Next Article in Journal
Research on the Comparison Properties of PDMS Specimens Demolding Processes and the Mechanical Performance of Hollow-Solid Ratios of Flexible Telescopic Rods
Next Article in Special Issue
A Comprehensive Study of CsSnI3-Based Perovskite Solar Cells with Different Hole Transporting Layers and Back Contacts
Previous Article in Journal
Dynamic Characteristic Model of Giant Magnetostrictive Transducer with Double Terfenol-D Rods
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

High-Efficiency CsPbBr3 Light-Emitting Diodes using One-Step Spin-Coating In Situ Dynamic Thermal Crystallization

1
College of Physics, Changchun University of Science and Technology, Changchun 130012, China
2
College of Information Technology, Jilin Normal University, Siping 136000, China
*
Author to whom correspondence should be addressed.
Micromachines 2023, 14(6), 1104; https://doi.org/10.3390/mi14061104
Submission received: 18 April 2023 / Revised: 19 May 2023 / Accepted: 22 May 2023 / Published: 24 May 2023
(This article belongs to the Special Issue Perovskite Materials and Devices: Past, Present and Future)

Abstract

:
All-inorganic perovskite materials (such as CsPbBr3) have received widespread attention because of their better stability than hybrid counterparts, but their poor film morphology and crystalline quality limit their application in perovskite light-emitting devices (PeLEDs). Some previous studies have attempted to improve the morphology and crystalline quality of perovskite films by heating the substrate, but there are still some problems such as inaccurate temperature control, excessive temperature is not conducive to flexible applications, and the mechanism of action is not clear. In this work, we used a one-step spin-coating, low-temperature in situ thermally assisted crystallization process, in which the temperature was accurately monitored using a thermocouple in the range of 23–80 °C, and explored the effect of the in situ thermally assisted crystallization temperature on the crystallization of the all-inorganic perovskite material CsPbBr3 and the performance of PeLEDs. In addition, we focused on the influence mechanism for the in situ thermally assisted crystallization process on the surface morphology and phase composition of the perovskite films and promote its possible application in inkjet printing and scratch coating methods.

1. Introduction

Metal halide perovskites have attracted increasing attention due to their exceptional optoelectronic properties such as high charge carrier mobility [1,2,3], narrow emission linewidth [4,5], readily tunable emission wavelengths [6,7], and low cost of solution preparation, which make them ideal for use in optoelectronic applications, including solar cells [8], light-emitting diodes (LEDs) [9,10,11], photodetectors [12], and lasers [13]. Since 2014, PeLEDs have enabled a major breakthrough, especially in the external quantum efficiency (EQE) for organic–inorganic hybrid lead bromide, green-emitting perovskite LEDs, which has recently exceeded 20% [14]. Unfortunately, hybrid perovskites, which contain small organic cations such as methylammonium (MA) or formamidinium (FA), are extremely moisture-sensitive, which leads to rapid degradation in LED performance and thus limits their prospects for practical applications [15,16,17]. Perovskites based on inorganic cesium cations, namely CsPbX3 (X = Cl, Br, and I), exhibit better thermal and chemical stability compared to their hybrid analogs, and may thus provide the base for high-performance LEDs with reasonable operational stability [18,19].
Currently, several different methods have been developed for fabricating dense CsPbBr3 perovskite films with uniform thickness, such as solution processing methods including one-step spin-coating [20,21], two-step spin-coating [22,23,24,25,26,27], spray coating [28,29,30,31,32], and ink-jet printing [33,34]. In these methods, substrate heating has some preliminary applications in the current research related to perovskite thin films. In 2015, the team of Prof. Aditya D. Mohite at Los Alamos National Laboratory reported the preparation of perovskite films with very large grain sizes using a hot-cast solution deposition process for the first time [35]. In contrast to the conventional spin-coating process for film formation followed by annealing, the hot-cast solution deposition technique coats the heated precursor solution directly onto the substrate at 180 °C. Since the in situ-assisted thermal crystallization temperature is much higher than the perovskite crystallization temperature, and the presence of excess solvent allows the perovskite crystals to grow for a long time, producing a large number of large size crystal particles. However, the process requires a solvent with a high boiling point, thus limiting its application on flexible devices. In 2016, on this basis, Prof. Erik M. J. Johansson’s team at Uppsala University reported a method for preparing perovskite solar cells in air [36]. The researchers spin-coated a heated precursor solution onto a preheated substrate under atmospheric conditions and annealed it to obtain a well-crystallized perovskite film. The results showed that the step of preheating the substrate has an effect on the thickness as well as the crystallinity of the perovskite films, but the substrate preheating temperature above 80 °C leads to poor homogeneity in the perovskite films. The effect of substrate temperature on the crystallization process for perovskite thin films in the one-step method was reported by Prof. Xiaobin Liu’s group at the University of Electronic Science and Technology in 2019 [37]. To accurately control the substrate temperature, the researchers heated the substrate to 150 °C and then quickly spin-coated the solution onto the substrate when it cooled to a predetermined temperature. They found that when the substrate temperature was controlled within a certain range, a highly crystalline perovskite layer could be formed. However, due to the experimental conditions, the temperature of the substrate continued to decrease during the spin-coating process, which affected the accuracy of the experimental analysis. Based on the investigations mentioned above, our team reported that low-temperature in situ dynamic thermal crystallization achieves high-performance CsPbBr3-based PeLED. We conducted the annealing process while evaporating the precursor (CsBr and PbBr2), and the fabricated devices demonstrated superior performance [38]. Furthermore, the EQE of green PeLED was over 26%, which implied the great potential for the commercializatioin [39].
In this work, we applied the in situ thermally assisted crystallization process to a one-step spin-coating method. The effect of in-situ preparation temperature on the crystallization of CsPbBr3 perovskite films during one-step preparation and the performance of PeLEDs were investigated. The substrate was heated continuously while spin-coating the perovskite film, and the temperature was accurately monitored using a thermocouple in the range of 23–80 °C. Using an analysis of the morphology and properties of CsPbBr3 films at different temperatures, we focus on the influence mechanism of the in situ thermally assisted crystallization process on the surface morphology and phase composition of the perovskite film and promote its possible application in inkjet printing and scratch coating methods.

