the electronic structures of the pseudo-cubic α-phase Cs4Pb4−xAgxI8Br4 and (f) the resulting bandgap as a function of the silver doping level. Here, we assumed that the silver atom may stay with CsPbI2Br as a substitutional dopant according to Chen et al.’s study [32]. First, except for Cs4Pb4−xAgxI8Br4 (x = 3), all the others display the promising ‘direct bandgap’ characteristics. Second, although the bandgap of CsPbI2Br is known to be ~1.8–1.9 eV [48,49,50], the DFT results exhibit the small value, 1.361 eV, indicating the typical underestimation of the bandgap in the PBE-based DFT calculation [51,52]. Hence, we need to focus on the trend of the bandgap instead of the exact value itself. Third, when Cs4Pb4−xAgxI8Br4 (x = 2), the bandgap is the largest, 1.966 eV. Fourth, when Cs4Pb4−xAgxI8Br4 (x = 4), i.e., 4[CsAgI2Br], the bandgap is the smallest, 1.060 eV. Finally, this trend of the bandgap is summarized in Figure 2f, providing the insightful silver-doping effect on the electronic structure of CsPbI2Br qualitatively.
In our experimental study, we introduced Ag atoms into the CsPbI2Br crystals by dissolving AgBr into the perovskite precursor solutions, indicating that there should be combined effects from both Ag+ cations and Br anions because our samples are mixed-halide perovskites. Here, in the perovskite precursor solutions, Br anions can serve as a processing additive/dopant because the bromine anions have a high Gutmann’s donor number (DN = 33.7, Lewis basicity) [33,34] affecting the crystallization of the perovskites via the modified interactions between DMSO (DN = 29.8) and the perovskite precursors in the solution state. Figure 3 shows each optical bandgap at the onset of the absorption, for which the Tauc plot was employed for clarity (Figure 4). Here, the film thickness (l) was estimated to be ~148 nm (0% AgBr and 1% AgBr) and ~332 nm (2% AgBr and 3% AgBr), respectively. For this purpose, the following absorption coefficient (α = ~5 × 104 cm−1 at 600 nm [53]) was employed: l = 2.302 × A b s / α , where ‘Abs’ denotes the absorbance. First of all, for the CsPbI2Br perovskite without AgBr, the bandgap is 1.84 eV, which falls in the general bandgap (~1.8–1.9 eV) of the CsPbI2Br perovskite [48,49,50]. Here, it is notable that CsPbI3 and CsPbBr3 have the bandgaps of ~1.73 eV and ~2.3 eV, respectively. However, when AgBr was employed into the perovskite precursor solutions, the resulting bandgap increased slightly from 1.87 eV (at 1% AgBr) to 1.95 eV (at 2% AgBr) and 1.96 eV (at 3% AgBr). Figure 3b shows the summary of the results, i.e., the bandgap as a function of the AgBr doping concentrations. It is worthy to remind that the bandgap is a key factor in the ‘stability and cost’ for practical solar cells, determining the theoretical PCE based on the Shockley–Queisser limit [54]. Importantly, Ravi et al. pointed out that, in CsPbX3 perovskite, the conduction band minimum (CBM) is dominantly affected by Pb 6p orbitals, whereas the valence band maximum (VBM) is mainly determined by anti-bonding hybridization Pb 6s and X np orbitals, specifically, the major effect from X np [55]. Therefore, the bandgap shift in Figure 4 could be explained as follows. Ag-doping may affect the CBM shift, whereas Br codoping may contribute to the VBM move. However, these effects will be very small because of the slight AgBr doping concentration of ~1–3%. Then, we measured the PL lifetime for the CsPbI2Br when AgBr was 0, 1, 2, and 3%. Accordingly, as shown in Figure 5, the PL lifetime was enhanced from 0.990 ns (at 0% AgBr) to 1.187 ns (at 3% AgBr) with increasing AgBr amounts, suggesting that the AgBr doping helps minimize the nonradiative transition according to the literature reports [22,23,24,25,26,27,28,29,30,32].
The FT-IR spectra were characterized for the CsPbI2Br thin film as a function of the AgBr doping in the perovskite precursor solutions. As shown in Figure 6a, we can find only the FT-IR peaks from the solvents remaining inside of the CsPbI2Br thin film. This observation indicates that the trace amounts of the solvent molecules may survive in the trapped state inside of the perovskite film, although the annealing temperature (>250 °C) was higher than the boiling points of each solvent molecules (DMSO: 189 °C and CB: 132 °C; their structures are given in Figure 6b). Here, the detailed peak assignment is as follows [56]: First, in the high frequency regions, 3025–2849 cm−1, =C-H and –C-H vibrations were observed from the DMSO and CB molecules inside of the CsPbI2Br crystals. At 1676 cm−1, a -C=C vibration from the aromatic ring of CB was detected. On the other hand, at 1387 cm−1, 1092 cm−1, and 738 cm−1, -CH3, S=O, and –C-Cl vibrations were displayed, respectively. Finally, the small peak at 463 cm−1 is ascribed to the molecular vibration of the antisolvent CB.
Importantly, Figure 6c shows the solvent engineering process [35] to explain the solvent trapping phenomena detected via the FT-IR. Here, the solvent engineering procedure is as follows. The antisolvent (chlorobenzene) dripping on top of the wet perovskite precursor film (solvent: DMSO) during spinning brings forth the fast crystallization and deposition of a perovskite film. Here, to understand the intermolecular interactions, the solubility parameter (δ) data [36] of the solvent, antisolvent, and perovskite are required. First, DMSO and CB have δ = 14.5 (cal/cm3)2 and δ = 9.5 (cal/cm3)2, respectively [57]. In the case of CsPbI2Br, we may estimate it from the water contact angle ( θ c = 32.76°) data reported by Chen and coworkers [58]. Li and Neumann [59] suggested the relation between the contact angle and surface energy, cos θ c = 1 + 2 γ s v / γ l v   exp β ¯ γ l v γ s v 2 , where γ l v , γ s v , and γ s l are the surface energies for liquid–vapor, solid–vapor, and