Next Article in Journal
Size Effect on Diffusion and Catalytic Performance of Mordenite in Dimethyl Ether Carbonylation
Previous Article in Journal
Catalytic Steam-Reforming of Glycerol over LDHs-Derived Ni–Al Nanosheet Array Catalysts for Stable Hydrogen Production
Previous Article in Special Issue
Synthesis of All-Inorganic Halide Perovskite Nanocrystals for Potential Photoelectric Catalysis Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of Asymmetrical CsPbBr3/TiO2 Nanocrystals with Enhanced Stability and Photocatalytic Properties

International School of Materials Science and Engineering (ISMSE), State Key Laboratory of Advanced Technology for Material Synthesis and Processing, Wuhan University of Technology, Wuhan 430070, China
*
Author to whom correspondence should be addressed.
Catalysts 2023, 13(7), 1048; https://doi.org/10.3390/catal13071048
Submission received: 5 June 2023 / Revised: 25 June 2023 / Accepted: 26 June 2023 / Published: 28 June 2023
(This article belongs to the Special Issue Hierarchically Porous Catalysts)

Abstract

:
Practical applications of CsPbX3 nanocrystals (NCs) are limited by their poor stability. The formation of a heterojunction between CsPbX3 NCs and oxides is an effective means to protect perovskite from polar solvents and other external factors. Significantly improving the stability and photocatalytic properties of the core/shell perovskite is very important for its application in photoelectric and photocatalytic technology. Here, we report the synthesis of asymmetrical CsPbBr3/TiO2 core–shell heterostructure NCs at the single-particle level by hot-injection liquid-phase synthesis and sol–gel method, where each CsPbBr3 NCs is partially covered by titanium dioxide. We tested not only the optical properties of the material but also the electrochemical impedance and photocurrent density of the material in sodium sulfate solution. It is shown that the type II arrangement is generated at the heterogeneous interface, which greatly facilitates the separation of electron–hole pairs and increases the carrier transport efficiency. Compared with CsPbBr3 NCs, CsPbBr3/TiO2 has higher photocatalytic efficiency. More crucially, due to the protection of the titanium dioxide shell, the product has higher long-term stability in humid air compared with bare CsPbBr3 NCs. The asymmetrical core–shell heterostructure prepared in this study not only improves the stability of CsPbX3 NCs but also provides some ideas for optoelectronic device applications and TiO2-based photocatalysts.

Graphical Abstract

1. Introduction

CsPbBr3 is a typical metal halide perovskite material that has become one of the most promising materials for optoelectronic applications, including solar cells [1,2,3,4], light emitting diodes (LED) [5,6,7,8,9,10], photodetectors [11], and lasers [12,13,14,15,16], due to its excellent photoelectric and catalytic properties such as narrow optical emission range, high photoluminescent quantum yield (PLQY), high photogenerated charge mobility, and tunable band gap. In recent years, the rapid development of research in this field has led to significant advances in the study of optoelectronic properties and materials. Inspired by the results of photovoltaic applications, CsPbBr3 nanocrystals (NCs) is a potential candidate for conducting efficient photocatalysis [17,18]. Since the pioneering work reported by Kovalenko et al. [19] in 2015, significant progress has been made in the preparation methods and photocatalytic applications of CsPbX3 NCs (X = Cl, Br, I).
With the growth in population and economic activities, the increasing demand for water resources and the aggravation of water pollution have further promoted water shortage on a global scale. In the case of abundant seawater resources, the use of catalysts to purify seawater can not only alleviate the shortage of freshwater resources but also greatly reduce the cost of the reaction system process. Solar-powered photocatalysis is one of the effective means to achieve green sustainable fuel preparation and high value-added chemical production in the future [20]. Halide perovskites have become a new type of photocatalyst material due to their advantages, such as high absorption efficiency, long carrier migration distance, and easy adjustment of component structure and morphology. Wang and coworkers reported the preparation of CsPbBr3 NCs, and the obtained CsPbBr3 NCs showed superior photocatalytic activity and excellent stability in a mixed medium of water/methanol vapor [21].
However, perovskite is often described as a crystalline liquid due to the soft and dynamic properties of the crystal lattice, so light-induced carriers are less likely to be trapped and scattered. At room temperature, polarons are formed by the coupling of electrons and holes caused by the ionic properties and strong structural dynamics of the PbX lattice, thus shielding the Coulomb potential and reducing the charge capture and scattering between polarons and between polarons and charged defects or optical phonons. Although there is a high defect tolerance in the lead halide perovskite NCs materials, its defects will still reduce its luminous performance. The high photoluminescence quantum yield of perovskite can only achieve 80–90% in the green–red fluorescence range of the spectrum. In addition, as an ionic crystal, perovskite NCs is not very stable in the process of purification and storage under harsh environmental conditions (such as air, polar solvents, heat, and light). They usually decompose or dissolve quickly, which greatly limits their practical application in photoelectric devices, catalysis, and other fields. The interaction between ligands and atoms on the perovskite NCs surface is highly dynamic. During purification, surface ligand shedding caused by polar solvents leads to an increase in nanocrystal defects, resulting in decomposition and loss of fluorescence properties [22]. In recent years, researchers have proposed many methods to solve the worldwide problem of poor stability of metal halide perovskite, such as surface modification, polymer encapsulation, silicon coating, atomic layer deposition passivation, etc. Imran et al. [23] reported the synthesis of asymmetric CsPbX3–Pb4S3Br2 NCs heterojunction by two-step direct synthesis using preformed subnanometer CsPbBr3 clusters. The existence of an epitaxial interface of Pb4S3Br2 material increases the structural rigidity of the perovskite lattice and improves the stability of perovskite material. Although surface passivation by organic ligands or polymers can significantly improve the stability of perovskites, these insulating shells severely limit the charge transport of metal halide perovskite themselves for optoelectronic, catalytic, and other applications.
In addition to polymers [24,25,26,27,28,29,30,31] and inert shell layers [16,32,33,34,35,36], semiconductors were also introduced to improve the stability and photophysical properties of perovskite. More importantly, by using another semiconductor with different conduction and valence bands, one can easily tune the bandgap or charge transfer processes and thus improve photophysical properties. Prior to shell growth of perovskite NCs, shell materials were carefully selected to design the core–shell structure to allow for favorable band arrangement. In type I core–shell structures, the shell material has a wider energy band gap than the core, so the edges of the core material of the conduction and valence bands are located in the energy gap of the shell, having a crossing band arrangement. The band gap of the core material determines the band gap energy of this heterogeneous structure, while the shell provides passivation of surface defects and limits electrons and holes within the core. This can improve the overall optical properties and stability of NCs [37]. Type II structures have a staggered arrangement in which the edges of the valence and conduction bands of the core material have lower or higher energies than the edges of the shell material. The band gap of type II structures is determined by the energy separation between the conduction band edge of one semiconductor and the valence band edge of another semiconductor. Carriers are spatially separated in these structures, thus offering potential benefits for applications, such as optoelectronics and photocatalysis. TiO2, as a chemically inert and nontoxic photocatalyst, is the most important material used in seawater desalination and wastewater treatment [38,39,40]. Due to its rapid electron–hole recombination, forming a type II heterojunction between CsPbBr3 and titanium dioxide can not only solve the problem of stability but also improve the charge transport efficiency of perovskite, laying the foundation for its photoelectric and photocatalytic applications.
The sol–gel method has been widely used for the modification of perovskite nanoparticles with various oxide materials, such as silica and titanium dioxide. However, since CsPbBr3 NCs are very sensitive to water or temperature and fluorescence of perovskites quenching occurs before the formation of silica or titanium oxide shell layers, it is challenging to modify their surfaces using conventional sol–gel methods. In the available reports, Li et al. [41] demonstrated that TiO2 shell coating improved the stability of perovskite but resulted in partial decomposition of CsPbBr3 NCs or the dissolution of corners and edges during hydrolysis and calcination. Ji et al. [42] studied the synthesis of highly stable CsPbBr3/TiO2 nanocomposites by a top-assisted low-temperature solvothermal method. Although the material improves the stability of the perovskite NCs and preserves the edge of the perovskite core from being destroyed, the agglomeration is severe and the size is large, which poses obstacles for the photoelectric and photocatalytic applications. For example, in perovskite light-emitting diode (PeLED) applications, small-size particles facilitate the formation of thin films with high uniformity. Therefore, we are eager to develop a method that can achieve heterogeneous coating of CsPbBr3 with TiO2 at the single-particle level to improve stability.
Herein, we certify the synthesis of an asymmetric CsPbBr3/TiO2 core–shell nanocrystals heterojunction at the single-particle level in anhydrous solvent and at low temperature, aiming to improve the stability of perovskite, thereby improving its photoelectric and photocatalytic performance. Based on the band alignment, it is confirmed that the obtained CsPbBr3/TiO2 heterojunction NCs has a type II heterostructure with enhanced carrier transport efficiency. It is demonstrated that the CsPbBr3/TiO2 NCs show good optoelectronic properties and enhanced water stability under water test conditions, which provides a reference value for the preparation of high-efficiency, long-life photovoltaic and luminescent devices.