2. Experimental Section

Materials Preparation. Cesium bromide (CsBr, 99.9%), lead bromide (PbBr2, 99.99%), and pre-patterned, indium-doped tin-oxide (ITO) substrates were purchased from Advanced Election Technology Co., Ltd. (Beijing, China), and used without further purification. PEDOT: PSS, 1,3,5-Tris(1-phenyl-1H-benzimidazol-2-yl) benzene synonym (TPBI, 99.5%), 8-hydroxyquinolinolato-lithium (Liq, 99.5%), and aluminum (Al) were purchased from Xi’an Polymer Light Technology Corp. (Xi’an, China) and used without further purification.
Device Fabrication. The ITO-coated substrates (25 mm × 25 mm) were cleaned using sequential sonication in acetone, ethanol, and deionized water for 15 min. in each solvent and then dried in an oven at 100 °C for 30 min. After 20 min UV ozone treatment for ITO, a PEDOT:PSS aqueous solution was spin-coated onto the ITO substrate using a one-step process (4000 rpm for 30 s) and then baked at 150 °C for 20 min in a dry environment. The specific process flow for the one-step spin-coating in situ thermally assisted crystallization process for CsPbBr3 is described in Figure S1. The substrates were loaded into a high-vacuum chamber (base pressure ≈ 2 × 10−6 torr) to deposit TPBI (50 nm), Liq (1 nm), and Al (100 nm) layer-by-layer. The substrates were kept at room temperature with the sample holder rotating at 30 rpm. All the functional materials were also sequentially deposited using thermal evaporation and growth at a standard deposition rate of 0.6 Å/s without substrate heating.
Characterizations. The CsPbBr3 films and PeLEDs were tested in the air without encapsulation. The thickness of each material layer was determined using an oscillating quartz thickness monitor (INFICON SQC-310C) located in the vacuum chamber to determine the deposition rate. Scanning electron microscopy (SEM) was conducted using a JSM-7500F field-emission scanning electron microscope (JEOL) to observe the morphology of the CsPbBr3 films. The surface roughness of the CsPbBr3 films was measured with atomic force microscopy (AFM) using NX10 atomic force microscopy (Park). X-ray diffraction (XRD) was conducted using a model D/max 2400 X-ray diffraction (Bruker). Photoluminescence (PL) spectra were obtained using an RF-5301PC fluorescence spectrophotometer (SHIMADZU). Electroluminescence (EL) spectra and CIE colorimetry values were collected using a PR655 SpectraScan spectroradiometer (Photo Research). The current-voltage luminance characteristic curves were obtained using a B290222A precision source/measure unit (Agilent, Santa Clara, CA, USA) in the air at room temperature.

