Next Article in Journal
Computational Approaches to the Electronic Properties of Noble Metal Nanoclusters Protected by Organic Ligands
Next Article in Special Issue
Performance Enhancement of All-Inorganic Carbon-Based CsPbIBr2 Perovskite Solar Cells Using a Moth-Eye Anti-Reflector
Previous Article in Journal
BiVO4 Ceramic Photoanode with Enhanced Photoelectrochemical Stability
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Controlled Growth of Porous InBr3: PbBr2 Film for Preparation of CsPbBr3 in Carbon-Based Planar Perovskite Solar Cells

1
School of Science, Northeast Electric Power University, Jilin 132012, China
2
State Key Laboratory of Superhard Materials, College of Physics, Jilin University, Changchun 130012, China
3
Beijing Key Lab of Cryo-Biomedical Engineering and Key Lab of Cryogenics, Technical Institute of Physics and Chemistry, Chinese Academy of Sciences, Beijing 100190, China
*
Authors to whom correspondence should be addressed.
Submission received: 15 August 2021 / Revised: 10 September 2021 / Accepted: 13 September 2021 / Published: 16 September 2021
(This article belongs to the Special Issue Advances in Nanomaterials for Perovskite Solar Cells)

Abstract

:
Due to the low solubility of CsBr in organic solvents, the CsPbBr3 film prepared by the multi-step method has holes and insufficient thickness, and the light absorption capacity and current density of the perovskite film hinder the further improvement in the power conversion efficiency (PCE) of CsPbBr3 solar cells. In this study, we introduced InBr3 into the PbBr2 precursor solution and adjusted the concentration of PbBr2, successfully prepared PbBr2 with a porous structure on the compact TiO2 (c-TiO2) substrate to ensure that it fully reacted with CsBr, and obtained the planar carbon-based CsPbBr3 solar cells with high-quality perovskite film. The results reveal that the porous PbBr2 structure and the increasing PbBr2 concentration are beneficial to increase the thickness of the CsPbBr3 films, optimize the surface morphology, and significantly enhance the light absorption capacity. Finally, the PCE of the CsPbBr3 solar cells obtained after conditions optimization was 5.76%.

1. Introduction

From the perspective of the preparation process, a very unique advantage of perovskite solar cells (PSCs) is that they have better processability [1,2,3,4]. Whether an organic-inorganic hybrid or all-inorganic perovskite solar cell, the film quality of the perovskite active layer (including morphology, grain shape, grain size, number of grain boundaries, whether there are pinholes, etc.) determines the PCE and durability of PSCs [5,6,7,8,9]. Generally, the prepared perovskite film should have a high-density microstructure, large grain size, low grain boundary density, and no pinholes or other voids in the film, to improve electron/hole mobility, achieve reduction in radiation recombination, and other purposes [1,10,11,12]. Both the preparation method of the perovskite film and the device structure (planar or mesoporous structure) can exert a significant impact on the quality and photoelectric performance of the perovskite film. In the process of preparing perovskite films by the two-step method, mesoporous substrates are often used to promote the PCE of PSCs. This is because the mesoporous layer formed by oxide nanoparticles has a higher specific surface area, and a large number of pores promote the improvement in the effective diffusivity of the lead halide precursor solution [13], so denser lead halide nanosized grains can be generated faster in the mesoporous layer, and the conversion process of lead halide to perovskite can be accelerated [1,14]. However, during this conversion process, due to some unconverted lead halide residues in the mesoporous layer and the possible formation of random crystals in the perovskite film [14,15], the efficiency of PSCs may be greatly affected. A method for fully converting the lead halide is a critical prerequisite for controlling the growth of perovskite films, or ensuring the perfect formation of crystals and enhancing the efficiency of PSCs [16,17,18].
Introducing a porous lead halide film into a planar structure is an effective method to replace the mesoporous structure in the two-step method [1]. The porous structure in the lead halide film can enlarge the contact area between the lead halide and other precursor solutions and increase the reaction rate, and it can reduce space-expansion-induced defects and residual stress during the expansion of the film [5]. Liu et al. [19] spin-coated a solution of PbI2 in DMF onto a substrate, then prepared a PbI2 film with a porous structure by adjusting the standing time. Ostwald ripening of the PbI2 crystallites can result in larger crystals and increase the void space between grains. The increase in porosity in the PbI2 film effectively promotes the transformation process of perovskite and reduces the residual amount of PbI2. The resulting CH3NH3PbI3 PSCs with inverted planar heterojunction have an efficiency of 15.7%, and there is almost no hysteresis. Dong et al. [20] confirmed that the porosity in PbI2 can affect the phase composition and film morphology of CsPbI2Br. Compared with other anti-solvents, the PbI2 (DMSO) film treated with green ethanol has higher porosity and randomly distributed crystals, which provide more diffusion paths and contact areas for the CsBr precursor, and create an essential prerequisite for ultimately obtaining CsPbI2Br film with high purity, high thickness, large grains, and full coverage. Although the CsPbBr3 film prepared by the multi-step method exhibits excellent moisture and thermal stability [21,22,23,24], it is also prone to problems such as low film coverage, thin film thickness, and the presence of impurity phases (such as CsPb2Br5) [5,25,26]. As mentioned above, a porous PbBr2 film with higher porosity can provide more diffusion paths for the CsBr precursor and increase the contact area between the two, which ultimately provides a favorable condition for the formation of CsPbBr3 with high purity phase and high coverage [27,28,29]. Zhao et al. [21] prepared porous PbBr2 films on SnO2 substrates by precisely controlling the crystallization temperature. The porous structure of PbBr2 not only allows the effective diffusion of CsBr solution to gain high-purity CsPbBr3, but also offers enough space for the growth of CsPbBr3 grains under stress-free conditions, resulting in high-quality CsPbBr3 film with larger grain size and lower grain boundary density. In our previous work, we prepared porous PbBr2 films on mesoporous TiO2 (m-TiO2) by introducing InBr3 into the PbBr2 precursor solution, and achieved In3+ or In cluster doping with CsPbBr3, thereby improving the growth quality of CsPbBr3 films and enhancing the efficiency of CsPbBr3 solar cells with mesoporous structure [30].
At present, it is a highly attractive choice to use carbon electrodes, which are low cost and highly stable, to construct CsPbBr3 PSCs without hole transport layer (HTL). Carbon, which matches the work function of perovskite, can effectively transport holes, and avoid the instability of traditional organic HTL (such as Spiro-MeOTAD) and the low conductivity of metal oxide HTL. At the same time, compared with Au electrodes, the carbon electrode is lower cost, and does not diffuse into the perovskite and cause the device performance to deteriorate as does the Ag electrode [31,32,33]. On the basis of our previous research results, in this work, we introduced InBr3 into a PbBr2 precursor solution and adjusted the concentration of PbBr2 to prepare a high-quality CsPbBr3 film with high thickness, high density, and no holes on c-TiO2, in order to further simplify the structure of the CsPbBr3 PSCs and optimize the growth quality of the CsPbBr3 film. When the concentration of InBr3 is 0.21 M and PbBr2 concentration is 1.3 M, the prepared carbon-based CsPbBr3 solar cells with a planar structure show the best 5.76% PCE with a short circuit current density (JSC) of 6.52 mA/cm2, an open circuit voltage (VOC) of 1.29 V, and a fill factor (FF) of 0.68.

