Next Article in Journal
Effect of Ti Doping on the Microstructure and Properties of SiCp/Al Composites by Pressureless Infiltration
Next Article in Special Issue
Effects on Metallization of n+-Poly-Si Layer for N-Type Tunnel Oxide Passivated Contact Solar Cells
Previous Article in Journal
Fabrication of Poly(ethylene furanoate)/Silver and Titanium Dioxide Nanocomposites with Improved Thermal and Antimicrobial Properties
Previous Article in Special Issue
Performance of Large Area n-TOPCon Solar Cells with Selective Poly-Si Based Passivating Contacts Prepared by PECVD Method
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of Size-Adjustable CsPbBr3 Perovskite Quantum Dots for Potential Photoelectric Catalysis Applications

International School of Materials Science and Engineering (ISMSE), State Key Laboratory of Advanced Technology for Materials Synthesis and Processing, Wuhan University of Technology, Wuhan 430070, China
*
Author to whom correspondence should be addressed.
Materials 2024, 17(7), 1607; https://doi.org/10.3390/ma17071607
Submission received: 12 March 2024 / Revised: 25 March 2024 / Accepted: 26 March 2024 / Published: 1 April 2024
(This article belongs to the Special Issue Advanced Materials for Solar Energy Utilization)

Abstract

:
As a direct band gap semiconductor, perovskite has the advantages of high carrier mobility, long charge diffusion distance, high defect tolerance and low-cost solution preparation technology. Compared with traditional metal halide perovskites, which regulate energy band and luminescence by changing halogen, perovskite quantum dots (QDs) have a surface effect and quantum confinement effect. Based on the LaMer nucleation growth theory, we have synthesized CsPbBr3 QDs with high dimensional homogeneity by creating an environment rich in Br ions based on the general thermal injection method. Moreover, the size of the quantum dots can be adjusted by simply changing the reaction temperature and the concentration of Br ions in the system, and the blue emission of strongly confined pure CsPbBr3 perovskite is realized. Finally, optical and electrochemical tests suggested that the synthesized quantum dots have the potential to be used in the field of photocatalysis.

1. Introduction

Over the past years, metal halide perovskite, a new type of semiconductor material, has attracted wide attention in the photovoltaic [1], luminescence [2] and detection [3] fields because of its excellent photoelectric properties. Compared with traditional semiconductors, perovskites have a special structure, and the general formula is ABX3 [4]. This paper focuses on metallic lead halide perovskites with an A-site of MA+, FA+ or Cs+; the B-site is occupied by metal cations such as Pb2+ or Sn2+, and they have an X-site of Cl, Br or I. The B-site atoms are in the body center of the cube and form a regular octahedral structure of BX64− with the X-site atoms, the A-site atoms are in the vertex of the cube, and correspondingly, the X-site atoms are in the six face centers of the cube [5,6,7,8,9].
This material is a direct bandgap semiconductor with advantages such as high carrier mobility, long charge diffusion distance, high defect tolerance and a low-cost solution preparation process. Today, perovskite solar cells have achieved a photoelectric conversion efficiency of more than 25.5%, and perovskite quantum dots (QDs) combine the properties of perovskite block materials and traditional semiconductor quantum dots with excellent photoelectric properties, such as high fluorescence quantum yield, extremely narrow half-peak width, adjustable band gap, full-spectrum coverage in visible light, etc. These properties make perovskite quantum dots one of the most promising photoelectric materials at present; they are therefore widely used in light-emitting diodes, solar cells, photodetectors and other fields [10,11,12,13]. Meanwhile, there are relevant examples of the application of quantum dots in the field of photocatalysis [14]. Similarly, composite materials such as perovskite quantum dots combined with other materials to form heterojunctions also have the potential for photocatalytic applications [15]. Quantum dots are quasi-zero-dimensional particles composed of a small number of atoms whose dimensions are close to their Bohr radii [16]. In quantum dots, charge carriers are restricted in all directions and are sometimes referred to as artificial atoms [17]. Quantum dots can be thought of as bridges between molecular clusters and bulk crystals depending on their size. For semiconductors, an extremely important parameter is the gap separating the conduction band (CB) and the valence band (VB), which is fixed for macroscopic crystals and is determined only by the material properties. However, for quantum dots, when the size is reduced to the nanometer level, the electrons are limited by the boundary during movement, and the corresponding energy level will be adjusted to match the change in size. When a quantum dot is excited by a photon of higher energy, the electrons in the VB experience energy fluctuations and are excited to the CB, while a hole is left in the valence band, and the formed electron–hole pair (charge carriers) will be attracted to each other by electrostatic action; the average distance between the electron and the hole is regarded as the Bohr radius (The Bohr radius of CsPbBr3 is 7 nm). The diffusion of charge carriers in the semiconductor lattice is related to the Bohr radius. When the size of the quantum dot is less than or close to its Bohr radius, the quantization of energy will be triggered, which will cause the electronic energy-level structure of the semiconductor material to change from a continuous energy level to a discrete energy level, and also cause the band gap between CB and VB to widen; this is the quantum-limited effect [18,19,20]. The quantum confinement effect can also be simply understood: the blue shift of the absorption and fluorescence spectra occurs as the size of the semiconductor decreases, and the blue shift of the spectrum becomes more obvious with a smaller size. In addition, a reduction in size will result in an increase in the proportion of atoms on the surface. Compared with bulk materials, quantum dots have a large specific surface area, and the surface atoms will form a large number of suspension bonds due to incomplete coordination [21,22]. The existence of suspension bonds may introduce defect levels in the semiconductor band gap and cause non-radiative recombination. At this time, the surface plays a decisive role in the structure and optical properties of the quantum dot [23,24,25,26]. In short, there are some properties distinguishing quantum dots compared to bulk materials.
Like traditional quantum dots, perovskite quantum dots can regulate the band gap through the quantum confinement effect to achieve tunable luminescence, and the absorption spectrum will gradually shift toward blue with the decrease in quantum dot size [27,28]. Another, lighter way is to adjust the luminescence through component engineering, that is, by adjusting the proportion of halogens. This is mainly because the CB of perovskite materials is contributed by the 6p orbital of Pb and the np orbital of halogens, and the 6p orbital contribution of Pb is a major part. The VB tops are mainly contributed by the 6s orbits of Pb and the np orbits of halogens [29]. For all metallic lead halide perovskites, the composition of lead is the same, so the band gap of perovskite can be continuously adjusted easily by adjusting the halogen composition, and, thus, the emission of perovskite in the entire visible band can be realized [30]. In addition, due to the soft ionic properties of halide perovskites, there is strong ionic interaction with surface ligands [31,32,33]. This makes it easy to exchange anions between perovskites with different halogens, and it is also easy and convenient to adjust the full spectrum of the visible light band of perovskites through composition engineering [34]. But, unfortunately, it is precisely because of this characteristic that the phase separation of perovskite mixed halogens occurs very easily in practical applications. Conversely, this problem can be fundamentally solved by regulating luminescence through quantum confinement effects. At the same time, how to prepare and control the size of quantum dots and ensure the uniformity of particle size has become a new problem to be solved urgently [35]. Based on previous work, we know that the synthesis of quantum dots follows LaMer nucleation growth theory. It mainly includes three parts; first, the monomer formation stage: because the precursor reacts under certain conditions, the monomer is formed quickly. Second, the nucleation stage: when the concentration of monomer exceeds the critical concentration of nucleation, it will be rapidly nucleated due to supersaturation, and a large amount of monomer will be consumed during nucleation, thereby reducing the monomer concentration to the critical concentration of nucleation again. Third, the growth stage: since the monomer concentration is lower than the critical concentration of nucleation, but still higher than the equilibrium concentration, within this range, QDs will carry out the growth stage [36]. Then, as the concentration of monomers in the solution is depleted, it enters the Ostwald ripening process. This process involves the dissolution of small-size QDs and the continued growth of large-size QDs, which will cause the widening of QDs [37]. The results show that accelerating the monomer formation rate can reduce the nucleation time of QDs, which is conducive to the formation of a more uniform QDs core. The homogeneous core can improve the uniformity of subsequent growth and inhibit the Ostwald ripening process, thus achieving the synthesis of QDs with a narrow size distribution.
Here, we add ZnBr2 to create a Br ion-rich reaction system on the basis of the general hot injection method, and effectively regulate the synthesis of CsPbBr3 QDs based on LaMer nucleation growth theory [38]. By simply adjusting the concentration of Br ions and changing the reaction temperature, CsPbBr3 QDs of different sizes with ultra-high particle size uniformity can be obtained easily. These CsPbBr3 QDs of similar size exhibit a certain self-assembly behavior due to their high homogeneity, which can be demonstrated by their XRD tests [39]. X-ray photoelectron spectroscopy (XPS) testing can also effectively prove that the Cs, Pb and Br elements in CsPbBr3 QDs have a reasonable chemical state environment. In addition, the ultraviolet–visible absorption (UV-vis) spectrum and photoluminescence (PL) spectrum can prove that, based on the quantum confinement effect, the corresponding luminous band of CsPbBr3 QDs is significantly blue-shifted with the reduction in their size [40]. Finally, optical and electrochemical tests show that the perovskite quantum dots have great potential for application in the field of photocatalysis.