3. Result and Discussion

The color change in the film is the most visual representation of the film’s surface. Photographs showing the CsPbBr3 films prepared at different in situ heat-assisted crystallization temperatures are shown in Figure S2. When the in situ thermally assisted crystallization temperature was increased, the color of the unannealed CsPbBr3 film gradually deepened from light to earthy, and the light transmission almost changed from fully transparent to semitransparent. As the in situ dynamic thermal crystallization temperature continued to rise above 70 °C, the CsPbBr3 film increased in haze and assumed a frosted state. The reason for the color change here may come from changes in the thickness of the CsPbBr3 film and the presence of derivatives of CsPbBr3 (e.g., CsPb2Br5 and Cs4PbBr6). In addition, the increase in film haze when the in situ heat-assisted crystallization temperature rises above 70 °C may originate from the change in surface roughness.
To analyze the effect of in situ thermally assisted crystallization on the film morphology, we characterized the surface roughness of the unannealed CsPbBr3 films using atomic force microscopy (AFM). Figure 1a–e shows the height AFM images of unannealed CsPbBr3 films prepared at different in situ thermally assisted crystallization temperatures with an image size of 10 × 10 μm. When the in situ thermally assisted crystallization temperature was increased from room temperature to 40 °C, the surface roughness of the films decreased from 4.1 nm to 3.1 nm and showed a dendritic morphology. Therefore, the lower in situ thermally assisted crystallization temperature may not accelerate solvent volatilization to influence the crystallization dynamic. In addition, when the temperature increased to 40 °C, the dendritic morphology on the surface of CsPbBr3 films almost disappeared, which indicated that the in situ thermally assisted crystallization temperature of 40 °C improved the agglomeration phenomenon and made the films become flatter and further reduced the surface defects sites of CsPbBr3 films. With the increase in temperature to 60 °C and 70 °C (as shown in Figure 1c,d), it can be seen more obviously that the pin-holes on the surface of the films increased rapidly and that island-like morphology started to be generated and increased significantly with the increase in temperature, while the surface roughness of the unannealed CsPbBr3 films increased from 3.1 nm to 4.6 nm. This implies that although the in situ thermally assisted crystallization process can accelerate the solvent volatilization to reduce perovskite agglomeration, once the evaporation rate is too fast, it will in turn lead to the increase in pin-holes on the films. These pin-holes existing among the PeLED devices may cause direct contact between the cathode and anode. Meanwhile, like grain boundaries, the pin-holes may contain many handing bonds acting as defect sites, which could cause serious nonradiative recombination. When the in situ thermally assisted crystallization temperature was further increased to 80 °C, the island-like morphology on the surface of the unannealed CsPbBr3 film became more obvious, accompanied by an enhancement in Ra, which may be due to the fact that the in situ thermally assisted crystallization temperature of 80 °C greatly accelerated the volatilization rate of the DMSO solvent, at which time the CsPbBr3 rapidly pre-crystallized around the nucleation site under the unannealed crystallization condition.
Figure 1a1–e1 shows that the AFM of CsPbBr3 films after annealing at different in situ thermally assisted crystallization temperatures provide the morphology of the annealed CsPbBr3 films. When the in situ thermally assisted crystallization temperature is below 60 °C, the surface of the films is relatively flat with Ra not exceeding 10 nm, while when the in situ thermally assisted crystallization temperature exceeds 70 °C, the surface of the CsPbBr3 films shows bumps of up to several hundred nanometers with a diameter of about 1 μm. It is also noted that when the in situ thermally assisted crystallization temperature rises to 80 °C, the bumps are denser, and the surface roughness of the film is larger. The generation of island-like morphology in perovskite electroluminescent devices can directly lead to contact between the perovskite layer and the electrode, resulting in carrier tunneling and destruction of the device. In addition, the laser confocal micrographs showing the unannealed and annealed CsPbBr3 films (Figure S4) at different in situ thermally assisted crystallization temperatures also show similar trends, which further support our conclusion.
In Figure S3, we summarize the surface roughness variation trend in CsPbBr3 films after annealing. It is found that the change in surface roughness of the films is small when the in situ thermally assisted crystallization temperature is below 60 °C. The in situ thermally assisted crystallization temperature above 60 °C accelerates the volatilization rate of the DMSO solvent, which leads to the growth of the films around the nucleation as sites. Therefore, for CsPbBr3 films prepared using DMSO solvent, an in situ thermally assisted crystallization temperature of 60 °C is the critical temperature for obtaining smooth CsPbBr3 films. Combining Figure 1b1,c1 to compare the surface morphology of the films prepared at the in situ thermally assisted crystallization temperatures of 40 °C and 60 °C, we can see that the in situ thermally assisted crystallization temperature of 40 °C results in lower surface grain boundary density, a lower density of defect states, and fewer pin-holes in the CsPbBr3 films, which may theoretically be the key to obtaining higher performance LEDs. The annealing temperature plays an important role in the variation in the morphology of the perovskite film. The elevated temperature may accelerate the evaporation of the solvent to influence the dynamics of crystallization. So, we attributed the disappearance of pin-holes to the difference in the annealing temperature.
According to the previous section, the surface morphology of CsPbBr3 films using the in situ thermally assisted crystallization process was improved, and the defect sites were significantly reduced. To further understand the reasons for this, the structure of the CsPbBr3 films prepared using the in situ thermally assisted crystallization process was analyzed.
Firstly, the unannealed and annealed CsPbBr3 films prepared at different in situ thermally assisted crystallization temperatures were characterized using X-ray diffraction (XRD) as shown in Figure 2a,b, respectively. As shown in Figure 2a, the intensity of the diffraction peak at the 2θ angle of 30.5°, which corresponds to the (200) crystal face of CsPbBr3, gradually increases when increasing the in situ thermally assisted crystallization temperature of the unannealed CsPbBr3 film. When the in situ thermally assisted crystallization temperature is above 60 °C, two diffraction peaks at 15.2° and 21.4° appear, corresponding to the (100) and (110) crystallographic planes of CsPbBr3, respectively. At the same time, the intensity of these two diffraction peaks increased with the increase in temperature. When the temperature rises to 70 °C and above, the films show diffraction peaks of Cs4PbBr6 at 12.7°, 22.4° and 28.6°. This is probably due to the fact that at a certain temperature, the unreacted CsBr in the precursor solution reacts with the CsPbBr3 first produced during the spin-coating process to form Cs4PbBr6. The reaction principle is shown in Equation (1). The increase in surface roughness of the films may therefore come from the production of Cs4PbBr6.
CsPbBr3 + 3CsBr = Cs4PbBr6
Since Cs4PbBr6 is easily converted to CsPbBr3 with thermal instability, we characterized the annealed CsPbBr3 films using XRD. The CsPbBr3 diffraction peaks are at 12.7° and 30.5°, which gradually increase in intensity as the temperature increases. The diffraction peak in Cs4PbBr6 disappears at 22.4°, which is mainly due to the conversion of Cs4PbBr6 into CsPbBr3 during annealing. It is noteworthy that the diffraction peak at 21.4° corresponding to the CsPbBr3(110) crystal plane increases in intensity as the temperature rises from 23 °C to 40 °C, while as the temperature rises from 40 °C to 80 °C, the intensity of the diffraction peak at the same position decreases instead. The CsPbBr3(220) crystal plane diffraction peak at 43.8° has almost the same behavior. Related studies suggest that CsPbBr3 films grown along the (110) direction have better EL properties [38]. Therefore, we found that the in situ thermally assisted crystallization process has a strong influence on the crystallographic orientation of the crystals in the one-step preparation of perovskite films and that high in situ thermally assisted crystallization temperatures for derivative-prone perovskite materials can lead to the formation of derivatives.
To verify that the island-like morphology is due to the Cs4PbBr6 crystals, we tested the Raman spectra of the corresponding positions. Figure 3a shows a Raman micrograph for in situ thermally assisted crystallization at a temperature of 80 °C. The red circles show the island-like morphology described previously, and the black circles show the non-island-like morphology. The island-like (red circle) and non-island-like (black circle) features in Figure 3b correspond to the characteristic Raman peaks in CsPbBr3 and Cs4PbBr6, respectively. There are three more significant Raman peaks for CsPbBr3 in the figure, a weaker shoulder peak at 56 cm−1 and two weak broad peaks at 126 cm−1 and 310 cm−1, where 56 cm−1 and 126 cm−1 correspond to the vibrational mode of the [PbBr6]4− normal octahedron and the motion mode of the Cs+ ion, and 310 cm−1 corresponds to the second-order phonon model for the normal octahedron. In contrast, Cs4PbBr6 has strong characteristic Raman peaks at 70 cm−1, 83 cm−1, and 126 cm−1, corresponding to the [PbBr6]4− ortho-octahedral mode of vibration. Notably, the disappearance of the theoretical strongest characteristic peak position in CsPbBr3 at 72 cm−1, as shown in the Figure, thus proves that the bulge on the film surface is produced by both CsPbBr3 and Cs4PbBr6. Considering the temperature sensitivity of Cs4PbBr6, the characteristic Raman peak in CsPbBr3 at 56 cm−1, shown in the red sample bar in Figure 3b, may come from the partial conversion of Cs4PbBr6 into CsPbBr3.