2. Materials and Methods

2.1. Materials

The fluorine-doped tin oxide coated glass (FTO, 6 Ω/□) was purchased from Opvtech New Energy Co., Ltd. (Yingkou, China). Titanium (IV) isopropoxide (99.9%), N,N-dimethylformamide (DMF, chromatographic grade, ≥99.9%), methanol (chromatographic grade, ≥99.9%), ethanol (chromatographic grade, ≥99.8%), and isopropanol (≥99.5%) were purchased from Aladdin (Shanghai, China). PbBr2 (99.99%) and CsBr (99.9%) were purchased from Xi’an Polymer Light Technology Corp (Xi’an, China). InBr3 (99.9%) was purchased from Macklin Biochemical Co., Ltd. (Shanghai, China). The commercial carbon paste was purchased from Shanghai MaterWin New Materials Co., Ltd. (Shanghai, China).

2.2. Device Fabrication

All the following processes were carried out in an air atmosphere. The c-TiO2 was spin-coated on pretreated FTO substrates using 0.2 M titanium isopropoxide in ethanol at 5000 rpm for 30 s and dried for 10 min at 120 °C, then the substrates were annealed at 500 °C for 1 h. Perovskite films were synthesized by a multistep solution-processing method. We added 0.21 mmol of InBr3 to 1 mL DMF solution of 1.0, 1.1, 1.2, 1.3, or 1.4 M PbBr2, which was stirred under 90 °C. Afterward, the mixed solution was spin-coated on the FTO/c-TiO2 substrate at 2000 rpm for 30 s and heated to 90 °C for 30 min to obtain InBr3:PbBr2 films with a porous structure. Then, the methanol solution of CsBr (0.07 M) was spin-coated on InBr3:PbBr2 film at 5000 rpm for 30 s and continuingly heated at 250 °C for 5 min. This step was performed 6 times. After the prepared sample was soaked in isopropanol for 30 min and annealed at 250 °C for 15 min, the blade coating method was applied to coat carbon paste onto the perovskite films and heated at 100 °C for 10 min to form the carbon back electrode with the effective area of 0.09 cm−2.

2.3. Characterization

The X-ray diffraction (XRD) patterns of the synthesized sample were investigated using an X-ray diffractometer (Cu Kα radiation, λ = 1.5418 Å, Rigaku D/max2500, Tokyo, Japan). The morphologies of the synthesized films were recorded by a scanning electron microscope (SEM, FEI MAGELLAN 400, FEI, Hillsboro, OR, USA), and the system was connected to an energy-dispersive X-ray spectroscopy (EDS). X-ray photoelectron spectroscopy (XPS) was performed by an X-ray photoelectron spectrometer system (K-Alpha, Thermo Fisher Scientific, Waltham, MA, USA) equipped with a monochromatic Al Kα X-ray source (1486.6 eV) operating at 100 W. The steady-state photoluminescence (PL) and the time-resolved photoluminescence (TRPL) decay spectrums of perovskite films were collected on a photoluminescence spectrometer (FLS980, Edinburgh Instruments, Livingston, UK) with a 473 nm excitation source. The absorption spectrum was measured by a UV–Vis spectrometer (UV-3600, Shimadzu, TKY, Japan) in the range of 200 to 800 nm. The current–voltage (J–V) characteristics were measured by a solar cells test system (XP3000, Sanyou, China) equipped with a Keithley 2400 source meter. The external quantum efficiency (EQE) was tested by an EQE measured system (QTest Station 1000A, CROWNTECH, INC., Macungie, PA, USA) under DC mode.