2. Materials and Methods

2.1. Materials

Cesium carbonate (Cs2CO3, Aladdin, Shanghai China, 99.9%), lead bromide (PbBr2, Aladdin, Shanghai China, 99.99%), zinc bromide (ZnBr2, Aladdin, Shanghai China, 99%), oleic acid (OA, Aladdin, Shanghai China, AR), oleylamine (OAm, Aladdin, Shanghai China, 80~90%), 1-octadecene (ODE, Aladdin, Shanghai China, >90% (GC)), n-hexane (Aladdin, Shanghai China, >99%), acetone (Aladdin, Shanghai China, >99%).

2.2. Methods

2.2.1. Preparation of CsOA Precursors

The Cs-oleate precursor solution was prepared in a 50 mL three-necked round-bottomed flask by dissolving Cs2CO3 (0.4 g) in a mixture of oleic acid (OA, 5 mL) and 1-octadecene (ODE, 15 mL). After drying with a vacuum pump for 1 h at 120 °C, it was heated under N2 protection to 140 °C until all Ca2CO3 dissolved to form Cs-oleate. The Cs-oleate precursor solution was preheated to 120 °C before injection.

2.2.2. Synthesis of CsPbBr3 Quantum Dots (QDs)

ODE (10 mL), PbBr2 (0.138 g) and ZnBr2 were loaded in a 50 mL 3-necked flask and vacuum-dried for 1 h at 120 °C, and then oleylamine (2 mL) and OA (2 mL) were injected under a N2 atmosphere. After the PbBr2 and ZnBr2 dissolved, the temperature was set to a different temperature, and the Cs-oleate solution (0.8 mL, 0.125 M in ODE) was quickly injected in. The reaction mixture was cooled by an ice-water bath after reacting for 30 s for the crystal growth. The product was collected by centrifugation. The amount of ZnBr2 added depended on the molar ratio of bromine/lead. When the molar ratio of Br:Pb = 6:1, ZnBr2 (0.17 g) was required, and when the molar ratio of Br:Pb = 10:1, twice as much was required.