The films prepared at different in situ thermally assisted crystallization temperatures were also characterized using Raman spectroscopy, as shown in Figure 3c. Only when the temperature reached 80 °C did the characteristic Raman peak in Cs4PbBr6 appear. This also demonstrates that the in situ thermally assisted crystallization temperature accelerates the volatilization rate of DMSO, but the excessively fast crystallization rate leads to preferential growth of CsPbBr3 at some locations on the substrate surface and results in the production of CsPbBr3 derivatives on this basis. The results of our energy spectroscopy (EDS) analysis of the island-like morphology on the film surface are shown in Figure 3d. We have carried out an EDS surface scan analysis of the island-like formations at the red squares, and it can be seen from Table S1 that the atomic percentages of Cs, Pb, and Br are 35.36%, 8.22%, and 56.42%, respectively, with elemental ratios close to 4:1:6, which is a ratio consistent with the elemental ratio in the Cs4PbBr6 molecular formula. The island-like morphology is more likely to be Cs4PbBr6.
To further demonstrate that the in situ thermally assisted crystallization process can suppress defect states in perovskite films, steady-state fluorescence (PL) spectra, as well as transient fluorescence (TRPL) spectra, were tested, both at room temperature with an excitation light wavelength of 365 nm. Figure 4a shows the steady-state normalized photoluminescence spectra of the films at different in situ thermally assisted crystallization temperatures. As can be seen in the figure, when the in situ thermally assisted crystallization temperature was below 40 °C, the films exhibited a PL peak in CsPbBr3 at 525 nm, and the full width at half maxima (FWHM) was narrower at 40 °C compared to the films prepared at 23 °C, indicating better crystallinity of the CsPbBr3 films at 40 °C. When the temperature rises to 60 °C, the PL peak in the film shows a significant red-shift, which, for perovskite films, is related to surface defect states and grain boundaries. Moreover, when the in situ thermally assisted crystallization temperature was higher than 60 °C, the crystallization of the prepared perovskite films was worse, which is consistent with the results discussed earlier. When the temperature increased to 80 °C, a PL luminescence peak at 514 nm Cs4PbBr6 can be found in the film, together with a shoulder peak at 522 nm. The appearance of the blue-shifted shoulder peak here is related to the quantum-limited effect. The reaction of CsPbBr3 with CsBr leads to a reduction in the crystal size of CsPbBr3 nested in the crystals of Cs4PbBr6 [40]. This supports the view that the characteristic CsPbBr3 Raman peaks in the Raman spectral analysis in the previous section may be due to the partial conversion of Cs4PbBr6 into CsPbBr3. Figure 4b shows the steady-state photoluminescence spectra of the films at in situ thermally assisted crystallization temperatures of 23 °C, 40 °C, and 60 °C. It can be seen that the PL intensity at 40 °C is the strongest, even higher than that at 60 °C where the film thickness is thicker, which proves that the films prepared using the in situ thermally assisted crystallization process at 40 °C have fewer defect sites and higher crystalline quality [41].
In addition to the degree of crystallinity and crystal orientation in the CsPbBr3 films grown at different temperatures, the nature of defects in the film is also a critical feature that affects the properties of the photo-generated carriers and PeLEDs performance with the device structure of ITO/PEDOT:PSS/perovskite. Therefore, the trap states of the CsPbBr3 films were investigated by evaluating their time-resolved photoluminescence (TRPL), which is presented in Figure 5. In addition, the TRPL curves in Figure 5 were fitted using biexponential decay functions, and the average decay times for the two components in each of the biexponential decay functions included a relatively short lifetime τ1 with an area under the TRPL curve of A1 and a longer lifetime τ2 with an area under the TRPL curve of A2. The average decay time τave was then calculated using the following expression:
τ a v e = A 1 τ 1 2 + A 2 τ 2 2 A 1 τ 1 + A 2 τ 2
The charge carrier lifetimes obtained for the CsPbBr3 films grown at RT, 40 °C, and 60 °C are listed in Table 1. The TRPL spectra show two characteristic decays. The faster decay, associated with lifetime τ1, is related to trap-assisted non-radiative recombination by defects sites in the perovskite film, especially on the surface of perovskite grains with a higher defect concentration. The slower decay, associated with lifetime τ2, is related to radiation recombination of the excited states inside the perovskite grains. Compared with room temperature, the carrier lifetime under the in situ thermally assisted crystallization temperature of 40 °C is slightly improved, where τ1 increased from 1.57 ns to 1.59 ns and τ2 increased from 13.05 ns to 14.03 ns. The increase in τ1 indicates that defect states on the surface of the CsPbBr3 film are reduced, resulting in less non-radiation recombination. On the other hand, the increase in τ2 demonstrates a longer carrier lifetime and indicates the improved quality of the CsPbBr3 film. However, when the in situ thermally assisted crystallization temperature is 60 °C, the carrier lifetime will be significantly decreased, where τ1 decreased from 1.59 ns to 1.30 ns and τ2 decreased from 14.03 ns to 8.38 ns. This shows that when the in situ heat-assisted crystallization temperature is too high, the defect density increases, leading to the internal radiation recombination decrease, and the quality of the CsPbBr3 film relatively decreases. Overall, the average lifetime of the CsPbBr3 film prepared at the 40 °C in situ thermally assisted crystallization temperature is 5.87 ns, which is significantly longer than the 4.41 ns lifetime of the sample prepared at 23 °C. This result can be explained by the observed reduction in grain boundary trap-assisted recombination. Meanwhile, the results for PLQY in Figure S5 were inconsistent with the TRPL measurement.
In order to gain a deeper understanding of the reasons for the enhancement in the PLQY of films prepared using the in situ thermally assisted crystallization process, here we use the space-charge-limited current (SCLC) method to quantify and compare the changes in defects density with and without the in situ thermally assisted crystallization processes. We prepared a single-hole carrier device with the structure ITO/PEDOT:PSS/CsPbBr3/CBP/MoO3/Ag.
From the I-V curve in Figure 6, three different areas can be clearly identified, including Ohmic, Trap Filled Limit, and SCLC. We calculated the hole and electron defect state density (Nt) for the CsPbBr3 film using a single-hole carrier device. The calculation method is carried out according to the following formula:
N t = 2 ε ε 0 V T F L eL 2  
where ε and ε0 represent the relative permittivity and vacuum permittivity of the perovskite, respectively (8.854 × 10−12 F/m), VTFL is the defect limit voltage (the defect state is completely restricted within this voltage range), and L represents film thickness, where the thickness of the CsPbBr3 film is obtained using a step meter test and e represents the elementary charge. The thickness of the CsPbBr3 film at different in situ heat-assisted crystallization temperatures is shown in Table S1. Therefore, we can calculate the defect state density for the perovskite film prepared at room temperature and the 40 °C in situ thermally assisted crystallization temperature as: 3.59 × 1016 cm−3 and 2.3 × 1016 cm−3, respectively. The defect state density for the film at the in situ heat-assisted crystallization temperature of 40 °C is significantly reduced.
By characterizing the electroluminescence properties of CsPbBr3 PeLEDs prepared under different in situ thermally assisted crystallization temperatures using one-step spin-coating, we can further analyze the in situ thermally assisted crystallization temperature and its influence on the properties of CsPbBr3 films. Figure 7a shows the current density–voltage characteristic curve for the electroluminescent device at RT and the 40 °C and 60 °C in situ thermally assisted crystallization temperatures. We found that when the in situ thermally assisted crystallization temperature is 40 °C, with the same operating voltage, the device exhibits the highest current density. Combined with our previous research results on the effect of in situ thermally assisted crystallization temperature on the morphology of CsPbBr3 films, it is further proved that CsPbBr3 films prepared using the low-temperature in situ thermally assisted crystallization process have fewer defects. When the same number of carriers are injected, the proportion of non-radiative recombination generated when filling defects is less. When the in situ heat-assisted crystallization temperature is 40 °C, its thickness is not much different from that of the CsPbBr3 film prepared at RT. Therefore, compared with the film prepared at 60 °C, the radiation recombination rate of the device is improved with the limitation in the carrier space, so it shows a higher current density. Figure 7b shows the current efficiency–voltage characteristic curve of PeLEDs at RT and the 40 °C and 60 °C in-situ thermally assisted crystallization temperatures. From the figure, we can see that at 40 °C, the current efficiency of the device is as high as 9.9 cd/A, which is almost 3 times (RT, 3.1 cd/A) that seen when the process method is not used. In Figure 7c, we can see that when the driving voltage is only 5 V at 40 °C, the device has reached a brightness of 6208 cd/m2, while it is only 1829 cd/m2 at RT. This further proves that the low-temperature in-situ thermal-assisted crystallization technology improves the surface morphology of the CsPbBr3 film, reduces surface defects, and therefore reduces the non-radiative recombination rate. Figure 7d shows the electroluminescence spectra for CsPbBr3 electroluminescent devices prepared at RT and the 40 °C and 60 °C in situ thermally assisted crystallization temperatures. All spectra show the equivalent luminescence peak position at 520 nm, which corresponds to the CIE color coordinate (0.11, 0.77). It shows that the low-temperature in-situ heat-assisted crystallization process has almost no effect on the position of the emission peak in the CsPbBr3 electroluminescent device. Meanwhile, the EQE measurement shown in Figure S6 also proved that the 40 °C in situ thermally assisted crystallization temperature delivered the best-performance device.