2. Results and Discussion

Firstly, cesium precursor (cesium oleate) was prepared by using cesium carbonate as cesium source, adding oleic acid and trin-octylphosphine ligand as raw materials. Then, lead bromide precursor was prepared by adding octadecene, oleic acid, and oleylamine to dissolve lead bromide powder. Cesium oleate was injected into the lead bromide precursor by hot-injection method. After a few seconds of reaction, the reaction solution was cooled to room temperature in an ice bath and centrifugally washed to obtain CsPbBr3 NCs. Next, we chose titanium butoxide as the titanium precursor, slowly dropped the diluted titanium precursor into the toluene solution of CsPbBr3 NCs at a certain speed, and stirred the solution at room temperature for 3 h. The hydrolysis reaction occurred, so that TiO2 was deposited on the surface of CsPbBr3 NCs, and then the reaction solution was placed in a Teflon-lined stainless-steel autoclave for 10 h. Finally, the CsPbBr3/TiO2 light yellow powder was obtained by centrifugation and drying (see Section 3 for details).
The morphologies of CsPbBr3 NCs and asymmetrical CsPbBr3/TiO2 NCs were analyzed by transmission electron microscopy (TEM). As shown in Figure 1A, the CsPbBr3 NCs materials exhibit highly uniform monodispersion with clear edges and uniform size, with an edge length of about 17 nm, indicating the preparation of high-quality perovskite nanocrystals. Figure 1B is the TEM image of asymmetrical CsPbBr3/TiO2 core–shell NCs. It can be obviously observed that each nanocrystalline particle has two regions with different image contrast, indicating that the material is composite structure. CsPbBr3/TiO2 NCs are composed of two different structural components, CsPbBr3 and TiO2, with an obvious heterogeneous interface, in which the dark part is CsPbBr3 NCs, and the light part is TiO2, which further indicates the formation of core–shell heterojunction. Importantly, after the titanium oxide coating, the edges of the perovskite are not dissolved, achieving precise coating at the single-particle level, as well as the distribution of these elements.
The crystal structures of CsPbBr3 NCs and CsPbBr3/TiO2 NCs were characterized using an X-ray diffractometer. As shown in Figure 2 for the X-ray diffraction analysis (XRD) of CsPbBr3 NCs and CsPbBr3/TiO2 NCs, the XRD peaks of CsPbBr3 NCs appeared at 2θ = 15.2°, 21.6°, 24.1°, 26.3°, 30.5°, 30.8°, 34.5°, 38.0°, 43.9°, 46.7°, and 54.4°, corresponding to (001), (010), (110), ( 2 ¯ 02), (210), (002), ( 1 ¯ 12), (102), (020), (021), and (401) diffraction planes, respectively. The diffraction peaks of CsPbBr3 NCs match well with the standard monoclinic perovskite phase (PDF#00-054-0751) of CsPbBr3 NCs, with splitting diffraction peaks at 2θ ≈ 30°. For CsPbBr3/TiO2 NCs, in addition to the XRD peak observed for the monoclinic phase of CsPbBr3 NCs, an additional weak diffraction peak (marked by blue circles in the Figure 2) can be observed at 2θ ≈ 25°, which corresponds to the (101) diffraction plane of anatase TiO2 (PDF#00-021-1272). The above analysis indicates the formation of anatase type TiO2 on the surface of CsPbBr3 NCs. The lower intensity of the XRD peak of titanium dioxide relative to the CsPbBr3 NCs diffraction peak indicates that the titanium dioxide is partially crystallized during the hydrothermal treatment and has weak crystallinity.
Since most applications of perovskite are related to its photoelectric and catalytic properties, it is desirable that materials have both environmental stability and efficient charge transport. To gain more insight into the material charge transfer kinetics, we recorded the photoluminescence (PL) emission spectra and time-resolved photoluminescence spectra (TRPL) of CsPbBr3 NCs and asymmetrical CsPbBr3/TiO2 NCs (Figure 3). As shown in Figure 3A, the photoluminescence emission peaks of CsPbBr3 NCs and CsPbBr3/TiO2 NCs at about 522 nm and 528 nm, respectively, can be observed. Compared with CsPbBr3 NCs, the PL peak of CsPbBr3/TiO2 NCs is slightly red-shifted by 6 nm. After TiO2 shell coating, the PL intensity decreases. Figure 3B shows the TRPL of CsPbBr3 NCs and asymmetrical CsPbBr3/TiO2 NCs. The fitted decay curves reveal that the CsPbBr3/TiO2 NCs (τ = 1.73 ns) have shorter carrier lifetime compared with CsPbBr3 (τ = 6.08 ns), and the decreased PL lifetime indicates improved charge diffusion and separation, i.e., the rapid charge transfer from CsPbBr3 NCs to the TiO2 shell, thus leading to enhanced photovoltaic performance. The reduced fluorescence lifetime further manifests that after TiO2 shell coating, the CsPbBr3/TiO2 core/shell NCs generate a new nonradiative path, probably an electron transfer from the conduction band of CsPbBr3 NCs to the conduction band of titanium dioxide, indicating the formation of type II heterojunctions, which greatly facilitates the separation of electron–hole pairs and CsPbBr3/TiO2 NCs with higher carrier mobility [43].