solid–liquid, respectively. The constant β ¯ is 0.000115 m4/mJ2 and γ l v is 72.8 mJ/m2 for water, respectively. Then, by inputting θ c = 32.76° into the aforementioned Li–Neumann’s equation, we may estimate γ s v = 63.05 mJ/m2. Then, from the relation of δ cal / cm 3 2 = 1.829058 γ s v [60,61,62], we obtained δ = 14.5 (cal/cm3)2 or δ SI   unit = δ × 2.0455 = 29.7 MPa1/2, respectively (Table 1). Hence, because CsPbI2Br and DMSO have (almost) the same solubility parameter, there is high probability that DMSO may be trapped in the CsPbI2Br. However, for the case of CB, the ‘solvent–antisolvent’ (DMSO and CB) molecules are miscible because of the entropy-driven mixing, affording CB to wash and remove the DMSO molecules during its dripping process [35]. However, when CB was dropped on top of the wet perovskite precursor film, CB could also be trapped into the crystal structure of CsPbI2Br, although CB and CsPbI2Br have two different polarities, i.e., CB is slightly polar (polarity index = 2.7) but CsPbI2Br and DMSO (polarity index = 7.2) are highly polar [63]. Hence, CB’s trapping could be understood based on the physical trap instead of the chemical affinity between the CB and the CsPbI2Br perovskite. On the other hand, DMSO can be trapped for two reasons, i.e., affinity and physical confinement. Accordingly, even after thermal annealing at 280° for 10 min, trace amounts of the solvents could be trapped, as demonstrated in the FT-IR spectra in Figure 6a.
Figure 7a shows the XRD patterns of CsPbI2Br as a function of the AgBr doping concentration. First, in the absence of AgBr, the CsPbI2Br perovskite thin film displays the typical (100) and (200) peaks [64]. However, by increasing the AgBr concentration (see ~2–3% AgBr), the other peaks such as (211), (300), and (222) are intensified, indicating that the crystallographic ordering decreases with the increasing AgBr concentrations. This observation implies that the crystallization kinetics were changed when the AgBr was introduced into the perovskite precursor solutions. Second, we estimated the crystallite size (D) by using Scherrer’s relation of D = 0.9 λ / B cos θ , where λ (=0.154 nm) is the wavelength of the X-ray, whereas B is the full width at half-maximum (FWHM) at the diffraction angle of θ. The results are summarized in Figure 7b and Table 2. As shown in Figure 7b, there was no clear linear trend when AgBr was introduced into the perovskite precursor, implying that, although the morphology might be changed through different crystallization kinetics, the crystallite size (i.e., the average single crystalline domains in the polycrystalline structure) were not much changed, but rather similar to all the conditions, whether doped or not.
Figure 8a,b show the stability test of the perovskite thin film: (a) CsPbI2Br without AgBr and (b) CsPbI2Br with 1% AgBr doping. These two samples were selected for this test because the surface morphologies were relatively uniform compared to the others (~2–3% AgBr-doped perovskite samples). Interestingly, both samples show the growth of the minor peaks at the (211) and (222) crystallographic planes with time, indicating that the orientational ordering decreases with time. Here, it is notable that, except for single crystalline perovskite thin films, all the polycrystalline films are thermodynamically metastable because the defect area (including polycrystalline nature) makes the surface energy increase. Hence, for the lowering of the Gibbs free energy, the sample can undergo phase transition. In this case, by decreasing the orientational order (i.e., increasing the (211)-(222) XRD peaks), the film may reduce its free energy. When we see the CsPbI2Br sample without the AgBr doping in Figure 8a, compared to the same sample (but different batch) in Figure 7 (black solid line), the additional strong peak at the (300) crystallographic plane was observed [i.e., a more orientational order because the (100), (200), and (300) planes are equivalent], indicating the batch-to-batch partial uncertainty depending on the drying process in the laboratory under ambient conditions. Importantly, the 1% AgBr-doped CsPbI2Br shows the structural stability (i.e., the XRD peak position is the same with aging time), but the 0% AgBr sample clearly shows the major peak’s shift to the left direction (i.e., a partial expansion of crystal; see the dotted red line box in Figure 8a). This aging effect data proves that the AgBr doping should contribute to the structural stability of the CsPbI2Br perovskite films. The results are reasonable because the AgBr addition increases the stability (a wider bandgap and improved tolerance to the environments).
Figure 9 shows the SEM images displaying the microstructural morphologies of the CsPbI2Br sample as a function of the AgBr doping concentration. First, the CsPbI2Br thin films (a) without the AgBr and (b) with the 1% AgBr are relatively uniform, whereas the other films with the ~2–3% AgBr are nonuniform, displaying the crystal domains and defect sites clearly. Probably, the samples (c and d) were grown very fast in the presence of high doping (~2–3% AgBr). However, it is worth reminding that, according to the XRD data in Figure 7, the crystallite size (the average single-crystalline domains) is not much different from sample-to-sample. The average crystallite size is 41.5 ± 3.4 nm and 39.3 ± 7.2 nm at the (100) and (200) crystallographic planes, respectively. However, as shown in Figure 9, the film processing condition should be optimized further for photonic devices, which will be included in our future work. Finally, we note that, without annealing (>250 °C), the CsPbI2Br sample shows phase impurities at room temperature due to polymorphism (see Table S1 and Figure S1 in Supplementary Materials).

