Next Article in Journal
Indocyanine Green-Loaded Quenched Nanoliposomes as Activatable Theranostics for Cancer
Previous Article in Journal
Androgen Receptors in Human Breast Cancer and Female Canine Mammary Tumors
Previous Article in Special Issue
Visible Light-Driven Photocatalysis of Al-Doped SrTiO3: Experimental and DFT Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Composite ZIF-8 with Cs3Bi2I9 to Enhance the Photodegradation Ability on Methylene Blue

1
Key Laboratory of Low-Dimensional Structural Physics and Application, Education Department of Guangxi Zhuang Autonomous Region, College of Physics and Electronic Information Engineering, Guilin University of Technology, Guilin 541004, China
2
School of Electronic Information and Automation, Guilin University of Aerospace Technology, Guilin 541004, China
*
Authors to whom correspondence should be addressed.
Molecules 2025, 30(7), 1413; https://doi.org/10.3390/molecules30071413
Submission received: 17 February 2025 / Revised: 12 March 2025 / Accepted: 19 March 2025 / Published: 22 March 2025

Abstract

:
The development of green, efficient, and reusable photocatalysts is important for pollution degradation. In recent years, Cs3Bi2I9 has been shown to be an effective photocatalyst. However, the rapid recombination of electrons and holes weakens the photocatalytic activity. In this work, the photogenerated electron transfer rate was promoted by ZIF-8 compositing with Cs3Bi2I9, which effectively improved the pollutant degradation. After 50 min of visible light irradiation, Cs3Bi2I9/ZIF-8 removed up to 98.2% of methylene blue (MB), which was 4.15 times higher than that of Cs3Bi2I9 alone. In addition, the Cs3Bi2I9/ZIF-8 composite still exhibited high photocatalytic activity after three cycling experiments. Our research offers a simple and efficient method for enhancing the photocatalytic activity of lead-free halide perovskites.

1. Introduction

While actively promoting industrialization, countries have also brought tremendous challenges to environmental protection and resource conservation [1,2,3]. In recent years, the excessive use of various chemical reagents in the pharmaceutical, agricultural, and manufacturing industries has led to increasingly serious environmental pollution, which is contrary to the original intent of sustainable development [4,5,6]. Among these chemicals, methylene blue (MB) is a typical organic pollutant. Its aqueous solution is alkaline and harmful, causing serious environmental problems due to its non-biodegradation and accumulation in the environment [7]. The common strategies used to treat MB in wastewater are adsorption [8], biodegradation [9], and photocatalytic degradation [10]. Adsorption methods mainly suffer from high costs and secondary pollution issues, while biodegradation methods have long degradation cycles and strict restraint of degradation environments. In contrast, photocatalytic degradation of MB is the most efficient, environmentally friendly, and practical method.
Many materials such as TiO2 [11], ZnO [12], CuO [13], Fe2O3 [14], ZnS [15], and graphene [16] have been widely used for photocatalytic degradation [17,18,19]. However, wide bandgap, low visible light utilization, and fast recombination of photogenerated electron and hole pairs are potential limitations that restrict the photocatalytic activity. Recently, lead-based perovskites have been shown to be an efficient visible-light photocatalyst [20,21,22]. However, the toxicity of lead and instability of these perovskites have hindered its application in photocatalysis [23,24]. Guo et al. used a novel photocatalyst CsPbBr3/UiO-66 to degrade 91.6% of 10 mg/L methyl orange in 90 min under visible light [25]. Gu et al. used nitrogen-doped graphene quantum dots as surfactants to prepare δ-phase CsPbI3 nanocrystals and degraded 96% of rhodamine B in 4 h [26]. Although water stability can be improved by ligand exchange and surface functionalization, it is more important to develop lead-free halide perovskite photocatalysts that are non-toxic and water-stable [27]. In recent years, bismuth-based perovskites [28], which combines low toxicity with excellent water stability, have been discovered in the family of non-lead halide perovskites. In particular, a pioneering study by Sagar M. Jain et al. who chose lead-free bismuth-based perovskite (CH3NH3)3Bi2I9 for the preparation of solar cells, successfully converted solar energy into electric energy, achieving a conversion efficiency of 22.3%, which attracted widespread attention [29]. This breakthrough not only highlights the great potential of bismuth-based perovskites in the field of solar energy conversion but also heralds their important role as efficient photocatalysts.
Among them, inorganic halide perovskite Cs3Bi2I9 has attracted much attention due to its fast carrier transport, high defect tolerance, very low defect density, high water stability, and low toxicity [30,31,32]. Its photoluminescence quantum yield is 2.3% [33]. Effectively separating these electron-hole (e-h) pairs and transferring carriers is a feasible method to increase the number of carriers participating in photocatalysis [34,35,36]. Dou et al. demonstrated excellent stability by using Cs3Bi2I9 composited with Ti3C2 and showed more than 97.3% rhodamine B removal efficiency under visible light [37]. Therefore, it is proved feasible to find a co-catalyst to enhance the rapid separation and transfer of photoexcited e-h pairs in Cs3Bi2I9 to improve the photocatalytic performance [38,39]. In recent years, ZIF-8 has been used in electron transport layers due to its simple synthesis method, good conductivity, good stability, large specific surface area, high carrier mobility, and abundant active center sites [40].
In this study, the photocatalytic degradation performance of MB in aqueous solution was investigated by using Cs3Bi2I9/ZIF-8 composites as photocatalysts. When ZIF-8 is combined with Cs3Bi2I9, ZIF-8 provides abundant charge transfer channels, and the photogenerated electrons in Cs3Bi2I9 can be rapidly transferred to the surface of ZIF-8, which is conducive to the effective transfer and separation of photogenerated carriers [41,42,43], thus greatly improving the photocatalytic efficiency of Cs3Bi2I9. Under visible light, Cs3Bi2I9/ZIF-8 can degrade 98.2% of MB (20 mg/L) within 50 min, much higher than Cs3Bi2I9 (<50%), and there was almost no decrease in the photocatalytic activity after three cycles of the experiment. In addition, the fluorescence performance of Cs3Bi2I9/ZIF-8 photocatalyst in water remains almost unchanged for 30 days, demonstrating its good water stability. This work demonstrates that, by compositing ZIF-8 with Cs3Bi2I9, a highly efficient water-stable perovskite-based photocatalyst is obtained, which opens up prospects for the photocatalytic application of non-lead perovskites.