3. Results and Discussion

Figure 1 displays the top-view SEM images prepared without and with introducing 0.21 M InBr3 in PbBr2 precursors of different solubility. When the PbBr2 precursor concentration was 1.0 M without introducing InBr3, the PbBr2 film produced was uniform, dense, crack-free, and had a smooth surface, and there were pores of different sizes in the film (Figure 1a). In contrast, after the introduction of InBr3 into the 1.0 M PbBr2 precursor solution, as shown in Figure 1b, the InBr3:CsPbBr3 film had a porous structure, the crystal grains were mostly block structures with obvious size differences, and there were relatively small gaps between the crystal grains. This indicates that InBr3 can transform the PbBr2 grown on the c-TiO2 substrate from a dense film to a relatively low-density porous film. Compared with the porous PbBr2 film formed on m-TiO2, the morphology of the PbBr2 film prepared on c-TiO2 changed significantly regardless of the introduction of InBr3, which reveals that both the TiO2 substrate and InBr3 can directly affect the morphology of the PbBr2 film and determine the quality of the subsequent CsPbBr3 film growth. With the further increase in the PbBr2 concentration (1.1~1.4 M), although the films prepared under various conditions were still porous films, the morphology and porosity of the films did not illustrate obvious differences. The columnar and cubic crystal grains grouped into clusters, with gaps between the clusters, which provide space for the subsequent growth and free expansion of CsPbBr3 crystals [13,34]. Based on the cross-sectional SEM images of the corresponding PbBr2 films (Figure S1, Supplementary Materials), it can be concluded that the thickness of the film (330, 400, 450, 500, and 550 nm), the grain size, and the gap between the crystal grains (clusters) all exhibit an increasing trend with the increase in the PbBr2 concentration, which provides the pivotal condition to enhance the diffusion rate of the CsBr precursor and increase the contact area and reaction efficiency between CsBr and PbBr2, and subsequently guarantee the production of high-quality CsPbBr3 films with high thickness, high purity, high density, and large crystal grains.
To validate the influence of porous lead bromide films made by introducing InBr3 on the morphologies of the CsPbBr3 films, the top-view and cross-sectional SEM images illustrated in Figure 2 were examined. When the PbBr2 precursor concentration was 1.0 M without introducing InBr3, that is, when the PbBr2 film was a flat film, the thickness of the obtained CsPbBr3 film was about 290 nm, the film showed poor uniformity and coverage, and many holes in the film (Figure 2a). With the further increase in PbBr2 concentration (1.1~1.4 M), the thickness of each perovskite film was 310, 330, 350, and 370 nm, respectively (Figure S2, Supplementary Material). When 0.21 M InBr3 was introduced into the PbBr2 precursor solution, with the increase in the PbBr2 concentration (1.0~1.3 M), the CsPbBr3 film showed a satisfying growth relationship with c-TiO2, and the film thickness significantly increased to approximately 430, 470, 500, and 540 nm, respectively, which reveals that the porous structure of the PbBr2 film is the main factor affecting the increase in the thickness of the CsPbBr3 film. Meanwhile, the coverage of the CsPbBr3 film gradually rose with the increase in the PbBr2 concentration, with the pores in the film gradually decreasing, the average size of the crystal grains increasing, and the density and uniformity of the perovskite film significantly improving. In particular, when the PbBr2 concentration was 1.3 M, the surface of the CsPbBr3 film was smooth and dense with almost no holes. As mentioned above, the porous structure of the PbBr2 film can increase the contact area between the CsBr precursor and PbBr2 and make the two fully react, thereby making the process of converting PbBr2 to CsPbBr3 faster and more complete. With the increase in the PbBr2 concentration, that is, as the gap between the PbBr2 crystal grain clusters gradually enlarges, more uniform perovskite nuclei with a narrower size distribution are generated in the porous PbBr2 film in a shorter time. Small grains with unfavorable growth orientation transfer their monomers through the solid-state grain boundary to at least a few larger grains with decent growth orientation, until smaller grains disappear to generate large grains with the largest surface and interface area, and finally producing a large-grain CsPbBr3 film that completely covers the substrate and has no holes [1,35]. However, as the concentration of the PbBr2 precursor containing InBr3 was further increased to 1.4 M, the thickness of the formed CsPbBr3 film further increased to about 580 nm, but the surface morphology of the CsPbBr3 film began to deteriorate, and holes appeared in the film again, which can adversely affect the JSC of CsPbBr3 solar cells [36].
To investigate the effect of porous PbBr2 film on the phase, structure, and element composition of the perovskite, the XRD and EDS mappings were employed to characterize the CsPbBr3 films. The XRD patterns demonstrated in Figure 3a reveal that all CsPbBr3 films are cubic structures (PDF# 54-0752) [21,37], which reflects that the introduction of InBr3 and the concentration change of the PbBr2 film will not change the phase of CsPbBr3. When the PbBr2 precursor concentration was 1.0 M without introducing InBr3, a strong diffraction peak belonging to the PbBr2-rich CsPb2Br5 phase appeared at 2θ of 11.7°. In contrast, after introducing InBr3 into the 1.0 M PbBr2 precursor, the diffraction peak belonging to CsPb2Br5 phase disappeared (CsPb2Br5 will not be completely consumed), but the diffraction peak belonging to Cs4PbBr6 phase appeared at 2θ of 12.7°. This is because after the porous PbBr2 film with low thickness is fully contacted with the CsBr solution and is completely consumed, the excess CsBr and CsPbBr3 continue to react and form a CsBr-rich Cs4PbBr6 phase. On the contrary, in the planar structure of PbBr2, the contact area and the reaction efficiency between CsBr and PbBr2 are relatively low, and thus PbBr2 is not fully consumed. After the introduction of InBr3, as the concentration of the PbBr2 precursor gradually increased from 1.1 to 1.4 M, the intensity of the diffraction peaks belonging to the CsPb2Br5 phase in the XRD patterns of each sample showed an upward trend, and no obvious diffraction peak belonging to Cs4PbBr6 phase was observed. The cross-sectional EDS mapping of the CsPbBr3 film shown in Figure 3b proves that Cs, Pb, Br, and In are evenly distributed in the analyzed area, and there is no Pb or In gathering in a certain area.