4. Conclusions

CsPbI2Br is an interesting light-harvesting next-generation semiconductor by positioning between CsPbI3 (with a low stability but high absorption) and CsPbBr3 (with a high stability but low absorption), displaying the bandgap of ~1.84–1.96 eV and improved stability compared to CsPbI3. First, we demonstrated the tunability of the CsPbI2Br bandgap by substituting Pb with Ag through the DTF calculation. Herein, the bandgap reached the maximum when Pb/Ag = 50:50% (i.e., x = 2 in Cs4Pb4−xAgxI8Br4). Second, when we added AgBr into the perovskite precursor solution from 1% to 3%, the bandgap increased from ~1.87 eV to ~1.96 eV, which is qualitatively in line with the DFT prediction. Third, the PL lifetime was enhanced by employing AgBr into the perovskite precursor solution. Fourth, the FT-IR spectra exhibited that the solvent/antisolvent molecules were trapped inside of the perovskite thin films. Fifth, when AgBr was introduced (or the aging time was increased), the minor XRD peaks at the (211) and (222) crystallographic planes were partially enhanced, indicating the diminished orientational ordering in the film. Sixth, when AgBr was incorporated into the perovskite film (with a thickness of ~148–332 nm), the crystallization kinetics were changed, affecting the morphologies of the films. Seventh, based on the XRD data, AgBr doping was observed to contribute to the enhanced stability of the CsPbI2Br perovskite structure. Finally, it should be valuable to study further all-inorganic CsPbI2Br materials for the development of perovskite solar cells with enhanced stability as our future works.

Supplementary Materials

The following supporting information [16,43] can be downloaded at: https://www.mdpi.com/article/10.3390/electronicmat5020005/s1, Table S1: Analysis of the selected area electron diffraction (SAED) image for CsPbI2Br without annealing at room temperature. Figure S1: HR-TEM image of CsPbI2Br with different scale bars: (a) 2 nm, (b) 5 nm, and (c) 10 nm. (d) SAED pattern of CsPbI2Br displaying the phase impurity at room temperature.

Author Contributions

Writing—original draft preparation, T.K.; writing—review and editing, J.Y.K.; conceptualization, T.K., M.A., D.M. and J.Y.K.; methodology, T.K. and J.Y.K.; software, T.K. and M.A.; formal analysis and investigation, T.K., M.A., D.M., A.T., S.T. and J.Y.K.; resources, A.T., S.T. and J.Y.K.; data curation, T.K.; 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. The financial supports from the Jimma Institute of Technology and Bonga University in Ethiopia are appreciated by T.K.