4. Conclusions

In this manuscript, we systematically investigated the effect of a one-step spin-coating method combined with the in situ thermally assisted crystallization process on the defect density for states of CsPbBr3. The results confirm that the surface defect density for states of CsPbBr3 films can be reduced using a low temperature, in situ thermally assisted crystallization process, which reduces the non-radiative composite rate. On this basis, we prepared CsPbBr3 light-emitting devices that can reach a maximum brightness of 6208 cd/m2 at 5 V. A maximum current efficiency of 9.9 cd/A can be obtained at the same time. Furthermore, for the light-emitting devices without the low temperature in situ thermally assisted crystallization process, an almost 3-fold increase in brightness and current efficiency was demonstrated. This result is one of the highest efficiencies ever achieved for an all-inorganic CsPbBr3 light-emitting device without additives.
Compared with the previous high-temperature, in situ thermal preparation process often used in the spin-coating process, we demonstrate that the introduced low-temperature, in situ thermally assisted crystallization process can effectively enhance the optoelectronic properties of PeLEDs without other adjacent functional layers. With flexible screen devices being mainstream for future display and lighting devices, the low-temperature, in-situ thermally assisted crystallization process proposed in this work further reduces the requirement for substrate temperature resistance performance and provides a new idea for the solution-based preparation of perovskite films.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/mi14061104/s1, Figure S1. The fabrication process of one-step spin-coating in-situ thermally assisted crystallization process. Figure S2. The color variation of perovskite film at different in situ heat-assisted crystallization temperatures. Figure S3. The roughness variation of perovskite film under different annealing temperature. Figure S4. The laser confocal micrographs of unannealed and annealed CsPbBr3 films at different in situ thermally assisted crystallization temperatures. Figure S5. PLQY measurement of the perovskite film under different annealing tem-perature. Figure S6. EQE measurement for the best performance PeLED device. Table S1. The concentration of elements in the perovskite film (red circle).

Author Contributions

Supervision: X.C. Methodology: C.C. Investigation: B.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the National Natural Science Foundation of China under Grant No. 11904031.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare that there is no conflict of interest.