To confirm our hypothesis, we investigated the electronic energy band structures of CsPbBr3 NCs and asymmetrical CsPbBr3/TiO2 NCs using ultraviolet–visible absorption spectroscopy (UV–Vis). Compared with the normal hydrogen electrode (NHE), the conduction band and valence band edge of CsPbBr3 NCs are ≈ −1.07 eV and ≈1.35 eV, respectively. The conduction and valence bands of CsPbBr3 are higher than those of titanium dioxide. Therefore, a type II band arrangement can be formed in the CsPbBr3/TiO2 heterostructure NCs. In this case, the lowest energy hole should be confined to the CsPbBr3 nucleus, and electrons can delocalize on the CsPbBr3 nucleus and the titanium dioxide shell, thus promoting charge transfer and reducing PLQY. From the UV–Vis absorption spectra (Figure 4A), it is clear that CsPbBr3 NCs and CsPbBr3/TiO2 NCs showed significant absorption peaks in ultraviolet range. Based on Einstein’s photoelectric effect and the Kubelka–Munk equation [44], the surface optical band gaps of CsPbBr3 NCs and CsPbBr3/TiO2 NCs were estimated to be 2.42 eV and 2.40 eV(Figure 4B), respectively, by extrapolating the linear region of the absorption edge in the UV–Vis spectrum to the energy axis intercept. The reduction in the electron transition band gap is consistent with the fluorescence lifetime results, which further proves that the charge transport characteristics of perovskite nanocrystals are optimized after titanium oxide coating. The work function (Φ) is another important parameter in the study of electron transfer in duplicate semiconductor heterojunctions. It can be calculated from the energy difference of the surface valence band maximum and the Fermi level of the material. According to the above data, the work functions of anatase TiO2 and CsPbBr3 NCs are calculated to be 6.0 eV and 4.58 eV, respectively, indicating that the Fermi level of anatase TiO2 is lower than that of CsPbBr3 NCs. When they come into contact, electrons flow from CsPbBr3 to anatase titanium dioxide, making the same Fermi level consistent and producing internal electric field at the CsPbBr3/TiO2 NCs interface. These results are consistent with the results of the fluorescence lifetime test. The coating of titanium oxide is beneficial to charge separation in perovskite and improves the catalytic activity.
Based on the energy levels of CsPbBr3 NCs and TiO2 NCs, we can obtain the band structure of the sample. Obviously, it was found that a type II nanocomposite was formed, in which both the conduction band and valence band levels of the semiconductor CsPbBr3 were higher than those of the semiconductor titanium dioxide. Once irradiated by light (λ = 546.5 nm), electrons in the valence band of CsPbBr3 leap to the conduction band and form holes in their original positions. Under the action of the internal electric field, the carriers generated and leapt to the depletion region will be separated rapidly in the heterojunction, where the electrons will move toward the TiO2 conduction band and the holes will transport to CsPbBr3, and the photocurrent output due to the photoresponse can be obtained through the electrochemical workstation. The holes from the TiO2 valence band will be transferred to the CsPbBr3 valence band under the action of the external electric field, and the electrons will migrate in the opposite direction, thus producing a stable photoresponse, and the optical signal will be converted into an electrical signal.
The electrochemical impedance spectra (ESI) and the photoresponsive current densities of CsPbBr3 NCs and CsPbBr3/TiO2 NCs materials were tested as shown Figure 5. The charge transport properties of CsPbBr3 NCs and CsPbBr3/TiO2 NCs were further investigated by photoelectrochemical studies. The Nyquist plot is generally used in EIS and the Nyquist plot of the most common equivalent circuit model is semicircular in shape. In the equivalent circuit model, Cpe represents the double-layer capacitance, Rs represents the solution resistance from the reference electrode to the working electrode, and Rc is the charge transfer resistance of the electrode. The diameter of the semicircle is equal to the carrier migration resistance, and the smaller the diameter of the semicircle, the larger the carrier migration rate. As shown in Figure 5A, the curvature of the impedance spectrum of CsPbBr3/TiO2 NCs is significantly greater than that of CsPbBr3 NCs. Additionally, because the curvature of the circle is inversely proportional to the radius, the CsPbBr3/TiO2 NCs has a smaller semicircle arc compared with CsPbBr3 NCs. In other words, CsPbBr3/TiO2 NCs have a smaller carrier migration resistance than CsPbBr3 NCs. To sum up, with the formation of TiO2 shells, the charge transfer resistance of the CsPbBr3/TiO2 NCs is significantly reduced. In order to record the photocurrent generated by the photoresponse and exclude the effect of other factors, the light source is usually switched on at certain intervals to record the photocurrent profile. As shown in Figure 5B, when the light source was turned off, the recorded photocurrent density was almost zero; when the light source was turned on after 30 s, the current instantaneously increased; and when the light source was turned off at an interval of 30 s, the current instantaneously decreased. Compared with CsPbBr3 NCs, CsPbBr3/TiO2 NCs have higher current density. This result further confirms that the formation of heterostructure between CsPbBr3 and TiO2 improves the carrier migration efficiency of the perovskite material and has better charge transport properties.