2. Results

The crystal structures of ZIF-8, Cs3Bi2I9, and Cs3Bi2I9/ZIF-8 were analyzed by XRD. The crystal structure of ZIF-8 characterized by XRD is shown in Figure 1a. The principal diffraction peaks of ZIF-8 are located at 2θ = 7.2°, 10.2°, 12.6°, 14.7°, 16.4°, 17.9°, 19.5°, 22.1°, 24.6°, 26.8°, 29.7°, 30.6°, which are attributed to the (001), (002), (112), (022), (013), (222), and (123) crystal planes, respectively [44]. The XRD patterns of Cs3Bi2I9 and Cs3Bi2I9/ZIF-8 are shown in Figure 1b. The diffraction peaks at 2θ = 21.11°, 25.84°, 27.52°, 29.72°, 32.35°, and 42.97°, respectively, correspond to (110), (202), (203), (204), (205), and (220) crystal planes of cubic-phase Cs3Bi2I9 [45]. This indicates that both the synthesized ZIF-8 and Cs3Bi2I9 exhibit high crystallinity and are consistent with the standard XRD patterns. In the XRD pattern of Cs3Bi2I9/ZIF-8, the characteristic peak at 7.2° attributed to ZIF-8 can be clearly resolved, while other peaks are overlapped with those belonging to Cs3Bi2I9. These results suggest that ZIF-8 was successfully loaded onto Cs3Bi2I9.
The morphologies of the ZIF-8, Cs3Bi2I9 and Cs3Bi2I9/ZIF-8 can be clearly observed through SEM images. As shown in Figure 2a, it can be observed that ZIF-8 is a rhombic dodecahedral crystal with the particle size between 50 and 100 nm [46]. Figure 2b shows that Cs3Bi2I9 exhibits a hexagonal structure with a larger lateral size of 200–400 nm. In the SEM image of Cs3Bi2I9/ZIF-8 [Figure 2c], it is observed that small-size ZIF-8 regular particles were stacked on Cs3Bi2I9 platelets. Figure 2d presents the typical TEM image of Cs3Bi2I9/ZIF-8. The boundary of the blurred Cs3Bi2I9 hexagon and the irregular boundary of ZIF-8 can be discernible. Through high-resolution TEM, two different lattice stripes of Cs3Bi2I9 and ZIF-8 can be further clearly distinguished. The typical lattice spacings are 0.214 and 0.301 nm [Figure 2e,f], corresponding to the (101) crystal face of ZIF-8 and the (204) crystal face of Cs3Bi2I9, respectively [47,48].
Water stability serves as a crucial indicator in assessing the suitability of a catalyst for applications in aqueous environments. As illustrated in Figure 3a, the water contact angles corresponding to Cs3Bi2I9 and Cs3Bi2I9/ZIF-8 are 43° and 88.7°, respectively. The significantly larger water contact angle observed for Cs3Bi2I9/ZIF-8 suggests a reduced interaction surface with water. Thus, to some extent, it can be indirectly determined that Cs3Bi2I9/ZIF-8 has a better water stability. Its optical properties were studied through UV-Vis absorption and PL tests. The band gap of ZIF-8 is 4.9 eV, which is insensitive to the visible-region light [49]. In Figure 3b, the absorption properties of Cs3Bi2I9/ZIF-8 are identical to those of Cs3Bi2I9, with absorption edges up to 600 nm. Based on Tauc plots [50,51], the bandgap (Eg) of Cs3Bi2I9 can be concluded to be 2.37 eV [inset of Figure 3b]. In Figure 3c, it is observed that the PL intensity of Cs3Bi2I9/ZIF-8 is significantly stronger than Cs3Bi2I9 with similar PL centers and shapes, and thus, it is demonstrated that many of the e-h pairs in Cs3Bi2I9 no longer recombine in an emissive manner after ZIF-8 compositing. From the inset of Figure 3c, it can be noticed that the PL shape and intensity of Cs3Bi2I9/ZIF-8 in water remain almost unchanged for 30 days, which indicates its excellent water stability. In addition, the arc radius of Cs3Bi2I9/ZIF-8 is smaller than that of ZIF-8 and pure Cs3Bi2I9 as measured by EIS [Figure 3d], suggesting that it possesses the lowest electron-transfer resistance, which is beneficial for the effective separation and migration of photogenerated carriers. In Figure 3e, the photocurrent response of Cs3Bi2I9 is enhanced after ZIF-8 compositing under the same visible illumination, again indicating an increase in conducting carriers and a decrease in radiative recombination carriers. In a word, such experiments show that after ZIF-8 compositing, more photogenerated carriers in Cs3Bi2I9 no longer undergo the radiative emission and can be effectively transferred to ZIF-8, which benefits their participation in the photocatalytic process.
ZIF-8 is a strong adsorbent which can almost completely adsorb some organic dyes within merely tens of minutes. However, by adsorption alone, MB is very hard to separate from water [52]. The high e-h recombination rate of Cs3Bi2I9 gives rise to a relatively low photodegradation activity. When ZIF-8 is combined with Cs3Bi2I9, the resulting composite exhibits enhanced effectiveness in the decontamination process for removing MB through the adsorption–photodegradation pathway [4]. As shown in Figure 4a, for ZIF-8, Cs3Bi2I9 and Cs3Bi2I9/ZIF-8, the adsorption–desorption equilibrium in the dark can be achieved in 2 h. One can find that the MB adsorption ability of Cs3Bi2I9 is merely ~8% and that of ZIF-8 is ~13%, and however, that of Cs3Bi2I9/ZIF-8 is slightly increased (~15%), which may be attributed to the increase in specific surface area due to the uniform distribution of ZIF-8 on Cs3Bi2I9.
When ZIF-8 is composited with Cs3Bi2I9, it provides more adsorption sites, which are able to capture and immobilize the target pollutants more efficiently, thus improving the adsorption efficiency. The increasing adsorption properties of CZ-1, CZ-2, and CZ-3 were attributed to the increase in ZIF-8 content. As illustrated in Figure 4b, it is evident that under visible light conditions, CZ-1, CZ-2, and CZ-3 exhibit photodegradation efficiencies of 86.3%, 93.4%, and 90.3%, respectively. When the content of ZIF-8 increased from 20 to 30 mg, the photocatalytic efficiency actually decreased, possibly due to too much ZIF-8 blocking the effective light absorption of Cs3Bi2I9. As shown in Figure 4c, Cs3Bi2I9/ZIF-8 shows an obviously higher MB photodegradation efficiency (93.4%) than Cs3Bi2I9 (40%), and along with the adsorption, the total MB removal ratio of Cs3Bi2I9/ZIF-8 (98.2%) is undoubtedly higher than that of sole Cs3Bi2I9 (<50%). Compared with Cs3Bi2I9, the improvement in photocatalytic performance of Cs3Bi2I9/ZIF-8 is attributed to the effective separation and transfer of photo-generated carriers, as discussed in Figure 3. Interestingly, sole ZIF-8 exhibits very weak photodegradation ability on MB, may be due to its abundant active sites [47]. Fitting the corresponding photodegradation kinetics using a first-order reaction equation:
l n C C 0 = k t
where k is the photocatalytic efficiency (in this discussion its unit min−1 is omitted). As shown in Figure 4d, the k value of CZ-2 is 0.05, which is 4.15 times and 8.90 times of Cs3Bi2I9 (k = 0.012) and ZIF-8 (k = 0.005), respectively. In order to investigate recycling property of the photocatalyst Cs3Bi2I9/ZIF-8, three cycles of photocatalytic experiments were conducted. As shown in Figure 4e, in the three cycles, the photodegradation efficiencies were 97%, 89%, and 85%, respectively. There is a slight decrease because the surface of Cs3Bi2I9/ZIF-8 adsorbs MB molecules and it is hard to completely remove them. In order to determine the mechanism of the photocatalytic reaction, free radical trapping experiments were performed. As shown in Figure 4f, benzoquinone (BQ), isopropanol (IPA), silver nitrate (AgNO3), and potassium iodide (KI) were used to capture superoxide radicals (•O2), hydroxyl radicals (•OH), electrons, and holes, respectively. There was no significant decrease in the catalytic activity of Cs3Bi2I9/ZIF-8 in the presence of IPA and KI. However, when in the presence of AgNO3 and BQ, the degradation of MB by Cs3Bi2I9/ZIF-8 was merely 24% and 12%, respectively, indicating that the catalytic activity decreased greatly. This indicates that •O2 radicals and electrons were the main active substances during the photocatalytic degradation process.
In addition, Table 1 compares the relevant photocatalysts used for MB degradation in several studies. It can be seen that many photocatalysts can effectively degrade MB, but the photocatalytic time and degradation efficiency are not satisfactory. Compared with other photocatalyst materials, Cs3Bi2I9/ZIF-8 can photodegrade high-concentration MB with a low dosage, as well as a high degradation efficiency with a short light exposure time. Obviously, Cs3Bi2I9/ZIF-8 is a very promising photocatalyst.
In order to investigate the mechanism of the photocatalysis, the valence band maxima (EVBM) and conduction band minima (ECBM) of ZIF-8 and Cs3Bi2I9 were calculated by UPS measurements [Figure 5a] with EVBM = −[21.22 − (Ecutoff − Eonset)]. For ZIF-8, the Ecutoff was 15.48 eV and the Eonset was 3.31 eV. The EVBM of ZIF-8 can be calculated as −9.05 eV according to the formula. Since the bandgap (Eg) of ZIF-8 is 4.9 eV, the ECBM of ZIF-8 can be calculated as −4.55 eV according to the formula Eg = ECBM − EVBM [59]. Typically, we use a standard hydrogen electrode (NHE) as a reference electrode to represent the energy levels, i.e., E(NHE) = −4.5 − E(vacuum), to express the relationship between the NHE and the vacuum energy levels, which results in the VBM and CBM values of 4.55 eV and −0.35 eV, respectively. Similarly, the VBM and CBM values of Cs3Bi2I9 were 2.01 eV and −0.39 eV, respectively, according to Figure 5a. Based on the above analysis, a possible mechanism for the photocatalytic removal of MB by Cs3Bi2I9/ZIF-8 can be proposed [Figure 5b]:
hν (on Cs3Bi2I9) → h+ + e
When Cs3Bi2I9/ZIF-8 is exposed to visible light, a large number of electrons and holes are generated in Cs3Bi2I9, and the electrons in the valence band (VB) are excited to the conduction band (CB) while the holes are left in the VB. As shown in Equation (2), ZIF-8 cannot be photoexcited here due to its large band gap. Due to the close contact at the heterojunction interface between ZIF-8 and Cs3Bi2I9, electrons are spontaneously transferred from Cs3Bi2I9 (ECBM = −0.39 eV) to ZIF-8 (ECBM = −0.35 eV). The value of −0.35 eV is lower than the potential of O2/•O2 (−0.33 eV); therefore, the light-generated electrons can react with dissolved O2 to form strongly oxidizing •O2 at the interface. The superoxide radical (•O2) preferentially oxidizes the thioether group (-S-) and amino group [-N(CH3)2] of MB, converting them into sulfoxide (-SO-) and sulfone (-SO2) groups, thereby destroying the phenothiazine ring conjugation structure and generating open-ring intermediates. These open-ring products undergo hydroxylation and decarboxylation reactions to gradually release CO2, forming small-molecule acids such as formic acid and oxalic acid, ultimately mineralizing into CO2 and H2O. Since the results show that in our experiment only electrons and •O2 radicals play an important role [Figure 4f], the photodegradation process can be represented in Equations (3) and (4).
O2 + e → •O2
*MB + •O2 → mineralization products
In addition to enhancing the electron transfer in Cs3Bi2I9/ZIF-8, ZIF-8 can also provide abundant active sites and increase the contact surface area between the active sites and MB, thereby facilitating the photocatalytic reaction. The EVBM of Cs3Bi2I9 is 2.01 eV, which is higher than the potential of H2O/h+ (1.99 eV), and however, the results in Figure 4f show that holes and •OH radicals barely take part in the photocatalytic process. Maybe it is because MB photodegradation is more likely to occur on the ZIF-8 surface of Cs3Bi2I9/ZIF-8.