Afterward, we coated the prepared CsPbBr3 films with commercial carbon paste, and dried them to obtain the planar HTL-free PSCs with the structure of FTO/c-TiO2/CsPbBr3/carbon; the cross-sectional SEM image of the device is shown in Figure 4a. In order to further explore the effect of PbBr2 porous film on the device performance, Figure 4b–f illustrates the J–V curves under reverse scanning of the PSCs prepared under the conditions of different concentrations of PbBr2 precursor solutions without and with introducing 0.21 M InBr3, and the corresponding parameters are summarized in Table 1. When the PbBr2 concentration was gradually increased from 1.0 to 1.4 M without introducing InBr3 into the precursor solution, the efficiency of each CsPbBr3 cell was at a relatively low level, and there was little difference in the results. It is noteworthy that the efficiency of the CsPbBr3 cell obtained under the condition of a PbBr2 concentration of 1.2 M was the highest, namely 1.79%, while the efficiency of the other cells was between 1.5% and 1.7%. After the introduction of InBr3 into the PbBr2 precursor solution, the efficiency of all PSCs was significantly improved. As the PbBr2 concentration increased from 1.0 to 1.3 M, JSC, VOC, FF, and PCE all showed a trend of increasing with the increase in PbBr2 concentration. When the PbBr2 concentration was 1.3 M, the champion device with a PCE of 5.76% showed a JSC of 6.52 mA/cm2, a VOC of 1.29 V, and an FF of 0.68. When the PbBr2 concentration was further increased to 1.4 M, the efficiency of the device decreased to 4.97%, and JSC, VOC, and FF also illustrated a slight decline, caused by the slight deterioration in the morphology of CsPbBr3. The statistics of photovoltaic parameters collected from 15 devices (Figure S3, Supplementary Materials) and the average photovoltaic parameters with a small standard deviation (Table S1, Supplementary Materials) demonstrate that the InBr3:CsPbBr3 PSCs synthesized with different PbBr2 concentrations obtained in this work have decent reliability and reproducibility. Figure 4g displays the EQE and the corresponding integrated photocurrent densities to verify the validity of the CsPbBr3 devices based on InBr3. The photoresponse edges of these five devices are all around 540 nm. When the PbBr2 concentration is 1.3 M and the maximum EQE value of the champion device is 83%, the EQE values in the light absorption wavelength range of 350–525 nm are significantly enhanced, which proves that the device has better charge injection ability and collection capacity. In addition, the integrated photocurrent density of each device is closer to the corresponding JSC, and the mismatch is less than 5%, which signifies the reliability of the J–V curves.
Furthermore, XPS was utilized to study the influence of InBr3 on the electronic state of CsPbBr3 films at a PbBr2 concentration of 1.3 M. As displayed in Figure 5a, compared with the pure CsPbBr3, in addition to Cs 3d, Pb 4f, and Br 3d, In 3d was found in the XPS spectra for InBr3: CsPbBr3 film. According to Figure 5b, In 3d5/2 and In 3d3/2 peaked at 444.9 eV and 452.3 eV, respectively, and the high-resolution core-level binding energies of Cs 3d, Pb 4f, and Br 3d all shifted toward higher values compared with those in the pure CsPbBr3, which indicates that part of the In(III) in the porous PbBr2 was doped into the perovskite lattice in the process of transforming CsPbBr3, and caused the chemical state of [PbBr6]4− octahedron to change [38]. The incorporation of In3+ or In cluster not only improves the spatial symmetry of the CsPbBr3 lattice structure, but also reduces vacancy defects, which is beneficial to the extraction and transfer process of the charge, and finally significantly optimizes the JSC of the device [30,38,39,40].
In the optical property test, the devices with the FTO/c-TiO2/CsPbBr3 structure were further characterized by a UV–visible spectrophotometer and PL. In Figure 6a, as expected, the absorption edges of all perovskite films are at around 533 nm, which is basically the same as the EQE test result. The light absorption capacity of the films shows a trend of first increasing and then decreasing with increasing PbBr2 concentration, which is consistent with the change law of the perovskite morphology and PCE. In other words, the increase in the thickness of the CsPbBr3 film and the optimization of the surface morphology can enhance the light absorption rate of the perovskite layer and reduce the light loss, which is conducive to generating more electrons to improve JSC. Using Tuac’s equation, we can calculate the corresponding band gap (Eg) patterns of the CsPbBr3 film; the curves of α2 versus the photon energy hν, as shown in Figure 6b; and the Eg of the prepared CsPbBr3 to be about 2.35 eV. Regardless of whether InBr3 was introduced into the PbBr2 precursor, the solubility of PbBr2 did not meaningfully change the Eg of CsPbBr3. Figure S4 (Supplementary Material) illustrates the PL spectra of the PSCs prepared under different concentrations of PbBr2 precursor solutions with introducing 0.21 M InBr3. All the InBr3:CsPbBr3 films presented a typical emission band around 523 nm, which is close to their optical band gap. The PL intensity decreased as the PbBr2 concentration increased from 1.0 to 1.3 M, and then increased as the PbBr2 concentration further increased to 1.4 M. The InBr3:CsPbBr3 film synthesized with the PbBr2 concentration of 1.3 M showed the lowest PL intensity. When the PbBr2 precursor concentration was 1.3 M with/without introducing InBr3, as shown in Figure 6c, the PL peaks for both the CsPbBr3 films were still at about 523 nm, which indicates that the change in the concentration of PbBr2 and the introduction of InBr3 do not change the band position. However, the perovskite prepared after the introduction of InBr3 into the PbBr2 precursor solution showed more observable strong quenching, which proves that the incorporation of In3+ or In cluster and the improvement of perovskite film growth quality are beneficial to faster electron injection and collection, and reduce the recombination of electrons and holes [41,42], which explains why the JSC and VOC of the PSCs were significantly improved after the introduction of InBr3 in Figure 4b–f. Figure 6d illustrates the TRPL decay curves, and the relevant detailed parameters of the corresponding CsPbBr3 films are listed in Table S2. The specific lifetime of the carriers can be determined by a bi-exponential formula I = A 1 e ( τ τ 0 ) / τ 1 + A 2 e ( τ τ 0 ) / τ 2 (A1 and A2 are the relative amplitude, respectively; τ1 and τ2 are the fast and slow decay time, respectively) [43], and the average lifetimes (τave) is calculated by the formula   τ ave = ( A 1 τ 1 2 + A 2 τ 2 2 ) / ( A 1 τ 1 + A 2 τ 2 ) [44]. Significantly decreased lifetimes of τ1, τ2, and τave of the InBr3:CsPbBr3 film (1.35, 3.06 and 2.49 ns, respectively) compared with the control CsPbBr3 (2.65, 5.61, and 3.95 ns, respectively) were observed, representing lower trap states in the InBr3:CsPbBr3 film and an accelerated electron transport process. This phenomenon is brought about by the enhancement in the growth quality of InBr3:CsPbBr3 film and the reduction in the defects of Pb2+ and Br after partial substitution of Pb2+ by In3+ or In cluster [25,39]. This is conducive to improving the VOC and fill factor (FF) of the devices [45].