3. Materials and Methods

3.1. Materials

The cesium carbonate (Cs2CO3, 99.9% trace metals basis, Aladdin), oleic acid (OA, 80–90%, Aladdin), oleylamine (OAm, 80–90%, Aladdin), titanium butoxide (C16H36O4Ti, TBOT, >99.0%, Aladdin), trinoctylphosphine (TOP, 90%, Aladdin), 1-octadecene (ODE, >90% (GC), Aladdin), lead (II) bromide (PbBr2, 99.0%, Aladdin), and toluene were purchased from Sinopharm Chemical Reagent Co. All reagents labeled Aladdin are from Aladdin Reagent (Shanghai) Co., Ltd., Shanghai, China.

3.2. Methods

Preparation of Cesium Oleate: In a typical synthesis, 400 mg of Cs2 CO3 (1.23 mmol), 15 mL ODE, 2 mL TOP, and 1.25 mL oleic acid were added into a 25 mL three-neck round bottom flask connected with a double row of tubes and dried under N2 at 120 °C for 1 h. Cesium carbonate powder completely dissolved, and the mixed solution became transparent, indicating that cesium carbonate and oleic acid completely reacted to produce cesium oleate (Cs-OA) solution. According to the purity of cesium carbonate, the purity of cesium oleate can be calculated to be 7.55%. The Cs-OA solution was stored at room temperature and preheated to 140 °C before synthesis of CsPbBr3 nanocrystals.
Synthesis and Purification of CsPbBr3 NCs: Briefly, 138 mg of PbBr2 (0.374 mmol) was mixed with 10.0 mL ODE in another 25 mL three-neck round bottom flask connected with a double row of tubes and was dried under vacuum at 120 °C for 1 h. Then, 1.0 mL OAm and 1.0 mL OA were slowly added into the flask under the N2 atmosphere. After the solution became clear, the temperature was raised to 170 °C, followed rapidly injection of 0.8 mL of Cs-OA. After 5 s, the solution turned yellow, and the reaction was quenched by immersing the flask into an ice-water bath. The quenched solution was centrifuged at 10,000× g rpm for 5 min. To remove supernatant, the centrifuged CsPbBr3 NCs were dispersed in 5 mL dry toluene, were washed twice with dry toluene, and dispersed into 25 mL of dry toluene for further use.
Preparation of asymmetrical CsPbBr3/TiO2 heterojunction: Typically, when the stirring rate of the reaction system is 1500 rpm, titanium butoxide solution (80 µL titanium butoxide/2 mL dry toluene) is added to a 10 mL CsPbBr3 NCs toluene solution. After stirring for 3 h, the reaction solution was transferred into a Teflon-lined stainless-steel autoclave, and then the solvothermal reaction process was carried out at 160 °C for 10 h. After cooling down to room temperature, the obtained product was centrifuged at 10,000× g rpm for 5 min to collect precipitates, washed twice with dry toluene, and dried at 40 °C for 6 h; then, it was ground to obtain an asymmetrical CsPbBr3/TiO2 powder sample.

3.3. Characterization

3.3.1. Materials Characterization

Transmission Electron Microscopy (TEM) characterization was used. TEM images were acquired on a JEOL JEM-1400 Plus electron microscope equipped with a thermionic gun at an accelerating voltage of 120 kV. The samples of CsPbBr3 NCs and CsPbBr3/TiO2 NCs were prepared by depositing a diluted nanocrystal suspension in toluene onto carbon-coated copper grids.