Data Availability Statement

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

Acknowledgments

The Indian Institute of Science (S. Karthikeyan) is acknowledged for providing the supercomputing resources.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Gidey, A.T.; Assayehegn, E.; Kim, J.Y. Hydrophilic Surface-Driven Crystalline Grain Growth of Perovskites on Metal Oxides. ACS Appl. Energy Mater. 2021, 4, 6923–6932. [Google Scholar] [CrossRef]
  2. Kim, J.Y.; Yoo, Y.; Kim, J.; Park, H.J.; Cho, W.; Lee, S.; Sung, Y.-E.; Bae, B. Phase Behavior of Quasi-2D Hybrid Lead Bromide Perovskite Precursor Solutions. ACS Appl. Opt. Mater. 2024, 2, 108–117. [Google Scholar] [CrossRef]
  3. Zhou, S.; Zhou, G.; Li, Y.; Xu, X.; Hsu, Y.J.; Xu, J.; Zhao, N.; Lu, X. Understanding Charge Transport in All-Inorganic Halide Perovskite Nanocrystal Thin-Film Field Effect Transistors. ACS Energy Lett. 2020, 5, 2614–2623. [Google Scholar] [CrossRef]
  4. Ahmadi, M.; Wu, T.; Hu, B. A Review on Organic–Inorganic Halide Perovskite Photodetectors: Device Engineering and Fundamental Physics. Adv. Mater. 2017, 29, 1605242. [Google Scholar] [CrossRef]
  5. Ai, B.; Fan, Z.; Wong, Z.J. Plasmonic–perovskite solar cells, light emitters, and sensors. Microsyst. Nanoeng. 2022, 8, 5. [Google Scholar] [CrossRef] [PubMed]
  6. Eperon, G.E.; Stranks, S.D.; Menelaou, C.; Johnston, M.B.; Herza, L.M.; Snaith, H.J. Formamidinium lead trihalide: A broadly tunable perovskite for efficient planar heterojunction solar cells. Energy Environ. Sci. 2014, 7, 982–988. [Google Scholar] [CrossRef]
  7. Gidey, A.T.; Kim, J.Y. Tuning the crystallization process of perovskite active layer using a functionalized graphene oxide for enhanced photovoltaic performance. J. Mater. Sci. Mater. Electron. 2020, 31, 12257–12268. [Google Scholar] [CrossRef]
  8. 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]
  9. 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]
  10. Chen, W.; Wu, Y.; Yue, Y.; Liu, J.; Zhng, W.; Yang, X.; Chen, H.; Bi, E.; Ashraful, I.; Grätzel, M.; et al. Efficient and stable large-area perovskite solar cells with inorganic charge extraction layers. Science 2015, 350, 944–948. [Google Scholar] [CrossRef]
  11. 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]
  12. Tan, Q.; Li, Z.; Luo, G.; Zhang, X.; Che, B.; Chen, G.; Gao, H.; He, D.; Ma, G.; Wang, J.; et al. Inverted perovskite solar cells using dimethylacridine-based dopants. Nature 2023, 620, 545–551. [Google Scholar] [CrossRef] [PubMed]
  13. Duan, L.; Walter, D.; Chang, N.; Bullock, J.; Kang, D.; Phang, S.P.; Weber, K.; White, T.; Macdonald, D.; Catchpole, K.; et al. Stability challenges for the commercialization of perovskite–silicon tandem solar cells. Nat. Rev. Mater. 2023, 8, 261–281. [Google Scholar] [CrossRef]
  14. Kulbak, M.; Cahen, D.; Hodes, G. How Important Is the Organic Part of Lead Halide Perovskite Photovoltaic Cells? Efficient CsPbBr3 Cells. J. Phys. Chem. Lett. 2015, 6, 2452–2456. [Google Scholar] [CrossRef] [PubMed]
  15. Eperon, G.E.; 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]
  16. 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]
  17. Liu, X.; Li, J.; Cui, X.; Wang, X.; Yang, D. The progress and efficiency of CsPbI2Br perovskite solar cells. J. Mater. Chem. C 2023, 11, 426–455. [Google Scholar] [CrossRef]
  18. Lim, E.L.; Yang, J.; Wei, Z. Inorganic CsPbI2Br halide perovskites: From fundamentals to solar cell optimizations. Energy Environ. Sci. 2023, 16, 862–888. [Google Scholar] [CrossRef]
  19. Huang, J.; Zhou, D.; Yan, H.; Meng, C.; Yang, Y.; Liu, J.; Wang, M.; Xu, P.; Peng, Z.; Chen, J.; et al. A multiple-coordination framework for CsPbI2Br perovskite solar cells. J. Mater. Chem. C 2024, 12, 4112–4122. [Google Scholar] [CrossRef]