2.2.3. Purification of CsPbBr3 Quantum Dots (QDs)

For the product in the centrifugation, the supernatant was left on the bench top under ambient conditions for ~3 h until the unreacted salts precipitated. Then the mixture was centrifuged to get the clear supernatant. With a proper amount of the clarified solution, ~5 mL of acetone was slowly added until the mixture became turbid to avoid decomposition of the QDs. Then the QDs were centrifuged at 10,000 rpm for 10 min and the precipitate was collected and dissolved in ~500 μL of hexane.

2.3. Characterization

2.3.1. Characterization via Transmission Electron Microscopy (TEM)

Images captured via transmission electron microscopy (TEM) were generated utilizing a JEOL JEM-1400 Plus microscope (Beijing, China). This instrument, featuring a thermionic emission source, was set to function at an acceleration power of 120 kV. For the examination, quantum dot (QD) specimens underwent a precise preparation process, involving the deposition of a thinned QD mixture in n-hexane onto grids coated with carbon on copper. Following this, the analysis of the QDs’ size distribution was conducted through the application of Nano Measurer software (1.2), bearing version number 1.02.0005.

2.3.2. Characterization via X-ray Diffraction (XRD)

X-ray diffraction (XRD) measurements were conducted utilizing a Bruker D8 ADVANCE diffractometer, from Karlsruhe, Germany, employing Cu-Kα radiation at a wavelength of 1.540598 Å. Data collection covered a 2θ range from 10° to 60°, utilizing a step increment of 0.05° and a dwell time of 0.5 s per step. The instrument was set to a voltage of 40 kV and a current of 40 mA. To prepare for XRD, a dense suspension of quantum dots (QDs) in n-hexane was drop-cast onto silicon dioxide/silicon (SiO2/Si) supports. Analysis of the XRD patterns was subsequently performed with Jade 6.5 software.

2.3.3. X-ray Photoelectron Spectroscopy (XPS) Characterization

Spectra from X-ray photoelectron spectroscopy (XPS) of perovskite quantum dots were captured with an AXIS SUPRA spectrometer from Kratos Analytical, part of the Shimadzu Group, located in Manchester, UK. This device features a Thermo Scientific Kα spectrometer and utilizes a monochromatic Al-Kα X-ray for illumination. Data collected were then transformed into VGD format and analyzed using Avantage software (599-31), release 5.9922. Calibration of the binding energy scale was aligned using the C 1 s peak, attributing a binding energy of 284.8 eV specifically to the carbon–carbon (C–C) bond.

2.3.4. Fluorescence Spectrum Characterization

Optical absorption spectra in the UV–visible range were obtained using a Shimadzu UV-1800 spectrophotometer (Kyoto, Japan), which was operated with UVProbe 2.52 software. Concurrently, the emission characteristics were assessed through steady-state photoluminescence (PL) spectra, recorded using a Shimadzu RF-6000 spectrophotometer (Kyoto, Japan) equipped with LabSolutions RF software (1.11). For CsPbBr3, the excitation wavelength (λex) utilized was 400 nm. The preparation for optical evaluation involved the dilution of quantum dot (QD) solutions in n-hexane, followed by their allocation into quartz cuvettes, each with a path length of 10 mm, to ensure accurate measurements.

2.3.5. Electrochemistry (Photo) Measurements

Photocurrent and electrochemical impedance analyses on halide perovskite quantum dots (QDs) were conducted using a Zennium workstation, supplied by Zahner Company of Kronach, Germany. This investigation utilized a tri-electrode configuration comprising a sample-based working electrode, a counter-electrode of platinum disk, and a Ag/AgCl reference electrode (in saturated KCl solution) with a potential of +0.1989 V relative to the Normal Hydrogen Electrode (NHE) within an electrolyte solution of 0.5 M Na2SO4. Illumination for the photocurrent assessments was provided by a 405 nm wavelength LED, with its intensity quantified using a Newport photometer (Irvine, CA, USA). The electrochemical impedance spectra were acquired at a potential of −0.1 V versus NHE, with application of a 10 mV signal over a frequency range from 100 mHz to 20 kHz, in a 0.5 M Na2SO4 solution.