3.3.2. Optical Characterization

UV–Vis absorption spectra of CsPbBr3 NCs and CsPbBr3/TiO2 NCs samples was collected by Shimadzu UV-3600. The steady-state photoluminescence (PL) spectra and steady-state and time-resolved photoluminescence (TRPL) spectra were measured on a Shimadzu RF-6000 spectrophotometer with LabSolutions RF software using an excitation wavelength (λex) of 420 nm for CsPbBr3 NCs and the asymmetrical CsPbBr3/TiO2 NCs. Nanocrystals samples were prepared by diluting NC solutions in toluene in quartz cuvettes with a path length of 10 mm.

3.3.3. (Photo)electrochemistry Characterization

For (photo)electrochemistry measurement, a Zennium electrochemical workstation (Germany, Zahner Company) was used, and the measurement was performed in a three-electrode setup with the working electrode of the sample electrode, counter electrode of platinum disk, and reference electrode of Ag/AgCl (saturated KCl). (EAg/AgCl = +0.1989V vs. NHE) in the (photo)electrochemical electrolyte (0.5 M Na2SO4). For photocurrent measurement, the light source was a 405 LED, the light intensity was tested with a Newport photometer. Electrochemical impedance spectra were measured in 0.5 M Na2SO4 at −0.1 V vs. NHE had an amplitude of 10 mV (Frequency: 100 mHz–20 kHz) [33].

4. Conclusions

In summary, CsPbBr3 NCs were firstly prepared by thermal injection liquid-phase synthesis, and then asymmetrical CsPbBr3/TiO2 core–shell NCs were synthesized by sol-gel method. The prepared CsPbBr3 NCs were uniformly distributed, homogeneous in size, and highly monodisperse, with a size of around 17 nm. The TEM images of asymmetrical CsPbBr3/TiO2 core–shell NC materials clearly show that the CsPbBr3 and TiO2 materials are distributed in different regions with different image contrast, indicating the formation of heterogeneous structures. The red shift of 6 nm in the photoluminescence emission peak and the decrease in the photoluminescence emission peak and fluorescence lifetime are due to the difference in grain size and charge migration between CsPbBr3 and TiO2. The work function calculation shows that the Fermi level of anatase TiO2 is lower than that of CsPbBr3 NCs. When they touch each other, electrons flow from CsPbBr3 to anatase titanium dioxide, making the same Fermi level consistent. The energy band structures of the materials were obtained by UV–Vis analysis, demonstrating that asymmetrical CsPbBr3/TiO2 NCs form a type II heterostructure with narrower forbidden band widths and shorter carrier migration paths, as well as higher carrier mobilities. By measuring the electrochemical impedance and photocurrent density of CsPbBr3 NCs and asymmetrical CsPbBr3/TiO2 NCs, it is further verified that asymmetrical CsPbBr3/TiO2 NCs have better carrier transport properties compared with CsPbBr3 NCs. This work provides a low-cost and easy-to-handle method to achieve good water stability and excellent optoelectronic properties of perovskite, which shows promising applications in photocatalysts and high-performance optoelectronic devices.

Author Contributions

Y.L. (Yong Liu) conceived project ideas and material synthetic designs. W.L., J.L., X.W., Y.L. (Yuqing Li) and J.H. performed material characterization and data analysis. Y.L. (Yong Liu) and W.L. discussed the results and contents for this work. W.L. and Y.L. (Yong Liu) wrote the original drafts, reviews, and edits. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China (52072281, Yong Liu); The Major Program of the National Natural Science Foundation of China (22293021, Yong Liu); the National innovation and entrepreneurship training program for college students (No. S202210497011 Yong Liu). Yong Liu gratefully acknowledges Youth Innovation Research Fund project of State Key Laboratory of Advanced Technology for Materials Synthesis and Processing, Wuhan University of Technology.

Data Availability Statement

Data are available in a publicly accessible repository that does not issue DOIs, or on request from the corresponding author.