  20. Jeong, M.J.; Jeon, S.W.; Kim, S.Y.; Noh, J.H. High Fill Factor CsPbI2Br Perovskite Solar Cells Via Crystallization Management. Adv. Energy Mater. 2023, 13, 2300698. [Google Scholar] [CrossRef]
  21. Ding, Y.; Guo, Q.; Geng, Y.; Dai, Z.; Wang, Z.; Chen, Z.; Guo, Q.; Zheng, Z.; Li, Y.; Zhou, E. A low-cost hole transport layer enables CsPbI2Br single-junction and tandem perovskite solar cells with record efficiencies of 17.8% and 21.4%. Nano Today 2022, 46, 101586. [Google Scholar] [CrossRef]
  22. Zhu, M.; Qin, L.; Xia, Y.; Liang, J.; Wang, Y.; Hong, D.; Tian, Y.; Tie, Z.; Jin, Z. Antimony doped CsPbI2Br for high-stability all-inorganic perovskite solar cells. Nano Res. 2024, 17, 1508–1515. [Google Scholar] [CrossRef]
  23. Ma, P.; Bie, T.; Liu, Y.; Yang, L.; Bi, S.; Wang, Z.; Shao, M. Zirconium Doping to Enable High-Efficiency and Stable CsPbI2Br All-Inorganic Perovskite Solar Cells. ACS Appl. Mater. Interfaces 2024, 16, 1217–1224. [Google Scholar] [CrossRef] [PubMed]
  24. Han, Y.; Zhao, H.; Duan, C.; Yang, S.; Yang, Z.; Liu, Z.; Liu, S.F. Controlled n-Doping in Air-Stable CsPbI2Br Perovskite Solar Cells with a Record Efficiency of 16.79%. Adv. Funct. Mater. 2020, 30, 1909972. [Google Scholar] [CrossRef]
  25. Zhang, W.; Xiong, J.; Li, J.; Daoud, W.A. Guanidinium Passivation for Air-Stable Rubidium- Incorporated Cs(1-x)RbxPbI2Br Inorganic Perovskite Solar Cells. Sol. RRL 2020, 4, 2000112. [Google Scholar] [CrossRef]
  26. Guo, Z.; Zhao, S.; Liu, A.; Kamata, Y.; Teo, S.; Yang, S.; Xu, Z.; Hayase, S.; Ma, T. Niobium Incorporation into CsPbI2Br for Stable and Efficient All-Inorganic Perovskite Solar Cells. ACS Appl. Mater. Interfaces 2019, 11, 19994–20003. [Google Scholar] [CrossRef] [PubMed]
  27. Yang, F.; Hirotani, D.; Kapil, G.; Kamarudin, M.A.; Ng, C.H.; Zhang, Y.; Shen, Q.; Hayase, S. All-Inorganic CsPb1−xGexI2Br Perovskite with Enhanced Phase Stability and Photovoltaic Performance. Angew. Chem. Int. Ed. 2018, 57, 12745–12749. [Google Scholar] [CrossRef]
  28. Liu, C.; Li, W.; Li, H.; Wang, H.; Zhang, C.; Yang, Y.; Gao, X.; Xue, Q.; Yip, H.-Y.; Fan, J.; et al. Structurally Reconstructed CsPbI2Br Perovskite for Highly Stable and Square-Centimeter All-Inorganic Perovskite Solar Cells. Adv. Energy Mater. 2019, 9, 1803572. [Google Scholar] [CrossRef]
  29. Duan, J.; Zhao, Y.; Yang, X.; Wang, Y.; He, B.; Tang, Q. Lanthanide Ions Doped CsPbBr3 Halides for HTM-Free 10.14%-Efficiency Inorganic Perovskite Solar Cell with an Ultrahigh Open-Circuit Voltage of 1.594 V. Adv. Energy Mater. 2018, 8, 1802346. [Google Scholar] [CrossRef]
  30. Lau, C.F.J.; Zhang, M.; Deng, X.; Zheng, J.; Bing, J.; Ma, Q.; Kim, J.; Hu, L.; Green, M.A.; Huang, S.; et al. Strontium-Doped Low-Temperature-Processed CsPbI2Br Perovskite Solar Cells. ACS Energy Lett. 2017, 2, 2319–2325. [Google Scholar] [CrossRef]
  31. Swarnkar, A.; Mir, W.J.; Nag, A. Can B-Site Doping or Alloying Improve Thermal- and Phase-Stability of All-Inorganic CsPbX3 (X = Cl, Br, I) Perovskites? ACS Energy Lett. 2018, 3, 286–289. [Google Scholar] [CrossRef]
  32. Chen, S.; Liu, X.; Wang, Z.; Li, W.; Gu, X.; Lin, J.; Yang, T.; Gao, X.; Kyaw, A.K.K. Defect Passivation of CsPbBr3 with AgBr for High-Performance All-Inorganic Perovskite Solar Cells. Adv. Energy Sustain. Res. 2021, 2, 2000099. [Google Scholar] [CrossRef]
  33. Gutmann, V. Solvent effects on the reactivities of organometallic compounds. Coord. Chem. Rev. 1976, 18, 225–255. [Google Scholar] [CrossRef]
  34. 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] [PubMed]
  35. 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. Nature Mater. 2014, 13, 897–903. [Google Scholar] [CrossRef] [PubMed]
  36. Hansen, C.M. Hansen Solubility Parameters—A User’s Handbook; CRC Press: Boca Raton, FL, USA, 2000. [Google Scholar]