4. Conclusions

By introducing InBr3 into a PbBr2 precursor solution, we successfully prepared planar carbon-based CsPbBr3 solar cells with higher perovskite film quality in this work. The introduction of InBr3 can transform the dense PbBr2 film originally grown on c-TiO2 into a porous structure. On this basis, appropriately increasing the PbBr2 concentration in the DMF precursor solution can enhance the thickness of the CsPbBr3 film and optimize the surface morphology without changing the phase of CsPbBr3, which can effectively alleviate the low thickness and poor quality of the CsPbBr3 film in the multi-step process. Due to the increase in film thickness and the incorporation of In3+ or In cluster, the light absorption capacity of the CsPbBr3 film is significantly improved. Therefore, the efficiency and EQE of the manufactured devices are significantly optimized. As InBr3 is introduced into the current flooding fluid and the PbBr2 concentration is 1.3 M, the small-area CsPbBr3 solar cell has the best PCE, which is 5.76%, and the maximum EQE value is 83%. This work provides a certain reference for the preparation of high-quality CsPbBr3 thin films, simplifying the structure and developing planar HTL-free CsPbBr3 solar cells.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/nano11092408/s1, Figure S1: Cross-sectional SEM images of PbBr2 films synthesized with different PbBr2 concentration: (a) 1.0 M w/ InBr3; (b) 1.1 M w/ InBr3; (c) 1.2 M w/ InBr3; (d) 1.3 M w/ InBr3; (e) 1.4 M w/ InBr3; Figure S2: Cross-sectional SEM images of CsPbBr3 films synthesized with different PbBr2 concentration: (a) 1.1 M w/o InBr3; (b) 1.2 M w/o InBr3; (c) 1.3 M w/o InBr3; (d) 1.4 M w/o InBr3; Figure S3: The schematic diagram of of 15 CsPbBr3 PSCs synthesized with different PbBr2 concentration: (a) JSC, (b) VOC, (c) FF and (e) PCE. Table S1: The average photovoltaic parameters of 15 CsPbBr3 PSCs synthesized with different PbBr2 concentration; Figure S4: PL spectra of the InBr3:CsPbBr3 synthesized with different PbBr2 concentration; Table S2: The fitted data of TRPL characterization.

Author Contributions

K.C., X.M., and Q.L. conceived the idea; K.C. and X.M. performed research and analyzed data; K.C. drafted the manuscript; H.X., B.F., D.Y., and Q.L. provided assistance for data acquisition and data analysis. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the Scientific Research Staring Foundation for the Doctors of the Northeast Electric Power University (No. BSJXM-2018223).