3. Results

The traditional synthesis of CsPbBr3 nanocrystals involves simple dissolution of PbBr2 and ligands in a solvent and hot injection of cesium oleate to obtain the corresponding halogen perovskite nanocrystals. Although monodisperse nanocrystals have been synthesized successfully, their sizes are not uniform, and effective size control methods are lacking. Based on the LaMer nucleation growth theory, it can be seen that this is an extremely rapid explosive nucleation process, accompanied by the Oswald maturation process in parallel with crystal growth. Therefore, a BR-rich reaction system was constructed by introducing ZnBr2, which can improve the formation of perovskite monomer and shorten the nucleation time. The homogeneous core is conducive to inhibiting the Oswald ripening process in the later stage of growth, avoiding size widening and obtaining a narrow size distribution of quantum dots. In addition, the size of the perovskite QDs can be adjusted by changing the reaction temperature during hot injection. As the size of QDs decreases, we obtain strongly confined CsPbBr3 QDs. Due to the great quantum confinement effect, the luminous wavelength of CsPbBr3 QDs shifts from green to blue. Compared with the component engineering of Cl/Br mixed halogens, the blue light formed based on dimension engineering fundamentally solves the problem of easy phase separation of mixed halogens. At the same time, with the addition of excessive Br ions, the Br vacancy defects on the surface of perovskite are greatly passivated [41]. In bromine-rich reaction systems, Pb2+ and Br ions are combined to form a complex of [PbBr3] firstly, and then the complex is transformed into an octahedron. Once the octahedron is formed, sufficient nucleation sites are provided for rapid nucleation with the addition of Cs-OA. We obtained CsPbBr3 QDs of different sizes at different reaction temperatures (150 °C, 130 °C, 110 °C, 90 °C) (Scheme 1). Based on the weak-to-strong quantum confinement effect, the continuously tunable emission of CsPbBr3 QDs from green light to blue light was realized.
Shown in Figure 1A–D are TEM images of CsPbBr3 QDs corresponding to reaction temperatures of 150 °C, 130 °C, 110 °C and 90 °C, respectively. The size of CsPbBr3 QDs can be visually observed through TEM images. It can be clearly seen from the TEM images that the size of the obtained CsPbBr3 QDs decreased with the gradual decrease in the reaction temperature during hot injection. In addition, we calculated the sizes of hundreds of quantum dots, made a bar chart, and placed it under the TEM images of quantum dots we obtained, arranged by the corresponding reaction temperatures (Figure 1E–H). According to the information in the bar chart, the average size and particle size distribution of CsPbBr3 QDs obtained, from high to low reaction temperature, are d = 5.7 ± 0.1 nm, d = 4.9 ± 0.05 nm, d = 3.9 ± 0.1 nm and d = 3.5±0.11 nm, respectively. In addition, comparing Figure S1 and Figure 1, it can be found that as the molar ratio of bromine to lead increases from Br: Pb = 6:1 to Br: Pb = 10:1, the sizes of QDs in the system with higher concentrations of Br ions at the same reaction temperature become smaller. This shows that in addition to the temperature condition, the change in Br ion content in the system also really affects the size of the quantum dot. At the same time, in an environment with high-energy particle impacts such as TEM, there are no white spots, which indicates the destruction of the structure, and small black patches, which indicate the collapse of the structure, leading to the precipitation of PbBr2. Studies indicate that increasing the concentration of halogen ions can improve the radiation resistance of halide perovskite QDs [42]. This may be attributed to the fact that with the increase in Br ion content in the system, more nucleation sites are formed while the surface defects are passivated. The increase in nucleation sites greatly accelerates the nucleation process and, thus, effectively inhibits Oswald ripening in some aspects, which means that the sizes of the perovskite quantum dots with extremely high size uniformity obtained by the improved method in this paper can be controlled simply and effectively. What is more interesting is that, compared with the monodisperse CsPbBr3 QDs synthesized by the general method, they show a chaotic arrangement (Figure S2). The improved CsPbBr3 QDs synthesized in this paper are arranged in an orderly manner on the substrate, and show a trend of self-assembly. It should be clear that this trend of self-assembly has high requirements for the dimensional uniformity of the assembled units, which is also evidence that the synthesized CsPbBr3 QDs have ultra-high dimensional uniformity. This high dimensional uniformity and size adjustability will also greatly facilitate the installation of CsPbBr3 QDs in channel materials, further developing their potential in areas such as catalysis.
It is noteworthy that the X-ray diffraction (XRD) pattern exhibits only two highly intense peaks, near the diffraction angles 2θ = 15° and 2θ = 30°. These peaks correspond to the (001) and (002) crystal faces of CsPbBr3 perovskite, respectively. The presence of these specific peaks, in accordance with the ICSD code 029073, confirms the cubic crystal structure of CsPbBr3 QDs (Figure S3). Just like the extremely ordered arrangement of quantum dots in the TEM images, CsPbBr3 QDs are arranged along the preferred orientation of the (001) crystal plane, with structural coherence, so that there are extremely strong diffraction peaks at the peaks of the (001) crystal plane and the (002) crystal plane in the XRD pattern. The special XRD pattern of the synthesized CsPbBr3 QDs is in good agreement with the special arrangement in the TEM images. Figure 2B–D and Figure S4 show the XPS spectrum of CsPbBr3 QDs. According to the curve of the spectrum, the chemical valence states of Cs, Pb and Br were analyzed [43]. Figure 2B shows the binding energy curve of the Cs 3d orbit, where the peaks corresponding to Cs 3d5/2 and Cs 3d3/2 are located near 724 eV and 738 eV, respectively. Similarly, Figure 2C shows the binding energy curve of the Pb 4f orbit, where the peaks corresponding to Pb 4f7/2 and Pb 4f5/2 are located near 138 eV and 143 eV, respectively. In addition, in Figure 2D, the coupling peaks near 68 eV and 69 eV come from the Br 3d orbit, representing Br 3d5/2 and Br 3d3/2, respectively. In the above peak location, the peak location of the Pb 4f track is similar to that recorded in reference [44].
Subsequently, we performed optical characterization of CsPbBr3 quantum dot samples at various reaction temperatures, including UV-vis and photoluminescence (PL) spectra. As the average particle size of CsPbBr3 QDs decreases, the quantum confinement effect changes from weak to strong, the continuous electronic energy levels split into discrete energy levels, and the energy levels are constantly widened. Shown in Figure 3A–D are CsPbBr3 quantum dot samples corresponding to reaction temperatures of 150 °C, 130 °C, 110 °C and 90 °C, respectively. We dispersed them in hexane and obtained the transition of excitation light from green to blue light under the excitation of an ultraviolet lamp. The corresponding UV-vis absorption peak and PL emission peak also gradually transited from a green band to a deep blue band, indicating that the band gap gradually widened. According to the detailed information shown in the UV-vis spectra and PL spectra in Figure 3, the emission peak positions of PL shift from 494 nm, 484 nm and 474 nm to 464 nm (deep blue light), while the corresponding absorption peak positions shift from 485 nm, 474 nm and 459 nm to 450 nm. It should be noted that the intrinsic PL excitation band of CsPbBr3 QDs should be around 520 nm. For details, see the illustration in Figure S5. The solution of CsPbBr3 QDs dispersed with larger sizes appears bright green under ultraviolet irradiation. The excited light color of the CsPbBr3 QDs solution obtained at the reaction temperature of 150 °C under the UV lamp is more like the transition color between green and blue, because the average particle size of the sample at this time was d ≈ 5.7 nm < 7 nm. At this time, an obvious quantum confinement effect appeared, and the original continuous electronic energy levels split into discrete energy levels. As the energy level widened, the band gap increased and the PL peak position shifted to blue. The quantum confinement effect can also increase the exciton binding energy of the crystal; the exciton binding energy in the halogen perovskite film is about 25 meV, and the excitons in the perovskite film will decompose into free electrons and holes at room temperature. For perovskite quantum dots, the exciton binding energy can reach hundreds of meV, and the increase in exciton binding energy increases the possibility of radiation recombination, which can significantly improve luminous performance. In addition, the adjustable band gap of CsPbBr3 QDs also has great potential in the field of photocatalysis. By adjusting the band gap of CsPbBr3 QDs and other materials to form heterostructures (Figure S7), it is easier to form a type II band structure, effectively realize the separation of electrons and holes, and improve the catalytic performance.
CsPbBr3 quantum dots (QDs), fabricated at a temperature of 150 °C, were subjected to assessments of photocurrent response and electrochemical impedance spectroscopy (EIS), as illustrated in Figure 4. The equivalent circuit analysis involved CPE as the notation for double-layer capacitance, with RS representing the resistance across the electrolyte between the reference and working electrodes, and RC indicating the resistance related to charge transfer at the electrode interface. The EIS patterns for the CsPbBr3 QDs showcased a distinct semicircle, a marker for charge transfer resistance. To accurately measure the photocurrent response exclusive of external effects, measurements were systematically taken with the initiation of the light source. Remarkably, when the illumination was discontinued, the photocurrent density approached zero, but surged immediately when the light was reinstated after a 30 s interval. Such periodic fluctuation in current, with light modulation every 30 s, underscores the CsPbBr3 perovskite’s adeptness at carrier transport. This efficiency in carrier migration underscores the potential of CsPbBr3 QDs for future photoelectrocatalysis applications.