  37. Hansen, C.M. 50 Years with solubility parameters—Past and future. Prog. Org. Coat. 2004, 51, 77–84. [Google Scholar] [CrossRef]
  38. ISO8037/IS3099; Educational Instruments and Equipment. The Bureau of Indian Standards: New Delhi, India, 1992.
  39. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169–11186. [Google Scholar] [CrossRef]
  40. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1997, 77, 3865–3868. [Google Scholar] [CrossRef] [PubMed]
  41. Chen, Y.; Shi, T.; Liu, P.; Xie, W.; Chen, K.; Xu, X.; Shui, L.; Shang, C.; Chen, Z.; Yip, H.-L.; et al. The distinctive phase stability and defect physics in CsPbI2Br perovskite. J. Mater. Chem. A 2019, 7, 20201–20207. [Google Scholar] [CrossRef]
  42. Shannon, R.D. Revised Effective Ionic Radii and Systematic Studies of Interatomic Distances in Halides and Chalcogenides. Acta Cryst. 1976, A32, 751–767. [Google Scholar] [CrossRef]
  43. Wang, B.; Novendra, N.; Navrotsky, A. Energetics, Structures, and Phase Transitions of Cubic and Orthorhombic Cesium Lead Iodide (CsPbI3) Polymorphs. J. Am. Chem. Soc. 2019, 141, 14501–14504. [Google Scholar] [CrossRef]
  44. Li, Z.; Yang, M.; Park, J.-S.; Wei, S.-W.; 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]
  45. Kebede, T.; Abebe, M.; Mani, D.; Paduvilan, J.K.; Thottathi, L.; Thankappan, A.; Thomas, S.; Kamangar, S.; Shaik, A.S.; Badruddin, I.A.; et al. Phase Behavior and Role of Organic Additives for Self-Doped CsPbI3 Perovskite Semiconductor Thin Films. Micromachines 2023, 14, 1601. [Google Scholar] [CrossRef]
  46. 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. [Google Scholar] [CrossRef]
  47. Idosa, D.A.; Abebe, M.; Mani, D.; Thankappan, A.; Thomas, S.; Aga, F.G.; Kim, J.Y. A Blue-Light-Emitting 3 nm-Sized CsPbBr3 Perovskite Quantum Dot with ZnBr2 Synthesized by Room-Temperature Supersaturated Recrystallization. Photonics 2023, 10, 802. [Google Scholar] [CrossRef]
  48. Xu, W.; He, F.; Zhang, M.; Nie, P.; Zhang, S.; Zhao, C.; Luo, R.; Li, J.; Zhang, X.; Zhao, S.; et al. Minimizing Voltage Loss in Efficient All-Inorganic CsPbI2Br Perovskite Solar Cells through Energy Level Alignment. ACS Energy Lett. 2019, 4, 2491–2499. [Google Scholar] [CrossRef]
  49. Lu, J.; Chen, S.C.; Zheng, Q. Defect passivation of CsPbI2Br perovskites through Zn(II) doping: Toward efficient and stable solar cells. Sci. China Chem. 2019, 62, 1044–1050. [Google Scholar] [CrossRef]
  50. Zeng, Z.; Zhang, J.; Gan, X.; Sun, H.; Shang, M.; Hou, D.; Lu, C.; Chen, R.; Zhu, Y.; Han, L. In Situ Grain Boundary Functionalization for Stable and Efficient Inorganic CsPbI2Br Perovskite Solar Cells. Adv. Energy Mater. 2018, 8, 1801050. [Google Scholar] [CrossRef]
  51. Yan, D.; Shi, T.; Zang, Z.; Zhou, T.; Liu, Z.; Zhang, Z.; Du, J.; Leng, Y.; Tang, X. Ultrastable CsPbBr3 Perovskite Quantum Dot and Their Enhanced Amplified Spontaneous Emission by Surface Ligand Modification. Small 2019, 15, 1901173. [Google Scholar] [CrossRef]
  52. 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]
  53. Mariotti, S.; Hutter, O.S.; Phillips, L.J.; Yates, P.J.; Kundu, B.; Durose, K. Stability and Performance of CsPbI2Br Thin Films and Solar Cell Devices. ACS Appl. Mater. Interfaces 2018, 10, 3750–3760. [Google Scholar] [CrossRef]
  54. Shockley, W.; Queisser, H.J. Detailed Balance Limit of Efficiency of p-n Junction Solar Cells. J. Appl. Phys. 1961, 32, 510–519. [Google Scholar] [CrossRef]
  55. Ravi, V.K.; Markad, G.B.; Nag, A. Band Edge Energies and Excitonic Transition Probabilities of Colloidal CsPbX3 (X = Cl, Br, I) Perovskite Nanocrystals. ACS Energy Lett. 2016, 1, 665–671. [Google Scholar] [CrossRef]
  56. Schrader, B. Infrared and Raman Spectroscopy: Methods and Applications; VCH: Weinheim, Germany, 1995. [Google Scholar]