Acknowledgments

The authors thank the Analytical and Testing Centre of Wuhan University of Technology for the UV–Vis characterizations.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kojima, A.; Teshima, K.; Shirai, Y.; Miyasaka, T. Organometal halide perovskites as visible-light sensitizers for photovoltaic cells. J. Am. Chem. Soc. 2009, 131, 6050–6051. [Google Scholar] [CrossRef]
  2. Jeon, N.J.; Noh, J.H.; Yang, W.S.; Kim, Y.C.; Ryu, S.; Seo, J.; Seok, S. Compositional engineering of perovskite materials for high-performance solar cells. Nature 2015, 517, 476–480. [Google Scholar] [CrossRef] [PubMed]
  3. Swarnkar, A.; Marshall, A.R.; Sanehira, E.M.; Chernomordik, B.D.; Moore, D.T.; Christians, J.A.; Chakrabarti, T.; Luther, J. Quantum dot-induced phase stabilization of α-CsPbI3 perovskite for high-efficiency photovoltaics. Science 2016, 354, 92–95. [Google Scholar] [CrossRef] [Green Version]
  4. Yang, W.S.; Park, B.W.; Jung, E.H.; Jeon, N.J.; Kim, Y.C.; Lee, D.U.; Shin, S.S.; Seo, J.; Kim, E.K.; Noh, J.H. Iodide management in formamidinium-lead-halide–based perovskite layers for efficient solar cells. Science 2017, 356, 1376–1379. [Google Scholar] [CrossRef] [Green Version]
  5. Song, J.; Li, J.; Li, X.; Xu, L.; Dong, Y.; Zeng, H. Quantum dot light-emitting diodes based on inorganic perovskite cesium lead halides (CsPbX3). Adv. Mater. 2016, 27, 7162–7167. [Google Scholar] [CrossRef] [PubMed]
  6. Li, J.; Xu, L.; Wang, T.; Song, J.; Chen, J.; Xue, J.; Dong, Y.; Cai, B.; Shan, Q.; Han, B. 50-fold EQE improvement up to 6.27% of solution-processed all-inorganic perovskite CsPbBr3 QLEDs via surface ligand density control. Adv. Mater. 2017, 29, 1603885. [Google Scholar] [CrossRef]
  7. Zhi-Kuang, T.; Reza Saberi, M.; May Ling, L.; Pablo, D.; Ruben, H.; Felix, D.; Michael, P.; Aditya, S.; Pazos, L.M.; Dan, C. Bright light-emitting diodes based on organometal halide perovskite. Nat. Nanotechnol. 2018, 9, 687–692. [Google Scholar]
  8. Zhang, X.; Xu, B.; Zhang, J.; Gao, Y.; Zheng, Y.; Wang, K.; Sun, X. All-inorganic perovskite nanocrystals for high-efficiency light emitting diodes: Dual-phase CsPbBr3-CsPb2Br5 Composites. Adv. Funct. Mater. 2016, 26, 4595–4600. [Google Scholar] [CrossRef]
  9. Song, T.; Cheng, H.; Fu, C.; He, B.; Zou, B.J.N. Influence of the active layer nanomorphology on device performance for ternary PbSxSe1−X quantum dots based solution-processed infrared photodetector. Nanotechnology 2016, 27, 165202. [Google Scholar] [CrossRef]
  10. Chiba, T.; Hoshi, K.; Pu, Y.J.; Takeda, Y.; Kido, J. High efficiency perovskite quantum-dot light-emitting devices by effective washing process and interfacial energy level alignment. ACS Appl. Mater. Inter. 2017, 9, 18054. [Google Scholar] [CrossRef] [PubMed]
  11. Dong, Y.; Gu, Y.; Zou, Y.; Song, J.; Xu, L.; Li, J.; Xue, F.; Li, X.; Zeng, H. Improving all-Inorganic perovskite photodetectors by preferred orientation and plasmonic effect. Small 2016, 12, 5622–5632. [Google Scholar] [CrossRef] [PubMed]
  12. Veldhuis, S.A.; Boix, P.P.; Yantara, N.; Li, M.; Sum, T.C.; Mathews, N.; Mhaisalkar, S.G. Perovskite materials for light-emitting diodes and lasers. Adv. Mater. 2016, 28, 6804–6834. [Google Scholar] [CrossRef] [PubMed]
  13. Fu, Y.; Zhu, H.; Schrader, A.W.; Liang, D.; Jin, S. Nanowire lasers of formamidinium lead halide perovskites and their stabilized alloys with improved stability. Nano Lett. 2016, 16, 1000–1008. [Google Scholar] [CrossRef] [Green Version]
  14. Zhu, H.; Fu, Y.; Meng, F.; Wu, X.; Gong, Z.; Ding, Q.; Gustafsson, M.V.; Trinh, M.T.; Jin, S.; Zhu, X. Lead halide perovskite nanowire lasers with low lasing thresholds and high quality factors. Nat. Mater. 2015, 14, 636–642. [Google Scholar] [CrossRef] [PubMed]
  15. Xing, G.; Mathews, N.; Lim, S.S.; Yantara, N.; Liu, X.; Sabba, D. Low-temperature solution-processed wavelength-tunable perovskites for lasing. Nat.Mater. 2014, 13, 476–480. [Google Scholar] [CrossRef] [PubMed]
  16. Zhong, Q.; Cao, M.; Hu, H.; Di, Y.; Chen, M.; Li, P.; Wu, L.; Qiao, Z. One-pot synthesis of highly stable CsPbBr3@SiO2 core-shell nanoparticles. ACS Nano 2018, 12, 8579–8587. [Google Scholar] [CrossRef] [PubMed]
  17. Protesescu, L.; Yakunin, S.; Bodnarchuk, M.I. Nanocrystals of cesium lead halide perovskites (CsPbX3, X = Cl, Br, and I): Novel optoelectronic materials showing bright emission with wide color gamut. Nano.Lett. 2015, 15, 3692–3696. [Google Scholar] [CrossRef] [Green Version]
  18. Xu, Y.F.; Yang, M.Z.; Chen, H.Y.; Liao, J.F.; Kuang, D. Enhanced solar-driven gaseous CO2 conversion by CsPbBr3 nanocrystal/Pd nanosheet Schottky-junction photocatalyst. ACS Appl. Energy Mater. 2018, 1, 5083–5089. [Google Scholar] [CrossRef]