Data Availability Statement

The data is available on reasonable request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dunlap-Shohl, W.; Zhou, Y.; Padture, N.P.; Mitzi, D.B. Synthetic Approaches for Halide Perovskite Thin Films. Chem. Rev. 2019, 119, 3193–3295. [Google Scholar] [CrossRef]
  2. Yuan, J.; Ling, X.; Yang, D.; Li, F.; Zhou, S.; Shi, J.; Qian, Y.; Hu, J.; Sun, Y.; Yang, Y.; et al. Band-Aligned Polymeric Hole Transport Materials for Extremely Low Energy Loss α-CsPbI3 Perovskite Nanocrystal Solar Cells. Joule 2018, 2, 2450–2463. [Google Scholar] [CrossRef] [Green Version]
  3. Jeon, I.; Seo, S.; Sato, Y.; Delacou, C.; Anisimov, A.; Suenaga, K.; Kauppinen, E.I.; Maruyama, S.; Matsuo, Y. Perovskite Solar Cells Using Carbon Nanotubes Both as Cathode and as Anode. J. Phys. Chem. C 2017, 121, 25743–25749. [Google Scholar] [CrossRef]
  4. Weber, S.; Rath, T.; Mangalam, J.; Kunert, B.; Coclite, A.M.; Bauch, M.; Dimopoulos, T.; Trimmel, G. Investigation of NiOX-hole Transport Layers in Triple Cation Perovskite Solar Cells. J. Mater. Sci. Mater. Electron. 2018, 29, 1847–1855. [Google Scholar] [CrossRef] [Green Version]
  5. Chen, W.; Li, X.; Li, Y.; Li, Y. A review: Crystal Growth for High-Performance All-Inorganic Perovskite Solar Cells. Energy Envi-ron. Sci. 2020, 13, 1971–1996. [Google Scholar] [CrossRef]
  6. Wang, K.; Liu, C.; Du, P.; Zhang, H.-L.; Gong, X. Efficient Perovskite Hybrid Solar Cells Through a Homogeneous High-Quality Organolead Iodide Layer. Small 2015, 11, 3369–3376. [Google Scholar] [CrossRef] [PubMed]
  7. Yuan, S.; Qiu, Z.; Gao, C.; Zhang, H.; Jiang, Y.; Li, C.; Yu, J.; Cao, B. High-Quality Perovskite Films Grown with a Fast Solvent-Assisted Molecule Inserting Strategy for Highly Efficient and Stable Solar Cells. ACS Appl. Mater. Interfaces 2016, 8, 22238–22245. [Google Scholar] [CrossRef]
  8. Hu, H.; Ren, Z.; Fong, P.W.; Qin, M.; Liu, D.; Lei, D.; Lu, X.; Li, G. Room-Temperature Meniscus Coating of >20% Perovskite Solar Cells: A Film Formation Mechanism Investigation. Adv. Funct. Mater. 2019, 29, 1900092. [Google Scholar] [CrossRef]
  9. Liu, Z.; Deng, K.; Zhu, Y.; Wang, M.; Li, L. Iodine Induced PbI 2 Porous Morphology Manipulation for High-Performance Planar Perovskite Solar Cells. Sol. RRL 2018, 2, 1800230. [Google Scholar] [CrossRef]
  10. Aristidou, N.; Eames, C.; Sanchez-Molina, I.; Bu, X.; Kosco, J.; Islam, M.S.; Haque, S.A. Fast oxygen diffusion and iodide defects mediate oxygen-induced degradation of perovskite solar cells. Nat. Commun. 2017, 8, 15218. [Google Scholar] [CrossRef]
  11. Yen, H.-J.; Liang, P.-W.; Chueh, C.-C.; Yang, Z.; Jen, A.K.-Y.; Wang, H.-L. Large Grained Perovskite Solar Cells Derived from Single-Crystal Perovskite Powders with Enhanced Ambient Stability. ACS Appl. Mater. Interfaces 2016, 8, 14513–14520. [Google Scholar] [CrossRef] [PubMed]
  12. Swarnkar, A.; Marshall, A.R.; Sanehira, E.M.; Chernomordik, B.D.; Moore, D.T.; Christians, J.A.; Chakrabarti, T.; Luther, J.M. Quantum dot-induced phase stabilization of -CsPbI3 perovskite for high-efficiency photovoltaics. Science 2016, 354, 92–95. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Zhou, Y.; Game, O.S.; Pang, S.; Padture, N.P. Microstructures of Organometal Trihalide Perovskites for Solar Cells: Their Evolution from Solutions and Characterization. J. Phys. Chem. Lett. 2015, 6, 4827–4839. [Google Scholar] [CrossRef] [PubMed]
  14. Ummadisingu, A.; Grätzel, M. Revealing the detailed path of sequential deposition for metal halide perovskite formation. Sci. Adv. 2018, 4, e1701402. [Google Scholar] [CrossRef] [Green Version]
  15. Scherer, G.W. Stress from crystallization of salt. Cem. Concr. Res. 2004, 34, 1613–1624. [Google Scholar] [CrossRef]
  16. Chen, Q.; Zhou, H.; Song, T.-B.; Luo, S.; Hong, Z.; Duan, H.-S.; Dou, L.; Liu, Y.; Yang, Y. Controllable Self-Induced Passivation of Hybrid Lead Iodide Perovskites toward High Performance Solar Cells. Nano Lett. 2014, 14, 4158–4163. [Google Scholar] [CrossRef]
  17. Carmona, C.R.; Gratia, P.; Zimmermann, I.; Grancini, G.; Gao, P.; Graetzel, M.; Nazeeruddin, M.K. High efficiency methylammonium lead triiodide perovskite solar cells: The relevance of non-stoichiometric precursors. Energy Environ. Sci. 2015, 8, 3550–3556. [Google Scholar] [CrossRef]
  18. Han, Q.; Ding, J.; Bai, Y.; Li, T.; Ma, J.-Y.; Chen, Y.-X.; Zhou, Y.; Liu, J.; Ge, Q.-Q.; Chen, J.; et al. Carrier Dynamics Engineering for High-Performance Electron-Transport-Layer-free Perovskite Photovoltaics. Chem 2018, 4, 2405–2417. [Google Scholar] [CrossRef] [Green Version]
  19. Liu, T.; Hu, Q.; Wu, J.; Chen, K.; Zhao, L.; Liu, F.; Wang, C.; Lu, H.; Jia, S.; Russell, T.P.; et al. Mesoporous PbI2 Scaffold for High-Performance Planar Heterojunction Perovskite Solar Cells. Adv. Energy Mater. 2016, 6, 1501890. [Google Scholar] [CrossRef]
  20. Dong, C.; Han, X.; Li, W.; Qiu, Q.; Wang, J. Anti-solvent assisted multi-step deposition for efficient and stable carbon-based CsPbI2Br all-inorganic perovskite solar cell. Nano Energy 2019, 59, 553–559. [Google Scholar] [CrossRef]
