Next Article in Journal
Catalytical Performance of Heteroatom Doped and Undoped Carbon-Based Materials
Next Article in Special Issue
One-Pot Fabrication of 2D/2D CdIn2S4/In2S3 Heterojunction for Boosting Photocatalytic Cr(VI) Reduction
Previous Article in Journal
Gold(I)-Catalyzed Direct Alkyne Hydroarylation in Ionic Liquids: Mechanistic Insights
Previous Article in Special Issue
Band Gap Engineering in Quadruple-Layered Sillén–Aurivillius Perovskite Oxychlorides Bi7Fe2Ti2O17X (X = Cl, Br, I) for Enhanced Photocatalytic Performance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Preparation of Two-Dimensional Layered CeO2/Bi2O3 Composites for Efficient Photocatalytic Desulfurization

1
School of Material Science and Engineering, Yancheng Institute of Technology, Yancheng 224051, China
2
Jiangsu Provincial Key Laboratory of Eco-Environmental Materials, Yancheng Institute of Technology, Yancheng 224051, China
3
School of Material Science and Engineering, Suzhou University of Science and Technology, Suzhou 215009, China
*
Authors to whom correspondence should be addressed.
Catalysts 2023, 13(5), 821; https://doi.org/10.3390/catal13050821
Submission received: 6 April 2023 / Revised: 26 April 2023 / Accepted: 27 April 2023 / Published: 29 April 2023
(This article belongs to the Special Issue Photocatalytic Nanomaterials for Environmental Purification)

Abstract

:
A two-dimensional layered CeO2/Bi2O3 composite was synthesized by microwave solvothermal method. X-ray diffraction (XRD), Raman spectroscopy, field emission scanning electron microscope (FESEM), transmission electron microscope (TEM), UV-Vis diffuse reflection spectroscopy (DRS), and X-ray photoelectron spectroscopy (XPS) were used to studied crystal structure, morphology, optical performance, elemental composition and the surface electronic state of the samples. The photocatalytic properties of the prepared samples were evaluated by photocatalytic desulfurization under visible light. When the molar ratio of Ce and Bi was 1:2, CeO2/Bi2O3 composite presented the highest photocatalytic desulfurization rate. Transient Photocurrent measurement, electrochemical impedance spectroscopy (EIS) and photoluminescence spectroscopy (PL) showed that CeO2 and Bi2O3 formed a heterojunction, which could promote the separation of photogenerated electrons and holes, improving the photocatalytic activity. Furthermore, it was found that the active species of hydroxyl radical (·OH) played an important role in the photocatalytic degradation of dibenzothiophene (DBT) based on the active species capture experiment. Finally, a plausible mechanism for the photocatalytic oxidative desulfurization of this nanocomposite was proposed.