3. Experimental Sections

3.1. Chemicals

Cesium acetate (CH3COOCs, 99.9%), 1-octadecene (ODE, 90%), iodotrimethylsilane (TMS-I, 97%), silver nitrate (AgNO3, 99.99%) and potassium iodide (KI, 98%) were purchased from Roan (Dundee, UK). Bismuth acetate [Bi(CH₃COO)₃, 99.99%] was purchased from Bidepharm (Shanghai, China). Oleic acid (OA, AR) was purchased from Aladdin (Tokyo, Japan). Oleylamine (OLA, 70%) and benzoquinone (BQ, 97%) were purchased from Macklin (Shanghai, China). Hexane (C6H14, 97%, AR), isopropyl alcohol (IPA, 99.7%, AR), sodium sulfate anhydrous (Na2SO4, 99%, AR), ethanol anhydrous (C2H5OH, 99%, GR), acetone (CH3COCH3, 99.5%, AR), and MB trihydrate were purchased from Xilong Scientific (Shantou, China). Zinc nitrate hexahydrate [Zn(NO3)2·6H2O] was purchased from Xilong Scientific. 2-methylimidazole (2-MIM, C4H6N2) was purchased from Macklin. Nafion solution (5%) was purchased from Chengxin Science and Technology (Beijing, China). The purified water was sourced from Wahaha (Hangzhou, China). All reagents and solvents were used as received without any further purification.