4. Conclusions

Based on the synthesis of perovskite by the hot injection method, we introduced more Br ions by adding ZnBr2 to create a bromine-rich reaction environment to regulate the nucleation and growth process of perovskite, and finally synthesized monodisperse, extremely uniform and adjustable CsPbBr3 QDs. Due to the presence of excessive Br ions in the solution, the reaction system had abundant nucleation sites, which greatly sped up the explosive nucleation process in the hot injection stage, and effectively inhibited the Oswald ripening process in the subsequent crystal growth. Thus, CsPbBr3 QDs with an extremely uniform size were obtained. We can regulate the size of CsPbBr3 QDs by adjusting the reaction temperature during hot injection and the amount of Br ions added. The general rule is as follows: when the content of Br ions in the system is constant, the size of the obtained quantum dots becomes smaller and smaller as the reaction temperature gradually decreases. Similarly, when the temperature is constant, the size of the quantum dot is negatively correlated with the Br ion content in the system to a certain extent. In addition, because the size is uniform enough, there is an interesting self-assembly arrangement, which can be clearly seen in TEM images and is to some extent supported by XRD. In addition, based on the quantum confinement effect, the size changes of CsPbBr3 QDs also bring special optical properties. When the size of the quantum dot enters the critical value of the Boer exciton radius (7 nm), with a further reduction in the average size, its photoluminescence wavelength shifts to a deep blue band. The results of spectral and structural characterization show that the emission of CsPbBr3 perovskite, from green light to blue light, can be continuously adjusted by adjusting the sizes of quantum dots. In general, on the one hand, from the perspective of structure, CsPbBr3 QDs can achieve more efficient catalytic performance by attaching to other materials, and the adjustable size property greatly facilitates this method. Furthermore, based on the change in band gap width caused by the size effect, the formation of a type II band structure with other materials also has the potential to be applied in the field of catalysis.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ma17071607/s1, Figure S1: (A–D) TEM images of CsPbBr3 QDs at different reaction temperatures (scale bar: 50 nm) and (E,F) their size-of-particle-diameter distribution diagrams; Figure S2: Image of CsPbBr3 QDs at different reaction temperatures under ultraviolet lamp excitation and the corresponding UV-vis and PL spectra; Figure S3: (a) HR-TEM image and its magnified local view; (b) the reciprocal space image after FFT; (c) the lattice image obtained by inverse FFT and the reciprocal space image; Figure S4: The XPS spectrum of CsPbBr3 QDs; Figure S5: Image of CsPbBr3 QDs under ultraviolet lamp excitation and the corresponding PL spectra; Figure S6: (a) TEM image and (b) EDS spectra of CsPbBr3 QDs; Figure S7: The relationship between absorbance and photon energy. The band gap (Bg) of CsPbBr3 QDs of different sizes.

Author Contributions

H.L. and Y.L. conceived projection and material synthetic design. H.L., X.W. and J.H. performed the experiment, sample testing, data collection and analysis. H.L., Q.L. and X.L. performed data discussion and formal analyses. H.L., J.L., C.L. and M.W. wrote, reviewed and edited the manuscript. All authors discussed the results. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Natural Science Foundation of China (52072281, Y.L.), the Major Program of the National Natural Science Foundation of China (22293021, Y.L.), the National Innovation and Entrepreneurship Training Program for College Students (No. S202210497011, Y.L.), Wuhan University of Technology Doctoral Research Initiation Fee (40120543, Y.L.), and Hubei Province’s state-allocated dual-class research platform fee (40120623, Y.L.).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