  19. Nedelcu, G.; Protesescu, L.; Yakunin, S.; Bodnarchuk, M.I.; Grotevent, M.J.; Kovalenko, M.V. Fast anion-exchange in highly luminescent nanocrystals of cesium lead halide perovskites (CsPbX3, X = Cl, Br, I). Nano Lett. 2015, 15, 5635–5640. [Google Scholar] [CrossRef]
  20. Chen, S.; Yin, H.; Liu, P.; Wang, Y.; Zhao, H. Stabilization and performance enhancement strategies for halide perovskite photocatalysts. Adv. Mater. 2023, 35, 2203836. [Google Scholar] [CrossRef]
  21. Xiao, M.; Hao, M.; Lyu, M.; Moore, E.G.; Zhang, C.; Luo, B.; Hou, J.; Lipton-Duffin, J.; Wang, L. Surface ligands stabilized lead halide perovskite quantum dot photocatalyst for visible light-driven hydrogen generation. Adv. Funct. Mater. 2019, 29, 1905683. [Google Scholar] [CrossRef]
  22. Liu, C.; Zeng, Q.; Wei, H.; Yu, Y.; Zhao, Y.; Feng, T.; Yang, B. Metal halide perovskite nanocrystal solar cells: Progress and challenges. Small Methods 2020, 4, 2000419. [Google Scholar] [CrossRef]
  23. Imran, M.; Peng, L.; Pianetti, A.; Pinchetti, V.; Ramade, J.; Zito, J.; Di Stasio, F.; Buha, J.; Toso, S.; Song, J.; et al. Halide perovskite-lead chalcohalide nanocrystal heterostructures. J. Am. Chem. Soc. 2021, 143, 1435–1446. [Google Scholar] [CrossRef]
  24. Zhang, H.; Wang, X.; Liao, Q. Embedding perovskite nanocrystals into a polymer matrix for tunable luminescence probes in cell imaging. Adv. Funct. Mater. 2017, 27, 1604382. [Google Scholar] [CrossRef]
  25. Wei, Y.; Deng, X.; Xie, Z.; Cai, X.; Liang, S.; Ma, P.A.; Hou, Z.Y.; Cheng, Z.Y.; Lin, J. Enhancing the stability of perovskite quantum dots by encapsulation in crosslinked polystyrene beads via a swelling–shrinking strategy toward superior water resistance. Adv. Funct. Mater. 2017, 27, 1703535. [Google Scholar] [CrossRef]
  26. Wang, Y.; Zhu, Y.; Huang, J.; Cai, J.; Zhu, J.; Yang, X.; Shen, J.; Jiang, H.; Li, C. CsPbBr3 perovskite quantum dots-bsed monolithic electrospun fiber membrane as an ultrastable and utrasensitive fluorescent sensor in aqueous medium. J. Phys. Chem. Lett. 2016, 7, 4253–4258. [Google Scholar] [CrossRef]
  27. Raja, S.N.; Bekenstein, Y.; Koc, M.A.; Fischer, S.; Zhang, D.; Lin, L.; Ritchie, R.O.; Yang, P.; Alivisatos, A.P. Interfaces, Encapsulation of perovskite nanocrystals into macroscale polymer matrices: Enhanced stability and polarization. ACS Appl. Mater. Inter. 2016, 8, 35523–35533. [Google Scholar] [CrossRef] [Green Version]
  28. Hou, S.; Guo, Y.; Tang, Y.; Quan, Q. Interfaces, synthesis and stabilization of colloidal perovskite nanocrystals by multidentate polymer micelles. ACS Appl. Mater. Inter. 2017, 9, 18417. [Google Scholar] [CrossRef]
  29. Liao, H.; Guo, S.; Cao, S.; Wang, L.; Gao, F.; Yang, Z.; Zheng, J.; Yang, W. A general strategy for in situ growth of all-Inorganic CsPbX3 (X = Br, I, and Cl) perovskite nanocrystals in polymer fibers toward significantly enhanced water/thermal stabilities. Adv. Opt. Mater. 2018, 6, 1800346. [Google Scholar] [CrossRef]
  30. Yang, X.; Xu, T.; Zhu, Y.; Cai, J.; Gu, K.; Zhu, J.; Wang, Y.; Shen, J.; Li, C. Preparation of CsPbBr3@PS composite microspheres with high stability by electrospraying. J. Mater. Chem. C 2018, 6, 7971–7975. [Google Scholar] [CrossRef]
  31. Meyns, M.; Perálvarez, M.; Heuer-Jungemann, A.; Hertog, W.; Ibáñez, M.; Nafria, R.; Genç, A.; Arbiol, J.; Kovalenko, M.V. Carreras. Polymer-enhanced stability of inorganic perovskite nanocrystals and their application in color conversion LEDs. ACS Appl. Mat. Inte. 2016, 8, 19579–19586. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Li, Z.; Long, K.; Huang, S.; Liang, L. Highly luminescent and ultrastable CsPbBr3 perovskite quantum dots incorporated into a silica/alumina monolith. Angew. Chem. 2017, 129, 8246–8250. [Google Scholar] [CrossRef]
  33. Wang, H.C.; Lin, S.Y.; Tang, A.C.; Singh, B.P.; Tong, H.C.; Chen, C.Y.; Lee, Y.C.; Tsai, T.L.; Liu, S. Mesoporous silica particles integrated with all-Inorganic CsPbBr3 perovskite quantum-dot nanocomposites (MP-PQDs) with high stability and wide color gamut used for backlight display. Angew. Chem. Int. 2016, 55, 7924–7929. [Google Scholar] [CrossRef]
  34. Dirin, D.N.; Protesescu, L.; Trummer, D.; Kochetygov, I.V.; Yakunin, S.; Krumeich, F.; Stadie, N.P.; Kovalenko, M.V. Harnessing defect-tolerance at the nanoscale: Highly luminescent lead halide perovskite nanocrystals in mesoporous silica matrixes. Nano Lett. 2016, 16, 5866–5874. [Google Scholar] [CrossRef] [Green Version]
  35. Sun, C.; Zhang, Y.; Ruan, C.; Yin, C.; Wang, X.; Wang, Y.; Yu, W. Efficient and stable white LEDs with silica-coated inorganic perovskite quantum dots. Adv. Mater. 2016, 28, 10088–10094. [Google Scholar] [CrossRef] [PubMed]
  36. Loiudice, A.; Saris, S.; Oveisi, E.; Alexander, D.; Buonsanti, R. CsPbBr3 QD/AlOx inorganic nanocomposites with exceptional stability in water, light, and heat. Angew. Chem. 2017, 129, 11099. [Google Scholar] [CrossRef] [Green Version]