3.2. Synthesis of ZIF-8, Cs3Bi2I9 and Cs3Bi2I9/ZIF-8

Synthesis of ZIF-8: First, 0.66 g (8.04 mmol) 2-MIM was added to 14.3 mL (0.353 mol) methanol. Then, 0.3 g (1.008 mmol) Zn(NO3)2·6H2O was added to 14.3 mL (0.353 mol) methanol, and the solution of 2-MIM in methanol was poured into the methanol solution of Zn(NO3)2·6H2O. The final mixture was stirred at 1000 rpm for 1 h and then allowed to stand for 24 h. After standing, the solution was centrifuged at 8000 rpm for 3 min. Subsequently, the precipitate was washed twice with methanol, centrifuged, and dried to obtain ZIF-8.
Synthesis of Cs3Bi2I9 and Cs3Bi2I9/ZIF-8: A total of 0.14 mmol cesium acetate, 0.2 mmol bismuth acetate, 1.5 mL (4.75 mmol) OA, and 0.37 mL (1.125 mmol) OLA were dissolved in 6 mL (0.01875 mol) ODE. The mixture was heated to 120 °C under N2 atmosphere. After maintaining the temperature for 30 min, the solution temperature was raised to 150 °C and 0.2 mL (1.7 mmol) TMS-I was rapidly injected to react for 1 min. After that, it was cooled in an ice water bath for 30 s. It was centrifuged at 8000 rpm for 15 min to obtain the precipitate. The precipitate was dispersed in hexane and centrifuged at 5000 rpm for another 15 min, and finally the precipitate was collected and dried in an oven at 60 °C for 24 h to obtain Cs3Bi2I9 powder. The synthesis process of Cs3Bi2I9/ZIF-8 is similar and shown in Figure 6, only differing in the fact that ZIF-8 was added into the cesium acetate/bismuth acetate solution when the temperature was heated to 120 °C. The obtained Cs3Bi2I9/ZIF-8 samples by adding 10, 20, and 30 mg ZIF-8 are, respectively, named as CZ-1, CZ-2 and CZ-3. If not specially noticed, Cs3Bi2I9/ZIF-8 below stands for CZ-2.

3.3. Adsorption and Photocatalytic Test

The adsorption capacity of the sample was evaluated in the dark. First, add 20 mg sample (ZIF-8, Cs3Bi2I9 or Cs3Bi2I9/ZIF-8) to 50 mL MB solution (20 mg/L) and stir at a certain speed. At intervals of 1 h, 3 mL of the solution is withdrawn, and the concentration of MB was quantified using a UV-Vis spectrophotometer (Shimadzu UV-2700, Kyoto, Japan) by measuring its absorbance at 664 nm. The adsorption–desorption equilibrium of all the samples could be reached in 2 h. Visible-light photodegradation of MB (50 mL, 20 mg/L) was performed under xenon lamp (PLS-SXE 300, PerfectLight, Beijing, China, 300 W, λ > 420 nm). After the adsorption–desorption equilibrium was achieved, 3 mL of solution was withdrawn each time to measure the current concentration of MB at intervals of 10 min. The MB removal (including adsorption and photodegradation) efficiency was calculated with the formula [26]
Removal   efficiency   ( % ) = C 0 C C 0 × 100 % = A 0 A A 0 × 100 %
where C0 and C are the initial and real-time concentration of MB, respectively. A0 and A represent the initial and real-time absorbance of MB at the wavelength of 664 nm, respectively. The recycling photodegradation tests were conducted by repeatedly collecting the photocatalysts through centrifugation at 8500 rpm and drying at 60 °C. Free radical capture experiments were conducted using IPA, AgNO3, KI, and BQ as scavengers to determine the photocatalytic active species.

3.4. Electrochemical Tests

Transient photocurrents and electrochemical impedances were measured using a standard three-electrode electrochemical workstation (CHI 660E, Shanghai CH Instrument Company, Shanghai, China). The electrolyte was Na2SO4 solution (0.2 M), and the light source was a xenon lamp (PLS-SXE300, 300 W). Graphite electrode was used as the counter electrode, AgCl electrode as the reference electrode, and the sample as the working electrode. The photocatalyst was deposited onto the fluorine-doped tin oxide (FTO) electrode. A pristine FTO electrode was acquired through sequential soak and sonication in acetone, followed by deionized water washing and sonication in anhydrous ethanol. The mixture of 10 mg sample, 100 μL deionized water, 100 μL anhydrous ethanol and 30 μL nafion was sonicated for 20 min. Subsequently, the mixture was dropwise added using a pipette gun, consequently leading to the formation of an orange-red coating on the electrically charged side of the FTO electrode. After the resultant FTO electrode was dried in an oven, the sample-coating FTO electrode was then used for tests.

3.5. Characterization

The X-ray diffractometer (XRD, Miniflex 600, Rigaku, Tokyo, Japan) was used to study the crystal structure of Cs3Bi2I9, ZIF-8 and the composite material Cs3Bi2I9/ZIF-8. Its scanning angle ranged from 5 to 60°, with a scanning speed of 2°/min. The elemental compositions of ZIF-8, Cs3Bi2I9 and Cs3Bi2I9/ZIF-8 were characterized using an X-ray photoelectron spectrometer (XPS, Thermo Scientific K-Alpha, Waltham, MA, USA). The morphology of ZIF-8 was analyzed using scanning electron microscopy (SEM, TESCAN MIRA LMS, Brno, Czech Republic). The microstructures of Cs3Bi2I9 and Cs3Bi2I9/ZIF-8 were characterized and the lattice spacing was determined using high resolution transmission electron microscope (TEM, JEM-2100F, Hitachi, Tokyo, Japan). The valence electron structure of Cs3Bi2I9 was measured using UV photoelectron spectroscopy (UPS, ESCALAB 250Xi, Thermo Scientific, Waltham, MA, USA). The UV-Vis absorption spectra of photocatalysts and organic pollutants were measured using a UV spectrophotometer (UV2700, Shimadzu, Kyoto, Japan), with the measurement wavelength range being 200~800 nm. The photoluminescence (PL) spectrum of the sample was measured using a fluorescence spectrometer (Edinburgh FL/FS900 Carry Eclipse, USA) from Agilent Technologies. An excitation wavelength of 380 nm is used, and the collected emission rage is 400–800 nm. The contact angle between Cs3Bi2I9, Cs3Bi2I9/ZIF-8 and water was measured by using SDC-200 angle measuring instrument of Dongwan Shengding Precision Instrument, Dongguan, China.