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

Acknowledgments

We are grateful to the School of Chemistry and Chemical and Life Sciences of Wuhan University of Technology for providing UV-vis and PL characterization and to the Center for Materials Research and Analysis of Wuhan University of Technology Materials for XPS and HR-TEM testing.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Yuan, J.; Hazarika, A.; Zhao, Q.; Ling, X.; Moot, T.; Ma, W.; Luther, J.M. Metal Halide Perovskites in Quantum Dot Solar Cells: Progress and Prospects. Joule 2020, 4, 1160–1185. [Google Scholar] [CrossRef]
  2. Solari, S.F.; Kumar, S.; Jagielski, J.; Shih, C.J. Monochromatic LEDs based on perovskite quantum dots: Opportunities and challenges. J. Soc. Inf. Disp. 2019, 27, 667–678. [Google Scholar] [CrossRef]
  3. Li, X.; Wang, Y.; Chen, T.; Xu, T.; Ye, S.; Liu, R.; Sun, Z.; Tan, C.; Lv, X.; Yang, J.; et al. Unencapsulated CsPbClBr2 Film Photodetectors Grown by Thermal Vacuum Deposition Exhibit Exceptional Environmental Stability in High-Humidity Air. ACS Appl. Energy Mater. 2022, 5, 8709–8716. [Google Scholar] [CrossRef]
  4. Akkerman, Q.A.; Manna, L. What Defines a Halide Perovskite? ACS Energy Lett. 2020, 5, 604–610. [Google Scholar] [CrossRef] [PubMed]
  5. Ceratti, D.R.; Tenne, R.; Bartezzaghi, A.; Cremonesi, L.; Segev, L.; Kalchenko, V.; Oron, D.; Potenza, M.A.C.; Hodes, G.; Cahen, D. Self-Healing and Light-Soaking in MAPbI3: The Effect of H2O. Adv. Mater. 2022, 34, 2110239. [Google Scholar] [CrossRef] [PubMed]
  6. Li, Y.; Ding, T.; Luo, X.; Tian, Y.; Lu, X.; Wu, K. Synthesis and Spectroscopy of Monodispersed, Quantum-Confined FAPbBr3 Perovskite Nanocrystals. Chem. Mater. 2019, 32, 549–556. [Google Scholar] [CrossRef]
  7. Saparov, B.; Mitzi, D.B. Organic–Inorganic Perovskites: Structural Versatility for Functional Materials Design. Chem. Rev. 2016, 116, 4558–4596. [Google Scholar] [CrossRef] [PubMed]
  8. Seth, S.; Ahmed, T.; De, A.; Samanta, A. Tackling the Defects, Stability, and Photoluminescence of CsPbX3 Perovskite Nanocrystals. ACS Energy Lett. 2019, 4, 1610–1618. [Google Scholar] [CrossRef]
  9. Zhang, J.; Zhong, Y. Origins of p-Doping and Nonradiative Recombination in CsSnI3. Angew. Chem. Int. Ed. 2022, 61, 106800. [Google Scholar]
  10. Fakharuddin, A.; Gangishetty, M.K.; Abdi-Jalebi, M.; Chin, S.H.; bin Mohd Yusoff, A.R.; Congreve, D.N.; Tress, W.; Deschler, F.; Vasilopoulou, M.; Bolink, H.J. Perovskite light-emitting diodes. Nat. Electron. 2022, 5, 203–216. [Google Scholar] [CrossRef]
  11. Jing, H.; Peng, R.; Ma, R.M.; He, J.; Zhou, Y.; Yang, Z.; Li, C.Y.; Liu, Y.; Guo, X.; Zhu, Y.; et al. Flexible Ultrathin Single-Crystalline Perovskite Photodetector. Nano Lett. 2020, 20, 7144–7151. [Google Scholar] [CrossRef] [PubMed]
  12. Kim, J.Y.; Lee, J.W.; Jung, H.S.; Shin, H.; Park, N.G. High-Efficiency Perovskite Solar Cells. Chem. Rev. 2020, 120, 7867–7918. [Google Scholar] [CrossRef] [PubMed]
  13. Tao, S.; Yang, D.; He, G.; Guo, D.; Yang, L.; Zheng, J.; Li, J.; Chen, J.; Ma, D. Photomultiplication-type perovskite photodetectors base on air-processed perovskite films. Org. Electron. 2023, 118, 106800. [Google Scholar] [CrossRef]
  14. Kannan, R.; Kim, A.R.; Eo, S.K.; Kang, S.H.; Yoo, D.J. Facile one-step synthesis of cerium oxide-carbon quantum dots/RGO nanohybrid catalyst and its enhanced photocatalytic activity. Ceram. Int. 2017, 43, 3072–3079. [Google Scholar] [CrossRef]
  15. 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. [Google Scholar] [CrossRef]
  16. Rossetti, R.; Nakahara, S.; Brus, L.E. Quantum size effects in the redox potentials, resonance Raman spectra, and electronic spectra of CdS crystallites in aqueous solution. J. Chem. Phys. 1983, 79, 1086–1088. [Google Scholar] [CrossRef]
  17. Kagan, C.R.; Murray, C.B. Charge transport in strongly coupled quantum dot solids. Nat. Nanotechnol. 2015, 10, 1013–1026. [Google Scholar] [CrossRef] [PubMed]
  18. Halperin, W.P. Quantum size effects in metal particles. Rev. Mod. Phys. 1986, 58, 533–606. [Google Scholar] [CrossRef]
  19. Feldheim, L.; Keating, C.D. Self-assembly of single electron transistors and related devices. Chem. Soc. Rev. 1998, 27, 1–12. [Google Scholar] [CrossRef]