References

  1. Wehrenfennig, C.; Eperon, G.E.; Johnston, M.B.; Snaith, H.J.; Herz, L.M. High Charge Carrier Mobilities and Lifetimes in Organolead Trihalide Perovskites. Adv. Mater. 2014, 26, 1584–1589. [Google Scholar] [CrossRef] [PubMed]
  2. Zeng, J.P.; Li, X.M.; Wu, Y.; Yang, D.D.; Sun, Z.G.; Song, Z.H.; Wang, H.; Zeng, H.B. Space-Confined Growth of CsPbBr3 Film Achieving Photodetectors with High Performance in All Figures of Merit. Adv. Funct. Mater. 2018, 28, 8. [Google Scholar] [CrossRef]
  3. Bao, C.; Chen, C.; Muhammad, M.; Ma, X.J.; Wang, Z.K.; Liu, Y.F.; Chen, P.; Chen, S.M.; Liu, B.; Wang, J.T.; et al. Hybrid perovskite charge generation layer for highly efficient tandem organic light-emitting diodes. Org. Electron. 2019, 73, 299–303. [Google Scholar] [CrossRef]
  4. Yuan, M.J.; Quan, L.N.; Comin, R.; Walters, G.; Sabatini, R.; Voznyy, O.; Hoogland, S.; Zhao, Y.B.; Beauregard, E.M.; Kanjanaboos, P.; et al. Perovskite energy funnels for efficient light-emitting diodes. Nat. Nanotechnol. 2016, 11, 872. [Google Scholar] [CrossRef]
  5. Xiao, Z.G.; Kerner, R.A.; Zhao, L.F.; Tran, N.L.; Lee, K.M.; Koh, T.W.; Scholes, G.D.; Rand, B.P. Efficient perovskite light-emitting diodes featuring nanometre-sized crystallites. Nat. Photonics 2017, 11, 108–115. [Google Scholar] [CrossRef]
  6. Sichert, J.A.; Tong, Y.; Mutz, N.; Vollmer, M.; Fischer, S.; Milowska, K.Z.; Cortadella, R.G.; Nickel, B.; Cardenas-Daw, C.; Stolarczyk, J.K.; et al. Quantum Size Effect in Organometal Halide Perovskite Nanoplatelets. Nano Lett. 2015, 15, 6521–6527. [Google Scholar] [CrossRef] [PubMed]
  7. Wang, H.R.; Li, X.M.; Yuan, M.J.; Yang, X.Y. Fast Postmoisture Treatment of Luminescent Perovskite Films for Efficient Light-Emitting Diodes. Small 2018, 14, 7. [Google Scholar] [CrossRef]
  8. Wang, H.X.; Lei, J.; Gao, F.; Yang, Z.; Yang, D.; Jiang, J.X.; Li, J.; Hu, X.H.; Ren, X.D.; Liu, B.; et al. Magnetic Field-Assisted Perovskite Film Preparation for Enhanced Performance of Solar Cells. ACS Appl. Mater. Interfaces 2017, 9, 21756–21762. [Google Scholar] [CrossRef]
  9. Chen, Y.M.; Zhou, Y.; Zhao, Q.; Zhang, J.Y.; Ma, J.P.; Xuan, T.T.; Guo, S.Q.; Yong, Z.J.; Wang, J.; Kuroiwa, Y.; et al. Cs4PbBr6/CsPbBr3 Perovskite Composites with Near-Unity Luminescence Quantum Yield: Large-Scale Synthesis, Luminescence and Formation Mechanism, and White Light-Emitting Diode Application. ACS Appl. Mater. Interfaces 2018, 10, 15905–15912. [Google Scholar] [CrossRef]
  10. Chen, C.; Wu, D.; Yuan, M.; Yu, C.; Zhang, J.; Li, C.N.; Duan, Y. Spectroscopic ellipsometry study of CsPbBr3 perovskite thin films prepared by vacuum evaporation. J. Phys. D-Appl. Phys. 2021, 54, 8. [Google Scholar] [CrossRef]
  11. Chen, S.M.; Chen, C.; Bao, C.; Mujahid, M.; Li, Y.; Chen, P.; Duan, Y. White Light-Emitting Devices Based on Inorganic Perovskite and Organic Materials. Molecules 2019, 24, 800. [Google Scholar] [CrossRef] [PubMed]
  12. Dou, L.T.; Yang, Y.; You, J.B.; Hong, Z.R.; Chang, W.H.; Li, G.; Yang, Y. Solution-processed hybrid perovskite photodetectors with high detectivity. Nat. Commun. 2014, 5, 6. [Google Scholar] [CrossRef]
  13. Veldhuis, S.A.; Boix, P.P.; Yantara, N.; Li, M.J.; 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]
  14. Lin, K.B.; Xing, J.; Quan, L.N.; de Arquer, F.P.G.; Gong, X.W.; Lu, J.X.; Xie, L.Q.; Zhao, W.J.; Zhang, D.; Yan, C.Z.; et al. Perovskite light-emitting diodes with external quantum efficiency exceeding 20 per cent. Nature 2018, 562, 245. [Google Scholar] [CrossRef] [PubMed]
  15. Yang, J.L.; Siempelkamp, B.D.; Liu, D.Y.; Kelly, T.L. Investigation of CH3NH3PbI3 Degradation Rates and Mechanisms in Controlled Humidity Environments Using in Situ Techniques. ACS Nano 2015, 9, 1955–1963. [Google Scholar] [CrossRef]
  16. Liu, H.W.; Wu, Z.N.; Shao, J.R.; Yao, D.; Gao, H.; Liu, Y.; Yu, W.L.; Zhang, H.; Yang, B. CsPbxMn1−xCl3 Perovskite Quantum Dots with High Mn Substitution Ratio. ACS Nano 2017, 11, 2239–2247. [Google Scholar] [CrossRef]
  17. Niu, G.D.; Guo, X.D.; Wang, L.D. Review of recent progress in chemical stability of perovskite solar cells. J. Mater. Chem. A 2015, 3, 8970–8980. [Google Scholar] [CrossRef]
  18. Song, J.Z.; Li, J.H.; Li, X.M.; Xu, L.M.; Dong, Y.H.; Zeng, H.B. Quantum Dot Light-Emitting Diodes Based on Inorganic Perovskite Cesium Lead Halides (CsPbX3). Adv. Mater. 2015, 27, 7162. [Google Scholar] [CrossRef]
  19. Zhang, L.Q.; Yang, X.L.; Jiang, Q.; Wang, P.Y.; Yin, Z.G.; Zhang, X.W.; Tan, H.R.; Yang, Y.; Wei, M.Y.; Sutherland, B.R.; et al. Ultra-bright and highly efficient inorganic based perovskite light-emitting diodes. Nat. Commun. 2017, 8, 8. [Google Scholar] [CrossRef]
  20. Wang, Z.B.; Cheng, T.; Wang, F.Z.; Dai, S.Y.; Tan, Z.A. Morphology Engineering for High-Performance and Multicolored Perovskite Light-Emitting Diodes with Simple Device Structures. Small 2016, 12, 4412–4420. [Google Scholar] [CrossRef]
  21. Song, L.; Huang, L.X.; Hu, Y.S.; Guo, X.Y.; Liu, X.Y.; Geng, C.; Xu, S.; Luan, N.N.; Bi, W.G.; Wang, L.S. Synergistic morphology control and non-radiative defect passivation using a crown ether for efficient perovskite light-emitting devices. J. Mater. Chem. C 2020, 8, 9986–9992. [Google Scholar] [CrossRef]
  22. Zhou, Z.X.; Zou, X.P.; Zhu, J.L.; Cheng, J.; Ren, H.Y.; Chang, C.C.; Yao, Y.J.; Chen, D.; Yu, X.; Li, G.D.; et al. Effect of Processing Technique Factors on Structure and Photophysical Properties of Perovskite Absorption Layer by Using Two-Step Spin-Coating Method. Crystals 2020, 10, 761. [Google Scholar] [CrossRef]