  57. Awol, N.; Amente, C.; Verma, G.; Kim, J.Y. Morphology and surface analyses for CH3NH3PbI3 perovskite thin films treated with versatile solvent–antisolvent vapors. RSC Adv. 2021, 11, 17789. [Google Scholar] [CrossRef]
  58. Chen, Q.; Lin, L.; Wang, Y.; Gao, Z.; Fu, Y.; Liu, Q.; Li, J.; He, D. Enhancement of photoelectric performance for CsPbI2Br solar cells by the synergistic effect of binary additives. J. Materiomics 2023, 9, 27–34. [Google Scholar] [CrossRef]
  59. Li, D.; Neumann, A.W. A Reformulation of the Equation of State for Interfacial Tensions. J. Colloid Interface Sci. 1990, 137, 304–307. [Google Scholar] [CrossRef]
  60. Nilsson, S.; Bernasik, A.; Budkowski, A.; Moons, E. Morphology and Phase Separation of Spin-Coated Films of Polyfluorene/PCBM Blends. Macromolecules 2007, 40, 8291–8301. [Google Scholar] [CrossRef]
  61. Kim, J.Y. Order–Disorder Phase Equilibria of Regioregular Poly(3-hexylthiophene-2,5-diyl) Solution. Macromolecules 2018, 51, 9026–9034. [Google Scholar] [CrossRef]
  62. Kim, J.Y. Phase Diagrams of Binary Low Bandgap Conjugated Polymer Solutions and Blends. Macromolecules 2019, 52, 4317–4328. [Google Scholar] [CrossRef]
  63. Kleiman, M.; Ryu, K.A.; Esser-Kahn, A.P. Determination of Factors Influencing the Wet Etching of Polydimethylsiloxane Using Tetra-n-butylammonium Fluoride. Macromol. Chem. Phys. 2016, 217, 284–291. [Google Scholar] [CrossRef]
  64. Atourki, L.; Bernabe, M.; Makha, M.; Bouabid, K.; Regragui, M.; Ihlal, A.; Abd-lefdil, M.; Mollar, M. Effect of doping on the phase stability and photophysical properties of CsPbI2Br perovskite thin films. RSC Adv. 2021, 11, 1440–1449. [Google Scholar] [CrossRef] [PubMed]
  65. Jeong, B.; Han, H.; Choi, Y.J.; Cho, S.H.; Kim, E.H.; Lee, S.W.; Kim, J.S.; Park, C.; Kim, D.; Park, C. All-Inorganic CsPbI3 Perovskite Phase-Stabilized by Poly(ethylene oxide) for Red-Light-Emitting Diodes. Adv. Funct. Mater. 2018, 28, 1706401. [Google Scholar] [CrossRef]
  66. Ma, S.; Kim, S.H.; Jeong, B.; Kwon, H.-C.; Yun, S.-C.; Jang, G.; Yang, H.; Park, C.; Lee, D.; Moon, J. Strain-Mediated Phase Stabilization: A New Strategy for Ultrastable α-CsPbI3 Perovskite by Nanoconfined Growth. Small 2019, 15, 1900219. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) Unit cell and supercell of CsPbI2Br with lattice parameters, a = b = 6.395 Å and c = 5.988 Å. (b) B-site substitutional doping by silver for the CsPbI2Br supercell with a size of 2 × 2 × 1.
Figure 1. (a) Unit cell and supercell of CsPbI2Br with lattice parameters, a = b = 6.395 Å and c = 5.988 Å. (b) B-site substitutional doping by silver for the CsPbI2Br supercell with a size of 2 × 2 × 1.
Electronicmat 05 00005 g001
Figure 2. Band structure of the pseudo-cubic α-phase Cs4Pb4−xAgxI8Br4 when (a) x = 0, (b) x = 1, (c) x = 2, (d) x = 3, and (e) x = 4. (f) Bandgap as a function of x in Cs4Pb4−xAgxI8Br4.
Figure 2. Band structure of the pseudo-cubic α-phase Cs4Pb4−xAgxI8Br4 when (a) x = 0, (b) x = 1, (c) x = 2, (d) x = 3, and (e) x = 4. (f) Bandgap as a function of x in Cs4Pb4−xAgxI8Br4.
Electronicmat 05 00005 g002
Figure 3. (a) UV–Vis spectra of CsPbI2Br as a function of the AgBr concentration. (b) Bandgap as a function of the AgBr doping concentration (%). Optical bandgap determination for CsPbI2Br as a function of the AgBr concentration by the Tauc plot.
Figure 3. (a) UV–Vis spectra of CsPbI2Br as a function of the AgBr concentration. (b) Bandgap as a function of the AgBr doping concentration (%). Optical bandgap determination for CsPbI2Br as a function of the AgBr concentration by the Tauc plot.
Electronicmat 05 00005 g003
Figure 4. Optical bandgap determination for CsPbI2Br as a function of the AgBr concentration (a) 0%, (b) 1%, (c) 2% and (d) 3% by the Tauc plot. Here, each arrow indicates the tangent line for determining the optical bandgap.