  20. Zhou, C.; Lin, H.; Tian, Y.; Yuan, Z.; Clark, R.; Chen, B.; van de Burgt, L.J.; Wang, J.C.; Zhou, Y.; Hanson, K.; et al. Luminescent zero-dimensional organic metal halide hybrids with near-unity quantum efficiency. Chem. Sci. 2018, 9, 586–593. [Google Scholar] [CrossRef]
  21. Ioannou, D.; Griffin, D.K. Nanotechnology and molecular cytogenetics: The future has not yet arrived. Nano Rev. 2010, 1, 5117. [Google Scholar] [CrossRef] [PubMed]
  22. Xu, Y.F.; Yang, M.Z.; Chen, B.X.; Wang, X.D.; Chen, H.Y.; Kuang, D.B.; Su, C.Y. A CsPbBr3 Perovskite Quantum Dot/Graphene Oxide Composite for Photocatalytic CO2 Reduction. J. Am. Chem. Soc. 2017, 139, 5660–5663. [Google Scholar] [CrossRef] [PubMed]
  23. Li, P.; Wei, Z.; Wu, T.; Peng, Q.; Li, Y. Au−ZnO Hybrid Nanopyramids and Their Photocatalytic Properties. J. Am. Chem. Soc. 2011, 133, 5660–5663. [Google Scholar] [CrossRef] [PubMed]
  24. Petronella, F.; Pagliarulo, A.; Striccoli, M.; Calia, A.; Lettieri, M.; Colangiuli, D.; Curri, M.L.; Comparelli, R. Colloidal Nanocrystalline Semiconductor Materials as Photocatalysts for Environmental Protection of Architectural Stone. Crystals 2017, 7, 30. [Google Scholar] [CrossRef]
  25. Pincik, E.; Matsumoto, T. Progress in applied surface, interface and thin film science 2017. Appl. Surf. Sci. 2018, 461, 1–2. [Google Scholar] [CrossRef]
  26. Ball, P.; Garwin, L. Science at the atomic scale. Nature 1992, 355, 761–764. [Google Scholar] [CrossRef]
  27. Fu, P.; Shan, Q.; Shang, Y.; Song, J.; Zeng, H.; Ning, Z.; Gong, J. Perovskite nanocrystals: Synthesis, properties and applications. Sci. Bull. 2017, 62, 369–380. [Google Scholar] [CrossRef] [PubMed]
  28. Protesescu, L.; Yakunin, S.; Bodnarchuk, M.I.; Krieg, F.; Caputo, R.; Hendon, C.H.; Yang, R.X.; Walsh, A.; Kovalenko, M.V. Nanocrystals of Cesium Lead Halide Perovskites (CsPbX3, X = Cl, Br, and I): Novel Optoelectronic Materials Showing Bright Emission with Wide Color Gamut. Nano Lett. 2015, 15, 3692–3696. [Google Scholar] [CrossRef]
  29. Endres, J.; Egger, D.A.; Kulbak, M.; Kerner, R.A.; Zhao, L.; Silver, S.H.; Hodes, G.; Rand, B.P.; Cahen, D.; Kronik, L.; et al. Valence and Conduction Band Densities of States of Metal Halide Perovskites: A Combined Experimental–Theoretical Study. J. Phys. Chem. Lett. 2016, 7, 2722–2729. [Google Scholar] [CrossRef]
  30. Liu, Y.; Li, F.; Li, Q.; Yang, K.; Guo, T.; Li, X.; Zeng, H. Emissions at Perovskite Quantum Dot/Film Interface with Halide Anion Exchange. ACS Photonics 2018, 5, 4504–4512. [Google Scholar] [CrossRef]
  31. Akkerman, Q.A.; D’Innocenzo, V.; Accornero, S.; Scarpellini, A.; Petrozza, A.; Prato, M.; Manna, L. Tuning the Optical Properties of Cesium Lead Halide Perovskite Nanocrystals by Anion Exchange Reactions. J. Am. Chem. Soc. 2015, 137, 10276–10281. [Google Scholar] [CrossRef]
  32. Jiang, Y.; Sun, C.; Xu, J.; Li, S.; Cui, M.; Fu, X.; Liu, Y.; Liu, Y.; Wan, H.; Wei, K.; et al. Synthesis-on-substrate of quantum dot solids. Nature 2022, 612, 679–684. [Google Scholar] [CrossRef]
  33. Uddin, M.A.; Glover, J.D.; Park, S.M.; Pham, J.T.; Graham, K.R. Growth of Highly Stable and Luminescent CsPbX3 (X = Cl, Br, and I) Nanoplates via Ligand Mediated Anion Exchange of CsPbCl3 Nanocubes with AlX3. Chem. Mater. 2020, 32, 5217–5225. [Google Scholar] [CrossRef]
  34. Wang, X.; Liu, W.; He, J.; Li, Y.; Liu, Y. Synthesis of All-Inorganic Halide Perovskite Nanocrystals for Potential Photoelectric Catalysis Applications. Catalysts 2023, 13, 1041. [Google Scholar] [CrossRef]
  35. Shekhirev, M.; Goza, J.; Teeter, J.D.; Lipatov, A.; Sinitskii, A. Synthesis of Cesium Lead Halide Perovskite Quantum Dots. J. Chem. Educ. 2017, 94, 1150–1156. [Google Scholar] [CrossRef]
  36. Chu, D.B.K.; Owen, J.S.; Peters, B. Nucleation and Growth Kinetics from LaMer Burst Data. J. Phys. Chem. A 2017, 121, 7511–7517. [Google Scholar] [CrossRef]
  37. Huang, X.; Guo, Q.; Yang, D.; Xiao, X.; Liu, X.; Xia, Z.; Fan, F.; Qiu, J.; Dong, G. Reversible 3D laser printing of perovskite quantum dots inside a transparent medium. Nat. Photonics 2020, 14, 82–88. [Google Scholar] [CrossRef]
  38. Dong, Y.; Qiao, T.; Kim, D.; Parobek, D.; Rossi, D.; Son, D.H. Precise Control of Quantum Confinement in Cesium Lead Halide Perovskite Quantum Dots via Thermodynamic Equilibrium. Nano Lett. 2018, 18, 3716–3722. [Google Scholar] [CrossRef]