  23. More, S.A.; Halor, R.; Raees, S.; Ghosh, S.S. A comparative study of two-step and three-step methods for coating organometallic lead halide perovskite thin films. J. Mater. Sci. Mater. Electron. 2020, 31, 17995–18003. [Google Scholar] [CrossRef]
  24. Li, S.N.; Ren, H.; Yan, Y. Boosting efficiency of planar heterojunction perovskite solar cells to 21.2% by a facile two-step deposition strategy. Appl. Surf. Sci. 2019, 484, 1191–1197. [Google Scholar] [CrossRef]
  25. Ip, A.H.; Quan, L.N.; Adachi, M.M.; McDowell, J.J.; Xu, J.X.; Kim, D.H.; Sargent, E.H. A two-step route to planar perovskite cells exhibiting reduced hysteresis. Appl. Phys. Lett. 2015, 106, 5. [Google Scholar] [CrossRef]
  26. Hongsith, K.; Wongrerkdee, S.; Ngamjarurojana, A.; Choopun, S. Efficiency enhancement of perovskite solar cell by using pre-heat treatment in two-step deposition method. Thin Solid Films 2019, 684, 9–14. [Google Scholar] [CrossRef]
  27. Xu, Y.Z.; Zhu, L.F.; Shi, J.J.; Lv, S.T.; Xu, X.; Xiao, J.Y.; Dong, J.; Wu, H.J.; Luo, Y.H.; Li, D.M.; et al. Efficient Hybrid Mesoscopic Solar Cells with Morphology-Controlled CH3NH3Pbl3-xClx Derived from Two-Step Spin Coating Method. ACS Appl. Mater. Interfaces 2015, 7, 2242–2248. [Google Scholar] [CrossRef]
  28. Liang, Z.R.; Zhang, S.H.; Xu, X.Q.; Wang, N.; Wang, J.X.; Wang, X.; Bi, Z.N.; Xu, G.; Yuan, N.Y.; Ding, J.N. A large grain size perovskite thin film with a dense structure for planar heterojunction solar cells via spray deposition under ambient conditions. RSC Adv. 2015, 5, 60562–60569. [Google Scholar] [CrossRef]
  29. Heo, J.H.; Lee, M.H.; Jang, M.H.; Im, S.H. Highly efficient CH3NH3PbI3-xClx mixed halide perovskite solar cells prepared by re-dissolution and crystal grain growth via spray coating. J. Mater. Chem. A 2016, 4, 17636–17642. [Google Scholar] [CrossRef]
  30. Bi, Z.N.; Liang, Z.R.; Xu, X.Q.; Chai, Z.S.; Jin, H.; Xu, D.H.; Li, J.L.; Li, M.H.; Xu, G. Fast preparation of uniform large grain size perovskite thin film in air condition via spray deposition method for high efficient planar solar cells. Sol. Energy Mater. Sol. Cells 2017, 162, 13–20. [Google Scholar] [CrossRef]
  31. Barrows, A.T.; Pearson, A.J.; Kwak, C.K.; Dunbar, A.D.F.; Buckley, A.R.; Lidzey, D.G. Efficient planar heterojunction mixed-halide perovskite solar cells deposited via spray-deposition. Energy Environ. Sci. 2014, 7, 2944–2950. [Google Scholar] [CrossRef]
  32. Su, Y.; Jiang, H.; Liu, Z.Q. A study on environment-friendly machining of titanium alloy via composite electrostatic spraying. Int. J. Adv. Manuf. Technol. 2020, 110, 1305–1317. [Google Scholar] [CrossRef]
  33. Liang, C.; Li, P.W.; Gu, H.; Zhang, Y.Q.; Li, F.Y.; Song, Y.L.; Shao, G.S.; Mathews, N.; Xing, G.C. One-Step Inkjet Printed Perovskite in Air for Efficient Light Harvesting. Sol. RRL 2018, 2, 9. [Google Scholar] [CrossRef]
  34. Yu, C.; Chen, C.; Wu, D.; Jiang, X.; Duan, Y. Research progress of inkjet printed perovskite optoelectronic devices. Chin. J. Liq. Cryst. Disp. 2021, 36, 158–175. [Google Scholar] [CrossRef]
  35. Nie, W.Y.; Tsai, H.H.; Asadpour, R.; Blancon, J.C.; Neukirch, A.J.; Gupta, G.; Crochet, J.J.; Chhowalla, M.; Tretiak, S.; Alam, M.A.; et al. High-efficiency solution-processed perovskite solar cells with millimeter-scale grains. Science 2015, 347, 522–525. [Google Scholar] [CrossRef]
  36. Sveinbjornsson, K.; Aitola, K.; Zhang, J.B.; Johansson, M.B.; Zhang, X.L.; Correa-Baena, J.P.; Hagfeldt, A.; Boschloo, G.; Johansson, E.M.J. Ambient air-processed mixed-ion perovskites for high-efficiency solar cells. J. Mater. Chem. A 2016, 4, 16536–16545. [Google Scholar] [CrossRef]
  37. Zhang, H.Y.; Zhao, C.; Li, D.; Guo, H.; Liao, F.Y.; Cao, W.; Niu, X.B.; Zhao, Y.Y. Effects of substrate temperature on the crystallization process and properties of mixed-ion perovskite layers. J. Mater. Chem. A 2019, 7, 2804–2811. [Google Scholar] [CrossRef]
  38. Chen, C.; Han, T.; Tan, S.; Xue, J.J.; Zhao, Y.P.; Liu, Y.F.; Wang, H.R.; Hu, W.; Bao, C.; Mazzeo, M.; et al. Efficient Flexible Inorganic Perovskite Light-Emitting Diodes Fabricated with CsPbBr3 Emitters Prepared via Low-Temperature in Situ Dynamic Thermal Crystallization. Nano Lett. 2020, 20, 4673–4680. [Google Scholar] [CrossRef]
  39. Wang, K.; Lin, Z.-Y.; Zhang, Z.; Jin, L.; Ma, K.; Coffey, A.H.; Atapattu, H.R.; Gao, Y.; Park, J.Y.; Wei, Z.; et al. Suppressing phase disproportionation in quasi-2D perovskite light-emitting diodes. Nat. Commun. 2023, 14, 397. [Google Scholar] [CrossRef]
  40. Quan, L.N.; Quintero-Bermudez, R.; Voznyy, O.; Walters, G.; Jain, A.; Fan, J.Z.; Zheng, X.L.; Yang, Z.Y.; Sargent, E.H. Highly Emissive Green Perovskite Nanocrystals in a Solid State Crystalline Matrix. Adv. Mater. 2017, 29, 6. [Google Scholar] [CrossRef]
  41. Chen, J.Z.; Xiong, Y.L.; Rong, Y.G.; Mei, A.Y.; Sheng, Y.S.; Jiang, P.; Hu, Y.; Li, X.; Han, H.W. Solvent effect on the hole-conductor-free fully printable perovskite solar cells. Nano Energy 2016, 27, 130–137. [Google Scholar] [CrossRef]
Figure 1. The AFM images showing the CsPbBr3 films which were prepared at different in situ thermally assisted crystallization temperatures. (ae) Unannealed and (a1e1) annealed.
Figure 1. The AFM images showing the CsPbBr3 films which were prepared at different in situ thermally assisted crystallization temperatures. (ae) Unannealed and (a1e1) annealed.