  21. Zhao, Y.; Duan, J.; Wang, Y.; Yang, X.; Tang, Q. Precise stress control of inorganic perovskite films for carbon-based solar cells with an ultrahigh voltage of 1.622 V. Nano Energy 2020, 67, 104286. [Google Scholar] [CrossRef]
  22. Tian, J.; Xue, Q.; Yao, Q.; Li, N.; Brabec, C.J.; Yip, H. Inorganic Halide Perovskite Solar Cells: Progress and Challenges. Adv. Energy Mater. 2020, 10, 2000183. [Google Scholar] [CrossRef]
  23. 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]
  24. Zhang, J.; Hodes, G.; Jin, Z.; Liu, F. All-Inorganic CsPbX 3 Perovskite Solar Cells: Progress and Prospects. Angew. Chem. Int. Ed. 2019, 58, 15596–15618. [Google Scholar] [CrossRef] [PubMed]
  25. Subhani, W.S.; Wang, K.; Du, M.; Wang, X.; Liu, F. Interface-Modification-Induced Gradient Energy Band for Highly Efficient CsPbIBr 2 Perovskite Solar Cells. Adv. Energy Mater. 2019, 9, 1803785. [Google Scholar] [CrossRef]
  26. Duan, J.; Zhao, Y.; He, B.; Tang, Q. High-Purity Inorganic Perovskite Films for Solar Cells with 9.72% Efficiency. Angew. Chem. Int. Ed. 2018, 57, 3787–3791. [Google Scholar] [CrossRef]
  27. Sun, K.; Hu, Z.; Shen, B.; Lu, C.; Huang, L.; Zhang, J.; Zhang, J.; Zhu, Y. Lewis Acid–Base Interaction-Induced Porous PbI2 Film for Efficient Planar Perovskite Solar Cells. ACS Appl. Energy Mater. 2018, 1, 2114–2122. [Google Scholar] [CrossRef]
  28. Cao, J.; Wang, F.; Yu, H.; Zhou, Y.; Lu, H.; Zhao, N.; Wong, C.-P. Porous PbI2 films for the fabrication of efficient, stable perovskite solar cells via sequential deposition. J. Mater. Chem. A 2016, 4, 10223–10230. [Google Scholar] [CrossRef]
  29. Jena, A.K.; Kulkarni, A.; Miyasaka, T. Halide Perovskite Photovoltaics: Background, Status, and Future Prospects. Chem. Rev. 2019, 119, 3036–3103. [Google Scholar] [CrossRef]
  30. Meng, X.; Chi, K.; Li, Q.; Feng, B.; Wang, H.; Gao, T.; Zhou, P.; Yang, H.; Fu, W. Fabrication of Porous Lead Bromide Films by Introducing Indium Tribromide for Efficient Inorganic CsPbBr3 Perovskite Solar Cells. Nanomaterials 2021, 11, 1253. [Google Scholar] [CrossRef]
  31. Kato, Y.; Ono, L.K.; Lee, M.V.; Wang, S.; Raga, S.R.; Qi, Y. Silver Iodide Formation in Methyl Ammonium Lead Iodide Perovskite Solar Cells with Silver Top Electrodes. Adv. Mater. Interfaces 2015, 7, 1500195. [Google Scholar] [CrossRef]
  32. Kulkarni, A.; Ünlü, F.; Pant, N.; Kaur, J.; Bohr, C.; Jena, A.K.; Öz, S.; Yanagida, M.; Shirai, Y.; Ikegami, M.; et al. Concerted Ion Migration and Diffusion-Induced Degradation in Lead-Free Ag3BiI6 Rudorffite Solar Cells under Ambient Conditions. Sol. RRL 2021, 5, 2100077. [Google Scholar] [CrossRef]
  33. Domanski, K.; Correa-Baena, J.-P.; Mine, N.; Nazeeruddin, M.K.; Abate, A.; Saliba, M.; Tress, W.; Hagfeldt, A.; Grätzel, M. Not All That Glitters Is Gold: Metal-Migration-Induced Degradation in Perovskite Solar Cells. ACS Nano 2016, 10, 6306–6314. [Google Scholar] [CrossRef]
  34. Schlipf, J.; Docampo, P.; Schaffer, C.J.; Körstgens, V.; Bießmann, L.; Hanusch, F.; Giesbrecht, N.; Bernstorff, S.; Bein, T.; Müller-Buschbaum, P. A Closer Look into Two-Step Perovskite Conversion with X-ray Scattering. J. Phys. Chem. Lett. 2015, 6, 1265–1269. [Google Scholar] [CrossRef]
  35. Thompson, C.V. Solid-State Dewetting of Thin Films. Annu. Rev. Mater. Res. 2012, 42, 399–434. [Google Scholar] [CrossRef]
  36. Yang, L.; Wang, J.; Leung, W.W.-F. Lead Iodide Thin Film Crystallization Control for High-Performance and Stable Solution-Processed Perovskite Solar Cells. ACS Appl. Mater. Interfaces 2015, 7, 14614–14619. [Google Scholar] [CrossRef] [PubMed]
  37. Liu, X.; Tan, X.; Liu, Z.; Ye, H.; Sun, B.; Shi, T.; Tang, Z.; Liao, G. Boosting the efficiency of carbon-based planar CsPbBr3 perovskite solar cells by a modified multistep spin-coating technique and interface engineering. Nano Energy 2019, 56, 184–195. [Google Scholar] [CrossRef]
  38. Liu, C.; Li, W.; Li, H.-Y.; Wang, H.; Zhang, C.; Yang, Y.; Gao, X.; Xue, Q.; Yip, H.-L.; Fan, J.; et al. Structurally Reconstructed CsPbI2 Br Perovskite for Highly Stable and Square-Centimeter All-Inorganic Perovskite Solar Cells. Adv. Energy Mater. 2019, 9, 1803572. [Google Scholar] [CrossRef]
  39. Wang, Z.-K.; Li, M.; Yang, Y.; Hu, Y.; Ma, H.; Gao, X.-Y.; Liao, L.-S. High Efficiency Pb-In Binary Metal Perovskite Solar Cells. Adv. Mater. 2016, 28, 6695–6703. [Google Scholar] [CrossRef] [PubMed]
  40. Chen, L.; Wan, L.; Li, X.; Zhang, W.; Fu, S.; Wang, Y.; Li, S.; Wang, H.-Q.; Song, W.; Fang, J. Inverted All-Inorganic CsPbI2Br Perovskite Solar Cells with Promoted Efficiency and Stability by Nickel Incorporation. Chem. Mater. 2019, 31, 9032–9039. [Google Scholar] [CrossRef]
  41. Ko, H.-S.; Lee, J.-W.; Park, N.-G. 15.76% efficiency perovskite solar cells prepared under high relative humidity: Importance of PbI2 morphology in two-step deposition of CH3NH3PbI3. J. Mater. Chem. A 2015, 3, 8808–8815. [Google Scholar] [CrossRef]
  42. Chen, J.; Zhao, X.; Kim, S.; Park, N. Multifunctional Chemical Linker Imidazoleacetic Acid Hydrochloride for 21% Efficient and Stable Planar Perovskite Solar Cells. Adv. Mater. 2019, 31, 1902902. [Google Scholar] [CrossRef] [PubMed]