4. Conclusions

In summary, Cs3Bi2I9/ZIF-8 photocatalysts with high catalytic activity were successfully prepared. The experimental results showed that the optimized Cs3Bi2I9/ZIF-8 had a synergistic adsorption and photodegradation removal rate of MB up to 98.2%, which was better than most materials reported previously. The excellent photocatalytic performance is attributed to the efficient separation of photogenerated e-h pairs and electron transfer in Cs3Bi2I9 by ZIF-8. In addition, it also advances in good water stability and reusability. The results also show that •O2 radicals are the main substances to oxidize MB in the photocatalytic degradation process. This work provides a promising strategy in the development of stable and cost-effective photocatalysts based on lead-free halide perovskites.

Author Contributions

T.T.; writing—review and editing, H.Z.; writing—original draft preparation, X.D. and H.Z.; data curation, H.W.; methodology and formal analysis, J.W.; investigation, L.J.; project administration, L.J.; funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the Natural Science Foundation of Guangxi Zhuang Autonomous Region (2025GXNSFAA069986), the Science and Technology Base and Talent Special Project of Guangxi Province (AD21220029), and the Research Foundation of Guilin University of Technology (GUTQDJJ2019011).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Saravanan, A.; Kumar, P.S.; Vo, D.-V.N.; Yaashikaa, P.R.; Karishma, S.; Jeevanantham, S.; Gayathri, B.; Bharathi, V.D. Photocatalysis for removal of environmental pollutants and fuel production: A review. Environ. Chem. Lett. 2020, 19, 441–463. [Google Scholar] [CrossRef]
  2. Deng, F.; Shi, H.; Guo, Y.; Luo, X.; Zhou, J. Engineering paths of sustainable and green photocatalytic degradation technology for pharmaceuticals and organic contaminants of emerging concern. Curr. Opin. Green Sustain. Chem. 2021, 29, 100465. [Google Scholar]
  3. Kumar, A.; Raizada, P.; Singh, P.; Saini, R.V.; Saini, A.K.; Hosseini-Bandegharaei, A. Perspective and status of polymeric graphitic carbon nitride based Z-scheme photocatalytic systems for sustainable photocatalytic water purification. Chem. Eng. J. 2020, 391, 123496. [Google Scholar]
  4. Kumar, M.; Barbhai, M.D.; Puranik, S.; Radha; Natta, S.; Senapathy, M.; Dhumal, S.; Singh, S.; Kumar, S.; Deshmukh, V.P.; et al. Combination of green extraction techniques and smart solvents for bioactives recovery. TrAC Trends Anal. Chem. 2023, 169, 117286. [Google Scholar]
  5. Yang, K.; Hao, L.; Hou, Y.; Zhang, J.; Yang, J.-H. Summary and application of Ni-based catalysts for electrocatalytic urea oxidation. Int. J. Hydrogen Energy 2024, 51, 966–981. [Google Scholar]
  6. Yu, Y.; Lin, Z.; Liu, D.; Hou, Y. Exploring the spatially heterogeneous impacts of industrial agglomeration on regional sustainable development capability: Evidence from new energy industries. Environ. Dev. Sustain. 2023, 26, 16657–16682. [Google Scholar]
  7. Rafiq, A.; Ikram, M.; Ali, S.; Niaz, F.; Khan, M.; Khan, Q.; Maqbool, M. Photocatalytic degradation of dyes using semiconductor photocatalysts to clean industrial water pollution. J. Ind. Eng. Chem. 2021, 97, 111–128. [Google Scholar]
  8. Jia, P.; Tan, H.; Liu, K.; Gao, W. Adsorption behavior of methylene blue by bone char. Int. J. Mod. Phys. B 2017, 31, 1744099. [Google Scholar]
  9. Theerthagiri, J.; Lee, S.J.; Karuppasamy, K.; Arulmani, S.; Veeralakshmi, S.; Ashokkumar, M.; Choi, M.Y. Application of advanced materials in sonophotocatalytic processes for the remediation of environmental pollutants. J. Hazard. Mater. 2021, 412, 125245. [Google Scholar]
  10. Dai, H.; Yuan, X.; Jiang, L.; Wang, H.; Zhang, J.; Zhang, J.; Xiong, T. Recent advances on ZIF-8 composites for adsorption and photocatalytic wastewater pollutant removal: Fabrication, applications and perspective. Coord. Chem. Rev. 2021, 441, 213985. [Google Scholar]
  11. Rojviroon, T.; Rojviroon, O.; Sirivithayapakorn, S. Photocatalytic decolourisation of dyes using TiO2 thin film photocatalysts. Surf. Eng. 2016, 32, 562–569. [Google Scholar] [CrossRef]
  12. Yildiz, S.; Canbaz, G.T.; Mihçiokur, H. Photocatalytic degradation of oxytetracycline using ZnO catalyst. Environ. Prog. Sustain. Energy 2024, 43, 14384. [Google Scholar] [CrossRef]
  13. Sahu, K.; Satpati, B.; Mohapatra, S. Facile fabrication of CuO nanosheets for photocatalytic applications. Appl. Phys. A 2021, 127, 361. [Google Scholar] [CrossRef]
  14. Uttam Pandit, V.R.; Parshuram Jadhav, G.K.; Sakharam Jawale, V.M.; Dubepatil, R.; Gurao, R.; Late, D.J. Synthesis and characterization of micro-/nano-α-Fe2O3 for photocatalytic dye degradation. RSC Adv. 2024, 14, 29099–29105. [Google Scholar] [CrossRef] [PubMed]
  15. Maaß, M.C.; Tasch, A.; Jooss, C.; Waitz, T. Photocatalytic hydrogen evolution using ZnS particles and LEDs. J. Chem. Educ. 2022, 99, 2086–2092. [Google Scholar] [CrossRef]
  16. Li, X.; Yu, J.; Wageh, S.; Al-Ghamdi, A.A.; Xie, J. Graphene in photocatalysis: A review. Small 2016, 12, 6640–6696. [Google Scholar] [CrossRef]
  17. Donga, C.; Mishra, S.B.; Abd-El-Aziz, A.S.; Mishra, A.K. Advances in graphene-based magnetic and graphene-based/TiO2 nanoparticles in the removal of heavy metals and organic pollutants from industrial wastewater. J. Inorg. Organomet. Polym. Mater. 2020, 31, 463–480. [Google Scholar] [CrossRef]
  18. Fatima, R.; Warsi, M.F.; Zulfiqar, S.; Ragab, S.A.; Shakir, I.; Sarwar, M.I. Nanocrystalline transition metal oxides and their composites with reduced graphene oxide and carbon nanotubes for photocatalytic applications. Ceram. Int. 2020, 46, 16480–16492. [Google Scholar] [CrossRef]
  19. Hong, X.; Wang, X.; Li, Y.; Fu, J.; Liang, B. Progress in graphene/metal oxide composite photocatalysts for degradation of organic pollutants. Catalysts 2020, 10, 921. [Google Scholar] [CrossRef]
  20. Mehra, S.; Mamta; Tawale, J.; Gupta, G.; Singh, V.N.; Srivastava, A.K.; Sharma, S.N. Evaluating Pb-based and Pb-free halide perovskites for solar-cell applications: A simulation study. Heliyon 2024, 10, 33243. [Google Scholar] [CrossRef]
  21. Thien, G.S.H.; Ab Rahman, M.; Yap, B.K.; Tan, N.M.L.; He, Z.; Low, P.L.; Devaraj, N.K.; Ahmad Osman, A.F.; Sin, Y.K.; Chan, K.Y. Recent advances in halide perovskite resistive switching memory devices: A transformation from lead-based to lead-free perovskites. ACS Omega 2022, 7, 39472–39481. [Google Scholar] [CrossRef]