Micromachines 14 01104 g001
Figure 2. XRD patterns for the (a) unannealed and (b) annealed CsPbBr3 films prepared at different in situ thermally assisted crystallization temperatures.
Figure 2. XRD patterns for the (a) unannealed and (b) annealed CsPbBr3 films prepared at different in situ thermally assisted crystallization temperatures.
Micromachines 14 01104 g002
Figure 3. (a) Raman micrograph showing the surface of the CsPbBr3 film after annealing at an in situ thermally assisted crystallization temperature of 80 °C. The red circle shows the island-like morphology described above, and the black circle shows the non-island-like morphology. Respectively with (b), when the in situ thermally assisted crystallization temperature is 80 °C, the red line and the black line in the Raman spectrum for the film sample correspond; (c) are the Raman spectra for the CsPbBr3 film samples at different in situ heat-assisted crystallization temperatures; (d) are the SEM images showing the surface of the CsPbBr3 film. The red box is the scanning area of EDS mapping, where the EDS map is above (d).
Figure 3. (a) Raman micrograph showing the surface of the CsPbBr3 film after annealing at an in situ thermally assisted crystallization temperature of 80 °C. The red circle shows the island-like morphology described above, and the black circle shows the non-island-like morphology. Respectively with (b), when the in situ thermally assisted crystallization temperature is 80 °C, the red line and the black line in the Raman spectrum for the film sample correspond; (c) are the Raman spectra for the CsPbBr3 film samples at different in situ heat-assisted crystallization temperatures; (d) are the SEM images showing the surface of the CsPbBr3 film. The red box is the scanning area of EDS mapping, where the EDS map is above (d).
Micromachines 14 01104 g003
Figure 4. (a) Steady-state normalized photoluminescence spectra showing the CsPbBr3 films at different in situ thermally assisted crystallization temperatures. (b) Steady-state photoluminescence spectra showing the CsPbBr3 film at 23 °C, 40 °C, and 60 °C in situ thermally assisted crystallization.
Figure 4. (a) Steady-state normalized photoluminescence spectra showing the CsPbBr3 films at different in situ thermally assisted crystallization temperatures. (b) Steady-state photoluminescence spectra showing the CsPbBr3 film at 23 °C, 40 °C, and 60 °C in situ thermally assisted crystallization.
Micromachines 14 01104 g004
Figure 5. Time-resolved photoluminescence spectra showing the CsPbBr3 films at different in situ thermally assisted crystallization temperatures.
Figure 5. Time-resolved photoluminescence spectra showing the CsPbBr3 films at different in situ thermally assisted crystallization temperatures.
Micromachines 14 01104 g005
Figure 6. (a,b) The SCLC diagrams for the single-hole carrier device prepared using CsPbBr3 thin film at RT and the 40 °C in situ thermally assisted crystallization temperature, respectively.
Figure 6. (a,b) The SCLC diagrams for the single-hole carrier device prepared using CsPbBr3 thin film at RT and the 40 °C in situ thermally assisted crystallization temperature, respectively.
Micromachines 14 01104 g006
Figure 7. EL properties of CsPbBr3 prepared at RT and the 40 °C and 60 °C in situ thermally assisted crystallization temperatures. (a) Current density–voltage characteristic curve; (b) current efficiency–voltage characteristic curve; (c) brightness–voltage characteristic curve; and (d) normalized electroluminescence spectra for CsPbBr3 light emitting devices prepared at different in situ thermally assisted crystallization temperature work on 5 V driving voltage.
Figure 7. EL properties of CsPbBr3 prepared at RT and the 40 °C and 60 °C in situ thermally assisted crystallization temperatures. (a) Current density–voltage characteristic curve; (b) current efficiency–voltage characteristic curve; (c) brightness–voltage characteristic curve; and (d) normalized electroluminescence spectra for CsPbBr3 light emitting devices prepared at different in situ thermally assisted crystallization temperature work on 5 V driving voltage.
Micromachines 14 01104 g007
Table 1. Carrier lifetimes for the CsPbBr3 films at different in situ thermally assisted crystallization temperatures.
Table 1. Carrier lifetimes for the CsPbBr3 films at different in situ thermally assisted crystallization temperatures.
RT40 °C60 °C
A1203.19202.19161.47
τ11.571.591.30
f175.2%63.7%48.8%
A28.0515.2026.28
τ213.0514.038.38
f224.8%36.3%51.2%
Average τ4.415.874.92
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

Zhang, B.; Chen, C.; Chen, X. High-Efficiency CsPbBr3 Light-Emitting Diodes using One-Step Spin-Coating In Situ Dynamic Thermal Crystallization. Micromachines 2023, 14, 1104. https://doi.org/10.3390/mi14061104

AMA Style

Zhang B, Chen C, Chen X. High-Efficiency CsPbBr3 Light-Emitting Diodes using One-Step Spin-Coating In Situ Dynamic Thermal Crystallization. Micromachines. 2023; 14(6):1104. https://doi.org/10.3390/mi14061104

Chicago/Turabian Style

Zhang, Buyue, Chen Chen, and Xinyu Chen. 2023. "High-Efficiency CsPbBr3 Light-Emitting Diodes using One-Step Spin-Coating In Situ Dynamic Thermal Crystallization" Micromachines 14, no. 6: 1104. https://doi.org/10.3390/mi14061104

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