  43. Yuan, H.; Zhao, Y.; Duan, J.; Wang, Y.; Yang, X.; Tang, Q. All-inorganic CsPbBr3 perovskite solar cell with 10.26% efficiency by spectra engineering. J. Mater. Chem. A 2018, 6, 24324–24329. [Google Scholar] [CrossRef]
  44. Pham, N.D.; Tiong, V.T.; Chen, P.; Wang, L.; Wilson, G.J.; Bell, J.; Wang, H. Enhanced perovskite electronic properties via a modified lead(ii) chloride Lewis acid–base adduct and their effect in high-efficiency perovskite solar cells. J. Mater. Chem. A 2017, 5, 5195–5203. [Google Scholar] [CrossRef] [Green Version]
  45. Ma, J.; Yang, G.; Qin, M.; Zheng, X.; Lei, H.; Chen, C.; Chen, Z.; Guo, Y.; Han, H.; Zhao, X.; et al. MgO Nanoparticle Modified Anode for Highly Efficient SnO2-Based Planar Perovskite Solar Cells. Adv. Sci. 2017, 4, 1700031. [Google Scholar] [CrossRef]
Figure 1. Top-view SEM images of PbBr2 films synthesized with different PbBr2 concentration: (a) 1.0 M w/o InBr3; (b) 1.0 M w/ InBr3; (c) 1.1 M w/ InBr3; (d) 1.2 M w/ InBr3; (e) 1.3 M w/ InBr3; (f) 1.4 M w/ InBr3.
Figure 1. Top-view SEM images of PbBr2 films synthesized with different PbBr2 concentration: (a) 1.0 M w/o InBr3; (b) 1.0 M w/ InBr3; (c) 1.1 M w/ InBr3; (d) 1.2 M w/ InBr3; (e) 1.3 M w/ InBr3; (f) 1.4 M w/ InBr3.
Nanomaterials 11 02408 g001
Figure 2. Top-view and cross-sectional (insets) SEM images of CsPbBr3 films synthesized with different PbBr2 concentration: (a) 1.0 M w/o InBr3; (b) 1.0 M w/ InBr3; (c) 1.1 M w/ InBr3; (d) 1.2 M w/ InBr3; (e) 1.3 M w/ InBr3; (f) 1.4 M w/ InBr3.
Figure 2. Top-view and cross-sectional (insets) SEM images of CsPbBr3 films synthesized with different PbBr2 concentration: (a) 1.0 M w/o InBr3; (b) 1.0 M w/ InBr3; (c) 1.1 M w/ InBr3; (d) 1.2 M w/ InBr3; (e) 1.3 M w/ InBr3; (f) 1.4 M w/ InBr3.
Nanomaterials 11 02408 g002
Figure 3. (a) XRD patterns of InBr3:CsPbBr3 films synthesized with different PbBr2 concentration; (b) elemental distribution mapping images of the InBr3:CsPbBr3 synthesized with the PbBr2 concentration of 1.3 M.
Figure 3. (a) XRD patterns of InBr3:CsPbBr3 films synthesized with different PbBr2 concentration; (b) elemental distribution mapping images of the InBr3:CsPbBr3 synthesized with the PbBr2 concentration of 1.3 M.
Nanomaterials 11 02408 g003
Figure 4. (a) Cross-sectional SEM image of the InBr3:CsPbBr3 device; (bf) J–V characteristics and (g) EQE spectra and integrated photocurrent densities for the InBr3:CsPbBr3 devices synthesized with different PbBr2 concentrations.
Figure 4. (a) Cross-sectional SEM image of the InBr3:CsPbBr3 device; (bf) J–V characteristics and (g) EQE spectra and integrated photocurrent densities for the InBr3:CsPbBr3 devices synthesized with different PbBr2 concentrations.
Nanomaterials 11 02408 g004
Figure 5. (a) XPS spectra and (b) Cs 3d, Pb 4f, Br 3d, and In 3d core spectra of the InBr3:CsPbBr3 synthesized with the PbBr2 concentration of 1.3 M.
Figure 5. (a) XPS spectra and (b) Cs 3d, Pb 4f, Br 3d, and In 3d core spectra of the InBr3:CsPbBr3 synthesized with the PbBr2 concentration of 1.3 M.
Nanomaterials 11 02408 g005
Figure 6. (a) UV–vis absorption spectra and (b) (αhν)2 vs. hν plots for the InBr3:CsPbBr3 devices synthesized with different PbBr2 concentrations; (c) PL and (d) TRPL decay spectra of InBr3:CsPbBr3 synthesized with the PbBr2 concentration of 1.3 M.
Figure 6. (a) UV–vis absorption spectra and (b) (αhν)2 vs. hν plots for the InBr3:CsPbBr3 devices synthesized with different PbBr2 concentrations; (c) PL and (d) TRPL decay spectra of InBr3:CsPbBr3 synthesized with the PbBr2 concentration of 1.3 M.
Nanomaterials 11 02408 g006
Table 1. Key J-V parameters of the corresponding CsPbBr3 solar cells.
Table 1. Key J-V parameters of the corresponding CsPbBr3 solar cells.
CsPbBr3JSC (mA/cm2)VOC (V)FFPCE (%)
1.0 Mw/o InBr32.97 1.09 0.50 1.62
w/ InBr33.82 1.25 0.57 2.72
1.1 Mw/o InBr33.47 0.97 0.49 1.79
w/ InBr34.70 1.26 0.57 3.38
1.2 Mw/o InBr33.40 1.04 0.47 1.66
w/ InBr35.51 1.27 0.60 4.20
1.3 Mw/o InBr33.22 1.07 0.46 1.58
w/ InBr36.52 1.29 0.68 5.76
1.4 Mw/o InBr32.94 1.04 0.50 1.53
w/ InBr35.97 1.27 0.65 4.97
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chi, K.; Xu, H.; Feng, B.; Meng, X.; Yu, D.; Li, Q. Controlled Growth of Porous InBr3: PbBr2 Film for Preparation of CsPbBr3 in Carbon-Based Planar Perovskite Solar Cells. Nanomaterials 2021, 11, 2408. https://0-doi-org.brum.beds.ac.uk/10.3390/nano11092408

AMA Style

Chi K, Xu H, Feng B, Meng X, Yu D, Li Q. Controlled Growth of Porous InBr3: PbBr2 Film for Preparation of CsPbBr3 in Carbon-Based Planar Perovskite Solar Cells. Nanomaterials. 2021; 11(9):2408. https://0-doi-org.brum.beds.ac.uk/10.3390/nano11092408

Chicago/Turabian Style

Chi, Kailin, Hansi Xu, Bingtao Feng, Xianwei Meng, Daoyu Yu, and Qian Li. 2021. "Controlled Growth of Porous InBr3: PbBr2 Film for Preparation of CsPbBr3 in Carbon-Based Planar Perovskite Solar Cells" Nanomaterials 11, no. 9: 2408. https://0-doi-org.brum.beds.ac.uk/10.3390/nano11092408

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