  22. Walker, B.; Kim, G.H.; Kim, J.Y. Pseudohalides in lead-based perovskite semiconductors. Adv. Mater. 2019, 31, 1807029. [Google Scholar]
  23. Kwak, J.I.; Lee, T.-Y.; An, Y.-J. Assessing the potential toxicity of hazardous material released from Pb-based perovskite solar cells to crop plants. J. Clean. Prod. 2023, 423, 138856. [Google Scholar]
  24. Zhu, Y.; Kang, Y.; Huang, H.; Zhuang, D.; Li, M.; Ling, Z.; Peng, K.; Zhai, L.; Zou, C. Systematic evaluation of the biotoxicity of Pb-based perovskite materials and perovskite solar cells. J. Mater. Chem. A 2024, 12, 2916–2923. [Google Scholar]
  25. Guo, L.; Zhang, C.; Tan, T.; Li, L.; Qian, Q.; Mo, Y.; Lei, X.; Cai, W.; Wang, Z. Effect of band structure on the photocatalytic performance of CsPbBr3/UiO-66 composites. J. Solid State Chem. 2024, 339, 124962. [Google Scholar]
  26. Gu, Y.; Du, X.; Hua, F.; Wen, J.; Li, M.; Tang, T. Nitrogen-doped graphene quantum dot-passivated δ-phase CsPbI3: A water-stable photocatalytic adjuvant to degrade rhodamine B. Molecules 2023, 28, 7310. [Google Scholar] [CrossRef]
  27. Saparov, B.; Hong, F.; Sun, J.-P.; Duan, H.-S.; Meng, W.; Cameron, S.; Hill, I.G.; Yan, Y.; Mitzi, D.B. Thin-film preparation and characterization of Cs3Sb2I9: A lead-free layered perovskite semiconductor. Chem. Mater. 2015, 27, 5622–5632. [Google Scholar]
  28. Lee, L.C.; Huq, T.N.; MacManus-Driscoll, J.L.; Hoye, R.L.Z. Research update: Bismuth-based perovskite-inspired photovoltaic materials. APL Mater. 2018, 6, 084502. [Google Scholar]
  29. Jain, S.M.; Edvinsson, T.; Durrant, J.R. Green fabrication of stable lead-free bismuth based perovskite solar cells using a non-toxic solvent. Commun. Chem. 2019, 2, 91. [Google Scholar]
  30. Ghosh, B.; Wu, B.; Mulmudi, H.K.; Guet, C.; Weber, K.; Sum, T.C.; Mhaisalkar, S.; Mathews, N. Limitations of Cs3Bi2I9 as lead-free photovoltaic absorber materials. ACS Appl. Mater. Interfaces 2018, 10, 35000–35007. [Google Scholar]
  31. Liu, J.; Nie, W.; Yan, L.; Hu, H.; Zhang, G.; Lin, P.; Hu, H.; Xu, L.; Wang, P.; Cui, C. Anisotropic optoelectronic properties of lead-free Cs3Bi2I9 single-crystal photodetectors. J. Phys. D Appl. Phys. 2024, 57, 335101. [Google Scholar] [CrossRef]
  32. Zhang, R.; Chen, Z.; Ma, J.; Zhang, P.; Liu, M.; Li, X.; Zhao, R.; Tang, J.; Ren, Z.; Li, S. An all-solid-state photo-rechargeable battery based on Cs3Bi2I9. Chem. Commun. 2023, 59, 2911–2914. [Google Scholar]
  33. Lou, Y.; Fang, M.; Chen, J.; Zhao, Y. Formation of highly luminescent cesium bismuth halide perovskite quantum dots tuned by anion exchange. Chem. Commun. 2018, 54, 3779–3782. [Google Scholar]
  34. Ji, J.W.; Zhang, L.J.; Yi, W.J.; Wang, Y.; Chen, J.H.; Yue, X.Z.; Yi, S.S. Re3P4@C/TiO2 ohmic junction boosts charge carrier separation for photocatalytic hydrogen evolution. Chem. Eng. J. 2024, 500, 157277. [Google Scholar] [CrossRef]
  35. Yali, Z.; Rong, W.; Aolin, L.; Jialei, H.; Zhilong, Z.; Shunhang, W. Constructing surface oxygen vacancy in the [Bi2O2]2+ layer defects mediated Bi2MoO6 enhanced visible light responsive photocatalytic activity. J. Chem. Phys. 2024, 161, 184707. [Google Scholar]
  36. Qiao, Y.N.; Zhang, Y.; Yuan, J.; Xue, H.M.; Jia, B. Construction of NiCo2S4/ZnCdS Schottky junction with unidirectional interfacial electron transfer for boosting photocatalytic hydrogen production. ACS Appl. Energy Mater. 2024, 7, 9921–9929. [Google Scholar] [CrossRef]
  37. Tang, T.; Dou, X.; Zhang, H.; Wang, H.; Li, M.; Hu, G.; Wen, J.; Jiang, L. Enhancing the photocatalytic activity of lead-free halide perovskite Cs3Bi2I9 by compositing with Ti3C2 MXene. Molecules 2024, 29, 5096. [Google Scholar] [CrossRef]
  38. Chen, S.; Yin, H.; Liu, P.; Wang, Y.; Zhao, H. Stabilization and performance enhancement strategies for halide perovskite photocatalysts. Adv. Mater. 2022, 35, 2203836. [Google Scholar]
  39. Fang, Z.; Yue, X.; Xiang, Q. Atomically contacted Cs3Bi2Br9 QDs@UiO-66 composite for photocatalytic CO2 reduction. Small 2024, 20, 2401914. [Google Scholar]
  40. Zhang, Z.; Luo, X.; Wang, B.; Zhang, J. Electron transport improvement of perovskite solar cells via a ZIF-8-derived porous carbon skeleton. ACS Appl. Energy Mater. 2019, 2, 2760–2768. [Google Scholar]
  41. Karim, M.R.; Choi, C.H.; Mohammad, A.; Yoon, T. Fabrication of high-performance supercapacitor of surface-engineered ZIF-8 for energy storage applications. J. Energy Storage 2024, 93, 112199. [Google Scholar]
  42. Shi, Y.; Zhou, H.; Dai, L.; Liu, D.; Du, W. Preparation of ordered macroporous ZIF-8-derived magnetic carbon materials and its application for lipase immobilization. Catalysts 2024, 14, 55. [Google Scholar] [CrossRef]
  43. Wang, Z.; Han, R.; Li, L.; Sun, J.; Yang, J.; Pan, M.; Wang, S. Electrochemical strategy based on the synergistic effect of ZIF-8 and MWCNTs for quantitation of tert-butylhydroquinone in oils and fried chips. Microchem. J. 2023, 185, 108268. [Google Scholar]
  44. Ayyob, M.; Sun, Z.; Liu, Y.Y.; Wang, Y.; Wang, A. Performance of amine-modified CdS-D@ZIF-8 nanocomposites for enhanced photocatalytic H2 evolution. Int. J. Hydrogen Energy 2023, 48, 25645–25659. [Google Scholar]
  45. Karthieka, R.R.; Prakash, T. Dose-dependent X-ray sensing behaviour of Cs3Bi2I9: PVDF-HFP nanocomposites. Phys. E Low-Dimens. Syst. Nanostructures 2021, 133, 114823. [Google Scholar]
  46. Ta, D.N.; Nguyen, H.K.D.; Trinh, B.X.; Le, Q.T.N.; Ta, H.N.; Nguyen, H.T. Preparation of nano-ZIF-8 in methanol with high yield. Can. J. Chem. Eng. 2018, 96, 1518–1531. [Google Scholar]
  47. Zheng, G.; Chen, Z.; Sentosun, K.; Pérez-Juste, I.; Bals, S.; Liz-Marzán, L.M.; Pastoriza-Santos, I.; Pérez-Juste, J.; Hong, M. Shape control in ZIF-8 nanocrystals and metal nanoparticles@ZIF-8 heterostructures. Nanoscale 2017, 9, 16645–16651. [Google Scholar]
  48. Pramod, A.K.; Kushvaha, S.S.; Batabyal, S.K. Lead-free Cs3Bi2I9 perovskite hexagonal microplates: A promising material solution-processed for ultraviolet self-powered photodetectors. J. Alloys Compd. 2024, 1006, 176320. [Google Scholar]