  39. Rainò, G.; Becker, M.A.; Bodnarchuk, M.I.; Mahrt, R.F.; Kovalenko, M.V.; Stöferle, T. Superfluorescence from lead halide perovskite quantum dot superlattices. Nature 2018, 563, 671–675. [Google Scholar] [CrossRef] [PubMed]
  40. Hu, H.; Guan, W.; Xu, Y.; Wang, X.; Wu, L.; Chen, M.; Zhong, Q.; Xu, Y.; Li, Y.; Sham, T.-K.; et al. Construction of Single-Atom Platinum Catalysts Enabled by CsPbBr3 Nanocrystals. ACS Nano 2021, 15, 13129–13139. [Google Scholar] [CrossRef] [PubMed]
  41. Wang, H.; Ye, F.; Sun, J.; Wang, Z.; Zhang, C.; Qian, J.; Zhang, X.; Choy, W.C.H.; Sun, X.W.; Wang, K.; et al. Efficient CsPbBr3 Nanoplatelet-Based Blue Light-Emitting Diodes Enabled by Engineered Surface Ligands. ACS Energy Lett. 2022, 7, 1137–1145. [Google Scholar] [CrossRef]
  42. Li, F.; Liu, Y.; Wang, H.; Zhan, Q.; Liu, Q.; Xia, Z. Postsynthetic Surface Trap Removal of CsPbX3 (X = Cl, Br, or I) Quantum Dots via a ZnX2/Hexane Solution toward an Enhanced Luminescence Quantum Yield. Chem. Mater. 2018, 30, 8546–8554. [Google Scholar] [CrossRef]
  43. Kim, H.; Bae, S.R.; Lee, T.H.; Lee, H.; Kang, H.; Park, S.; Jang, H.W.; Kim, S.Y. Enhanced Optical Properties and Stability of CsPbBr3 Nanocrystals Through Nickel Doping. Adv. Funct. Mater. 2021, 31, 2102770. [Google Scholar] [CrossRef]
  44. Liu, A.; Bi, C.; Li, J.; Zhang, M.; Cheng, C.; Binks, D.; Tian, J. High Color-Purity and Efficient Pure-Blue Perovskite Light-Emitting Diodes Based on Strongly Confined Monodispersed Quantum Dots. Nano Lett. 2023, 23, 2405–2411. [Google Scholar] [CrossRef]
Scheme 1. Illustrative representation of the CsPbBr3 QDs synthesis process.
Scheme 1. Illustrative representation of the CsPbBr3 QDs synthesis process.
Materials 17 01607 sch001
Figure 1. (AD) TEM images of CsPbBr3 QDs at different reaction temperatures, respectively (scale bar: 50 nm) and (EH) their size-of-particle-diameter distribution diagrams.
Figure 1. (AD) TEM images of CsPbBr3 QDs at different reaction temperatures, respectively (scale bar: 50 nm) and (EH) their size-of-particle-diameter distribution diagrams.
Materials 17 01607 g001
Figure 2. (A) XRD pattern and (BD) XPS spectra of CsPbBr3 QDs.
Figure 2. (A) XRD pattern and (BD) XPS spectra of CsPbBr3 QDs.
Materials 17 01607 g002
Figure 3. (AD) images of CsPbBr3 QDs at different reaction temperatures under ultraviolet lamp excitation and the corresponding UV-vis and PL spectra.
Figure 3. (AD) images of CsPbBr3 QDs at different reaction temperatures under ultraviolet lamp excitation and the corresponding UV-vis and PL spectra.
Materials 17 01607 g003
Figure 4. (A) Electrochemical impedance spectroscopy (Nyquist plot) of CsPbBr3 QDs synthesized at a reaction temperature of 150 °C. The inset illustrates the equivalent circuit model. (B) Instantaneous photocurrent response of CsPbBr3 QDs in neutral water (0.5 M Na2SO4) at −0.1 V vs. NHE.
Figure 4. (A) Electrochemical impedance spectroscopy (Nyquist plot) of CsPbBr3 QDs synthesized at a reaction temperature of 150 °C. The inset illustrates the equivalent circuit model. (B) Instantaneous photocurrent response of CsPbBr3 QDs in neutral water (0.5 M Na2SO4) at −0.1 V vs. NHE.
Materials 17 01607 g004
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

Li, H.; He, J.; Wang, X.; Liu, Q.; Luo, X.; Wang, M.; Liu, J.; Liu, C.; Liu, Y. Synthesis of Size-Adjustable CsPbBr3 Perovskite Quantum Dots for Potential Photoelectric Catalysis Applications. Materials 2024, 17, 1607. https://doi.org/10.3390/ma17071607

AMA Style

Li H, He J, Wang X, Liu Q, Luo X, Wang M, Liu J, Liu C, Liu Y. Synthesis of Size-Adjustable CsPbBr3 Perovskite Quantum Dots for Potential Photoelectric Catalysis Applications. Materials. 2024; 17(7):1607. https://doi.org/10.3390/ma17071607

Chicago/Turabian Style

Li, Hang, Jiazhen He, Xiaoqian Wang, Qi Liu, Xuemin Luo, Mingwei Wang, Jinfeng Liu, Chengqi Liu, and Yong Liu. 2024. "Synthesis of Size-Adjustable CsPbBr3 Perovskite Quantum Dots for Potential Photoelectric Catalysis Applications" Materials 17, no. 7: 1607. https://doi.org/10.3390/ma17071607

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