  37. Low, J.; Yu, J.; Jaroniec, M.; Wageh, S.; Al-Ghamdi, A. Heterojunction Photocatalysts. Adv. Mater. 2017, 29, 1601694. [Google Scholar] [CrossRef]
  38. Tobaldi, D.M.; Dvoranova, D.; Lajaunie, L.; Rozman, N.; Figueiredo, B.; Seabra, M.P.; Skapin, A.S.; Calvino, J.J.; Brezova, V.; Labrincha, J.A. Graphene-TiO2 hybrids for photocatalytic aided removal of VOCs and nitrogen oxides from outdoor environment. Chem. Eng. J. 2021, 405, 126651. [Google Scholar] [CrossRef]
  39. Jamjoum, H.A.A.; Umar, K.; Adnan, R.; Razali, M.R.; Mohamad Ibrahim, M.N. Synthesis, characterization, and photocatalytic activities of graphene oxide/metal oides nanocomposites: A review. Front. Chem. 2021, 9, 752276. [Google Scholar] [CrossRef]
  40. Hunge, Y.M.; Yadav, A.A.; Khan, S.; Takagi, K.; Suzuki, N.; Teshima, K.; Terashima, C.; Fujishima, A. Photocatalytic degradation of bisphenol A using titanium dioxide@nanodiamond composites under UV light illumination. Colloid Interface Sci. 2021, 582, 1058–1066. [Google Scholar] [CrossRef]
  41. Li, Z.Z.; Hofman, E.; Li, J.; Davis, A. Photoelectrochemically active and environmentally stable CsPbBr3/TiO2 core/shell nanocrystals. Adv. Funct. Mater. 2017, 28, 1704288. [Google Scholar] [CrossRef]
  42. Ji, Y.; Wang, M.; Yang, Z.; Qiu, H.; Bhatti, A. Trioctylphosphine-assisted pre-protection low-temperature solvothermal synthesis of highly stable CsPbBr3/TiO2 Nanocomposites. J. Phys. Chem. Lett. 2021, 12, 3786–3794. [Google Scholar] [CrossRef] [PubMed]
  43. Long, R.; Fang, W.H.; Prezhd, O.V. Strong interaction at the perovskite/TiO2 interface facilitates ultrafast photoinduced charge separation: A nonadiabatic molecular dynamics sudy. J. Phys. Chem. C 2017, 121, 3797–3806. [Google Scholar] [CrossRef]
  44. Zhang, Y.; Wang, X.; Chen, Y.; Gao, Y. Improved electroluminescence performance of quantum dot light-emitting diodes: A promising hole injection layer of Fe-doped NiO nanocrystals. Opt. Mater. 2020, 107, 110158. [Google Scholar] [CrossRef]
Figure 1. (A,B) Transmission electron microscopy (TEM) images of CsPbBr3 NCs and asymmetrical CsPbBr3/TiO2 NCs, respectively.
Figure 1. (A,B) Transmission electron microscopy (TEM) images of CsPbBr3 NCs and asymmetrical CsPbBr3/TiO2 NCs, respectively.
Catalysts 13 01048 g001
Figure 2. X-ray diffraction (XRD) patterns of CsPbBr3 NCs and asymmetrical CsPbBr3/TiO2 NCs.
Figure 2. X-ray diffraction (XRD) patterns of CsPbBr3 NCs and asymmetrical CsPbBr3/TiO2 NCs.
Catalysts 13 01048 g002
Figure 3. (A) Photoluminescence (PL) spectra and (B) PL time-resolved photoluminescence (TRPL) spectra of CsPbBr3 NCs and asymmetrical CsPbBr3/TiO2 NCs.
Figure 3. (A) Photoluminescence (PL) spectra and (B) PL time-resolved photoluminescence (TRPL) spectra of CsPbBr3 NCs and asymmetrical CsPbBr3/TiO2 NCs.
Catalysts 13 01048 g003
Figure 4. (A) Ultraviolet–visible spectroscopy (UV–Vis) of CsPbBr3 NCs and asymmetrical CsPbBr3/TiO2 NCs. (B) Absorbance versus photon energy and the determined bandgap Eg of CsPbBr3 NCs and asymmetrical CsPbBr3/TiO2 NCs.
Figure 4. (A) Ultraviolet–visible spectroscopy (UV–Vis) of CsPbBr3 NCs and asymmetrical CsPbBr3/TiO2 NCs. (B) Absorbance versus photon energy and the determined bandgap Eg of CsPbBr3 NCs and asymmetrical CsPbBr3/TiO2 NCs.
Catalysts 13 01048 g004
Figure 5. (A) Electrochemical impedance spectra (Nyquist plot) of CsPbBr3 NCs and CsPbBr3/TiO2 NCs: inset is the equivalent circuit model and Z′ and −Z″ are the imaginary part and real part impedances, respectively. (B) Transient photocurrent responses to on–off illumination of CsPbBr3 NCs and CsPbBr3/TiO2 NCs electrodes at −0.1 V versus NHE in neutral water (0.5 M Na2SO4).
Figure 5. (A) Electrochemical impedance spectra (Nyquist plot) of CsPbBr3 NCs and CsPbBr3/TiO2 NCs: inset is the equivalent circuit model and Z′ and −Z″ are the imaginary part and real part impedances, respectively. (B) Transient photocurrent responses to on–off illumination of CsPbBr3 NCs and CsPbBr3/TiO2 NCs electrodes at −0.1 V versus NHE in neutral water (0.5 M Na2SO4).
Catalysts 13 01048 g005
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

Liu, W.; Liu, J.; Wang, X.; He, J.; Li, Y.; Liu, Y. Synthesis of Asymmetrical CsPbBr3/TiO2 Nanocrystals with Enhanced Stability and Photocatalytic Properties. Catalysts 2023, 13, 1048. https://doi.org/10.3390/catal13071048

AMA Style

Liu W, Liu J, Wang X, He J, Li Y, Liu Y. Synthesis of Asymmetrical CsPbBr3/TiO2 Nanocrystals with Enhanced Stability and Photocatalytic Properties. Catalysts. 2023; 13(7):1048. https://doi.org/10.3390/catal13071048

Chicago/Turabian Style

Liu, Wanli, Jinfeng Liu, Xiaoqian Wang, Jiazhen He, Yuqing Li, and Yong Liu. 2023. "Synthesis of Asymmetrical CsPbBr3/TiO2 Nanocrystals with Enhanced Stability and Photocatalytic Properties" Catalysts 13, no. 7: 1048. https://doi.org/10.3390/catal13071048

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