  49. Pambudi, F.I. Electronic properties of heterometallic zeolitic imidazolate framework and its encapsulation with Ni, Pd and Pt. Inorg. Chem. Commun. 2022, 143, 109798. [Google Scholar]
  50. Chen, X.; Chen, C.; Zang, J. Bi2MoO6 nanoflower-like microsphere photocatalyst modified by boron doped carbon quantum dots: Improving the photocatalytic degradation performance of BPA in all directions. J. Alloys Compd. 2023, 962, 171167. [Google Scholar]
  51. Chen, X.; Chen, C.; Zang, J. Hybrid of carbon quantum dots modified g-C3N4 nanosheets and MoS2 nanospheres: Indirectly promote hydroxyl group production for efficient degradation of methyl orange. Diam. Relat. Mater. 2023, 139, 110385. [Google Scholar]
  52. Hong, Y.; Wang, B.; Hu, S.; Lu, S.; Wu, Q.; Fu, M.; Gu, C.; Wang, Y. Preparation and photocatalytic performance of Zn2SnO4/ZIF-8 nanocomposite. Ceram. Int. 2023, 49, 11027–11037. [Google Scholar]
  53. Jing, H.P.; Wang, C.C.; Zhang, Y.W.; Wang, P.; Li, R. Photocatalytic degradation of methylene blue in ZIF-8. RSC Adv. 2014, 4, 54454–54462. [Google Scholar] [CrossRef]
  54. Kitchamsetti, N.; Didwal, P.N.; Mulani, S.R.; Patil, M.S.; Devan, R.S. Photocatalytic activity of MnTiO3 perovskite nanodiscs for the removal of organic pollutants. Heliyon 2021, 7, 07297. [Google Scholar]
  55. Wang, Y.; Wang, L.; Liu, R.; Li, X. Casein templated synthesis of porous perovskite and its application in visible-light photocatalytic degradation of methylene blue. Mater. Sci. Semicond. Process. 2019, 103, 104597. [Google Scholar]
  56. Bresolin, B.M.; Ben Hammouda, S.; Sillanpää, M. An emerging visible-light organic–inorganic hybrid perovskite for photocatalytic applications. Nanomaterials 2020, 10, 115. [Google Scholar] [CrossRef] [PubMed]
  57. Kumar, M.; Vaish, R.; Elqahtani, Z.M.; Kebaili, I.; Al-Buriahi, M.S.; Sung, T.H.; Hwang, W.; Kumar, A. Piezo-photocatalytic activity of Bi2VO5.5 for methylene blue dye degradation. J. Mater. Res. Technol. 2022, 21, 1998–2012. [Google Scholar]
  58. Tichapondwa, S.M.; Newman, J.P.; Kubheka, O. Effect of TiO2 phase on the photocatalytic degradation of methylene blue dye. Phys. Chem. Earth Parts A/B/C 2020, 118–119, 102900. [Google Scholar]
  59. Yao, M.; Yu, H.; Ding, K.; Du, Z.; Zhang, W.; Xia, Y.; Pan, X.; Wang, W.; Liu, X.; Li, S.; et al. Quantum dot/ZIF-67/ZIF-8 heterostructure for efficient photocatalytic hydrogen production. ACS Appl. Nano Mater. 2024, 7, 11445–11454. [Google Scholar]
Figure 1. XRD patterns of (a) ZIF-8 and (b) Cs3Bi2I9 and Cs3Bi2I9/ZIF-8. The simulated and standard XRD patterns of ZIF-8 and Cs3Bi2I9 are used for comparison.
Figure 1. XRD patterns of (a) ZIF-8 and (b) Cs3Bi2I9 and Cs3Bi2I9/ZIF-8. The simulated and standard XRD patterns of ZIF-8 and Cs3Bi2I9 are used for comparison.
Molecules 30 01413 g001
Figure 2. SEM images of (a) ZIF-8, (b) Cs3Bi2I9, (c) Cs3Bi2I9/ZIF-8. (d) High-resolution TEM image of Cs3Bi2I9/ZIF-8. High-resolution TEM images of (e) ZIF-8 and (f) Cs3Bi2I9 in Cs3Bi2I9/ZIF-8 composite.
Figure 2. SEM images of (a) ZIF-8, (b) Cs3Bi2I9, (c) Cs3Bi2I9/ZIF-8. (d) High-resolution TEM image of Cs3Bi2I9/ZIF-8. High-resolution TEM images of (e) ZIF-8 and (f) Cs3Bi2I9 in Cs3Bi2I9/ZIF-8 composite.
Molecules 30 01413 g002
Figure 3. (a) Water contact angles of Cs3Bi2I9 and Cs3Bi2I9/ZIF-8. (b) UV–Vis absorption spectra of Cs3Bi2I9 and Cs3Bi2I9/ZIF-8. The inset is the corresponding Tauc plot of Cs3Bi2I9. (c) PL spectra of Cs3Bi2I9 and Cs3Bi2I9/ZIF-8. The inset is the PL spectra of Cs3Bi2I9/ZIF-8 on the 1st, 15th, and 30th day. (d) Electrochemical impedance measurements of ZIF-8, Cs3Bi2I9 and Cs3Bi2I9/ZIF-8. The inset is the equivalent circuit diagram, where Rs, Rct and CPE denote the total resistance of the external circuit, the charge transfer resistance and constant phase element. (e) The visible-light photocurrent responses of Cs3Bi2I9 and Cs3Bi2I9/ZIF-8.
Figure 3. (a) Water contact angles of Cs3Bi2I9 and Cs3Bi2I9/ZIF-8. (b) UV–Vis absorption spectra of Cs3Bi2I9 and Cs3Bi2I9/ZIF-8. The inset is the corresponding Tauc plot of Cs3Bi2I9. (c) PL spectra of Cs3Bi2I9 and Cs3Bi2I9/ZIF-8. The inset is the PL spectra of Cs3Bi2I9/ZIF-8 on the 1st, 15th, and 30th day. (d) Electrochemical impedance measurements of ZIF-8, Cs3Bi2I9 and Cs3Bi2I9/ZIF-8. The inset is the equivalent circuit diagram, where Rs, Rct and CPE denote the total resistance of the external circuit, the charge transfer resistance and constant phase element. (e) The visible-light photocurrent responses of Cs3Bi2I9 and Cs3Bi2I9/ZIF-8.
Molecules 30 01413 g003
Figure 4. (a) The MB adsorption performances in the dark. (b) Photocatalytic degradation of MB of Cs3Bi2I9/ZIF-8 with different ZIF-8 mass. (c) The MB photodegradation of different samples and (d) the corresponding photodegradation kinetics behavior. (e) Photocatalytic cycle tests of Cs3Bi2I9/ZIF-8. (f) The impact of photocatalytic effect of Cs3Bi2I9/ZIF-8 after scavengers addition.
Figure 4. (a) The MB adsorption performances in the dark. (b) Photocatalytic degradation of MB of Cs3Bi2I9/ZIF-8 with different ZIF-8 mass. (c) The MB photodegradation of different samples and (d) the corresponding photodegradation kinetics behavior. (e) Photocatalytic cycle tests of Cs3Bi2I9/ZIF-8. (f) The impact of photocatalytic effect of Cs3Bi2I9/ZIF-8 after scavengers addition.
Molecules 30 01413 g004
Figure 5. (a) UPS spectra of ZIF-8 and Cs3Bi2I9. (b) The mechanism of photocatalytic process of Cs3Bi2I9/ZIF-8.
Figure 5. (a) UPS spectra of ZIF-8 and Cs3Bi2I9. (b) The mechanism of photocatalytic process of Cs3Bi2I9/ZIF-8.
Molecules 30 01413 g005
Figure 6. The synthesis process of Cs3Bi2I9/ZIF-8.
Figure 6. The synthesis process of Cs3Bi2I9/ZIF-8.
Molecules 30 01413 g006
Table 1. Comparison of MB degradation effects of photocatalysts under visible or UV Light.
Table 1. Comparison of MB degradation effects of photocatalysts under visible or UV Light.
CatalystIlluminationQuantityC0EfficiencyCatalytic TimeRef.
ZIF-8Ultraviolet light25 mg10 mg/L82.3%120 min[53]
MnTiO3Visible light25 mg10 mg/L89.7%180 min[54]
LaMnO3Visible light5 mg5 mg/L95%120 min[55]
MAIPbVisible light25 mg10 mg/L90%60 min[56]
BiVO4Visible light0.25 g5 mg/L81%240 min[57]
TiO2Ultraviolet light0.15 g10 mg/L81.4%100 min[58]
Cs3Bi2I9/ZIF-8Visible light20 mg20 mg/L93.4%50 minThis work
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