Figure 4. Optical bandgap determination for CsPbI2Br as a function of the AgBr concentration (a) 0%, (b) 1%, (c) 2% and (d) 3% by the Tauc plot. Here, each arrow indicates the tangent line for determining the optical bandgap.
Electronicmat 05 00005 g004
Figure 5. PL lifetime of the CsPbI2Br thin film as a function of the AgBr concentration.
Figure 5. PL lifetime of the CsPbI2Br thin film as a function of the AgBr concentration.
Electronicmat 05 00005 g005
Figure 6. (a) FT-IR spectra of CsPbI2Br with or without AgBr when processed with dimethyl sulfoxide (DMSO) and chlorobenzene (CB). (b) Chemical structures of DMSO and CB. (c) Solvent engineering process: when CB is dripping on top of the wet perovskite (precursor) film, DMSO can be washed away. During this process, some solvent molecules could be trapped in the perovskite thin film.
Figure 6. (a) FT-IR spectra of CsPbI2Br with or without AgBr when processed with dimethyl sulfoxide (DMSO) and chlorobenzene (CB). (b) Chemical structures of DMSO and CB. (c) Solvent engineering process: when CB is dripping on top of the wet perovskite (precursor) film, DMSO can be washed away. During this process, some solvent molecules could be trapped in the perovskite thin film.
Electronicmat 05 00005 g006
Figure 7. CsPbI2Br as a function of AgBr doping: (a) XRD patterns and (b) crystallite size. The PDF reference card number is 80-4039 [65,66].
Figure 7. CsPbI2Br as a function of AgBr doping: (a) XRD patterns and (b) crystallite size. The PDF reference card number is 80-4039 [65,66].
Electronicmat 05 00005 g007
Figure 8. Structural stability test by the XRD: (a) CsPbI2Br without AgBr and (b) CsPbI2Br with 1% AgBr.
Figure 8. Structural stability test by the XRD: (a) CsPbI2Br without AgBr and (b) CsPbI2Br with 1% AgBr.
Electronicmat 05 00005 g008
Figure 9. SEM images: (a) CsPbI2Br without AgBr, (b) CsPbI2Br with the 1% AgBr (c) CsPbI2Br with the 2% AgBr, and (d) CsPbI2Br with the 3% AgBr.
Figure 9. SEM images: (a) CsPbI2Br without AgBr, (b) CsPbI2Br with the 1% AgBr (c) CsPbI2Br with the 2% AgBr, and (d) CsPbI2Br with the 3% AgBr.
Electronicmat 05 00005 g009
Table 1. Solubility parameter, molecular weight, density, and molar volume of solvent, non-solvent, and CsPbI2Br. δ (SI unit) = δ × 2.0455.
Table 1. Solubility parameter, molecular weight, density, and molar volume of solvent, non-solvent, and CsPbI2Br. δ (SI unit) = δ × 2.0455.
Chemical δ (MPa1/2) δ [(cal/cm3)2]MW (g/mol) ρ (g/cm3) V ^ i (cm3/mol)
DMSO29.714.578.131.1071.03
CB19.49.5112.561.11101.41
CsPbI2Br29.714.5673.824.79 a140.67
a This is estimated based on the lattice parameters, a = b = 0.640 nm and c = 0.597 nm.
Table 2. Crystallite size in CsPbI2Br as a function of the AgBr doping concentration.
Table 2. Crystallite size in CsPbI2Br as a function of the AgBr doping concentration.
PlaneAgBr (%)2θ (°)B (rad)D (nm)
(100)0150.00303646.1
1150.00360138.8
2150.00333441.9
3150.00357339.1
(200)0300.00364045.4
1300.00431933.2
2300.00314045.7
3300.00434733.0
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

Kebede, T.; Abebe, M.; Mani, D.; Thankappan, A.; Thomas, S.; Kim, J.Y. Silver-Doped CsPbI2Br Perovskite Semiconductor Thin Films. Electron. Mater. 2024, 5, 56-70. https://doi.org/10.3390/electronicmat5020005

AMA Style

Kebede T, Abebe M, Mani D, Thankappan A, Thomas S, Kim JY. Silver-Doped CsPbI2Br Perovskite Semiconductor Thin Films. Electronic Materials. 2024; 5(2):56-70. https://doi.org/10.3390/electronicmat5020005

Chicago/Turabian Style

Kebede, Tamiru, Mulualem Abebe, Dhakshnamoorthy Mani, Aparna Thankappan, Sabu Thomas, and Jung Yong Kim. 2024. "Silver-Doped CsPbI2Br Perovskite Semiconductor Thin Films" Electronic Materials 5, no. 2: 56-70. https://doi.org/10.3390/electronicmat5020005

Article Metrics

Back to TopTop