Tang, T.; Zhang, H.; Wang, H.; Dou, X.; Wen, J.; Jiang, L. Composite ZIF-8 with Cs3Bi2I9 to Enhance the Photodegradation Ability on Methylene Blue. Molecules 2025, 30, 1413. https://doi.org/10.3390/molecules30071413

AMA Style

Tang T, Zhang H, Wang H, Dou X, Wen J, Jiang L. Composite ZIF-8 with Cs3Bi2I9 to Enhance the Photodegradation Ability on Methylene Blue. Molecules. 2025; 30(7):1413. https://doi.org/10.3390/molecules30071413

Chicago/Turabian Style

Tang, Tao, Haoran Zhang, Hexu Wang, Xiaoyu Dou, Jianfeng Wen, and Li Jiang. 2025. "Composite ZIF-8 with Cs3Bi2I9 to Enhance the Photodegradation Ability on Methylene Blue" Molecules 30, no. 7: 1413. https://doi.org/10.3390/molecules30071413

APA Style

Tang, T., Zhang, H., Wang, H., Dou, X., Wen, J., & Jiang, L. (2025). Composite ZIF-8 with Cs3Bi2I9 to Enhance the Photodegradation Ability on Methylene Blue. Molecules, 30(7), 1413. https://doi.org/10.3390/molecules30071413

Article Metrics

Back to TopTop