1. Introduction

Nowadays, environmental pollution and climate change caused by fossil energy combustion have become the focus of global attention. In particular, the combustion of sulfides in motor vehicle fuel generates a large amount of sulfur oxides. These sulfur oxides can cause acid fog, acid rain and haze, lead to environmental pollution and threaten human health [1,2]. Therefore, it is necessary to develop deep desulfurization technology to remove the sulfide in fuel oil and decrease the emission of sulfur oxides from the source. At present, hydrodesulfurization (HDS) is the main technology, widely used in modern industry to reduce the sulfur content of fuel. It requires high pressure, high temperature and high hydrogen consumption. Dibenzothiophene (DBT) and its derivatives are the main sulfur species in diesel and gasoline. However, HDS technology experiences difficult in removing them [3,4]. In order to solve these problems, several hydrogen-free desulfurization technologies have been developed, such as oxidative desulfurization (ODS) [5], adsorptive desulfurization (ADS) [6], extractive desulfurization (EDS) [7], chemical desulfurization (CDS) [8] and biological desulfurization (BDS) [9], etc. Among these, the ODS method [10], as a deep desulfurization technology, has attracted wide attention because it can selectively convert organosulfur compound into corresponding sulfones and sulfate ions under moderate reaction conditions. In recent years, the photocatalytic oxidation desulfurization (PODS) method has shown extraordinary potential due to its special advantages, such as carbon neutrality, chemical reactions driven by photons instead of high pressure and high temperature, and the utilization of abundant available solar energy [11,12]. In this technique, oxidizing reagents such as molecular oxygen(O2) [13], hydrogen peroxide(H2O2) [14], ozone (O3) [15], and air [16] are used to convert organosulfur compounds present in motor fuels into their corresponding sulfones or SO42−, which can be removed through adsorption or extraction to achieve deep desulfurization. In addition, the activity of photocatalytic reactions mainly depends on the photocatalyst, so photocatalytic materials need to have stable structures and excellent optical properties Therefore, a series of excellent photocatalysts for photocatalytic desulfurization have been developed. Hussain’s group developed a deep aerobic photocatalytic oxidation desulfurization technology and synthesized a series of excellent photocatalysts such as LaVO4 [17,18], Ag3VO4 [19], V2O5 [20], etc.; Zhou et al. [21] prepared a Ag2O/Na-g-C3N4 heterojunction for photocatalytic desulfurization of thiophene in fuel. Belousov et al. [22] synthetized nanosized Bi2WxMO1−xO6 solid solutions to degrade DBT in model fuel; Zhang et al. [23] prepared a CeF3/g-C3N4 heterojunction photocatalyst with up-conversion performance for photocatalytic removal of DBT from model oil.
Currently, a rare earth oxide (CeO2) is considered as one of the semiconductor materials with various application prospects, especially in the field of photocatalysis, due to its unique 4f orbital and abundant electronic energy levels. However, single CeO2 still has some disadvantages, such as large band gap, low absorption in visible region and high electron–hole recombination. Nevertheless, these shortcomings can be overcome by coupling other semiconductors, promoting charge separation and increasing carrier lifetime to improve photocatalytic activity [24]. Therefore, a series of Ceria-based photocatalytic composites have been developed for photocatalytic desulfurization. Mousavi-Kamazani et al. [25] synthetized Cu2O-CeO2 nanocomposites, which has 84% photocatalytic desulfurization efficiency within 180 min under visible light; Radwan et al. [26] prepared Fe2O3-CeO2 nanocomposites, and research showed that, when the loading amount of Fe2O3 was 15%, the composite had the highest photocatalytic desulfurization activity; Chen’s group [27] prepared well-aligned a CeO2/TiO2 nanotube array photocatalyst and, when used for photocatalytic oxidation of benzo-thiophene (BT) under visible light irradiation, it was found that more than 90% of sulfur compounds in model oil were removed.
The semiconductor oxide Bi2O3 has recently attracted the attention of researchers due to its unique optical and electrical properties. It is well known that bismuth oxide has six crystal forms, denoted as α-Bi2O3 (monoclinic), β-Bi2O3 (tetragonal), δ-Bi2O3 (cubic bcc), γ-Bi2O3 (orthorhombic), ε-Bi2O3 (orthorhombic), and ῳ-Bi2O3 (triclinic), respectively. Among them, α-Bi2O3 is considered as a promising photocatalyst for water decomposition and organic photocatalytic degradation under visible light illumination, with good stability, non-toxicity and small band gap [28]. Dong et al. [29] successfully prepared a porous nanosheet structure α-Bi2O3 photocatalyst by the biomimetic-synthesis assisted hydrothermal method, which displayed efficient degradation toward different pollutant molecules; Gupta et.al. [30] synthesized α-Bi2O3 nanosheets by a simple annealing assisted thermal decomposition method, which exhibited high photocatalytic degradation of Rhodamine B. Therefore, CeO2 coupled with α-Bi2O3 to construct a heterojunction structure is expected to improve photocatalytic activity. In addition, two-dimensional (2D) semiconductor photocatalysts have been widely investigated due to their good light absorption characteristics, shorter electron and hole migration paths [31,32]. Lu et al. [33] synthesized N-doped two-dimensional CeO2-TiO2 nanosheets by a biological template method, which exhibited excellent photocatalytic desulfurization activity; Li et al. [34] utilized attapulgite-CeO2 decorated two-dimensional MoS2 to prepare nanocomposite, and the degradation rate of DBT can reach 95% under 3 h irradiation.
Herein, in this paper, we prepared a two-dimensional layered CeO2/Bi2O3 composite using a microwave solvothermal method. Then the activity of the photocatalysts were evaluated by removing DBT from the model oil. In addition, the crystal structure, morphology, optical performance, elemental composition and surface electronic state of the prepared samples were investigated by XRD, Raman spectroscopy, FESEM, TEM, DRS and XPS. Meanwhile, the separation efficiency of photogenerated electrons and holes was studied by photocurrent measurement, EIS and PL. Finally, the possible mechanism of photocatalytic desulfurization was discussed.

2. Results and Discussion

2.1. XRD Analysis

Figure 1 shows the XRD patterns of the prepared photocatalysts. The diffraction peaks near 28.6°, 33.1°, 47.5°, 56.3° and 59.1°correspond to (111), (200), (220), (311) and (222) crystal planes of cubic fluorite structure CeO2 (PDF#34-3094), respectively. Accordingly, the main characteristic peaks at 24.6° 25.8°, 26.9°, 27.4 °, 28.0 °, 33.0°, 33.2°, 34.0°, 35.0°, 37.6°, 40.0°, 42.4°, 45.1°, 46.3°, 48.6°, 52.4°,54.7°, 57.9°, 59.1°, 62.5° and 71.4 are indexed to (−102), (002), (−112), (−121), (012), (−122), (−202), (022), (−212), (−113), (−222), (−123), (023), (041), (−104), (−332), (−241), (024), (150), (104) and (−161) the crystal planes of monoclinic α-Bi2O3 (PDF#71-2274), respectively. In CeBi-2 and CeBi-1 composites, the main characteristic peaks are similar to CeO2, and some low-intensity Bi2O3 characteristic peaks appear. However, in CeBi-0.5 composite, the main characteristic peaks of the composite are similar to α-Bi2O3, and some low-intensity CeO2 characteristic peaks appear. It is worth noting that, with the increase of the Bi2O3 component, the characteristic diffraction peaks of the monoclinic α-Bi2O3 were gradually enhanced. Moreover, compared with pure CeO2 and Bi2O3, some peak positions of the composites are slightly shifted, which is due to the strong interfacial interaction between CeO2 and Bi2O3 [35]. This indicates that a heterostructure may be formed at the interface between CeO2 and Bi2O3.

2.2. Raman Spectra Analysis

Figure 2 shows the Raman characteristic peaks of CeO2, Bi2O3 and CeO2/Bi2O3 composites in the wavenumber range of 100–650 cm−1. The peak at 460 cm−1 in the pure CeO2 corresponds to the symmetrical stretching vibration of Ce-O-Ce, which belongs to the triple degenerate F2g vibration mode of CeO2 in the cubic fluorite phase [36]. However, for pure α-Bi2O3, the characteristic peak centered at 117 cm−1 is attributed to the vibration modes of bismuth atoms; while the characteristic peaks at 138 cm−1 and 150 cm−1 correspond to Bi-O stretching vibration modes; the bands at 183 cm−1, 210 cm−1, 284 cm−1, 313 cm−1, 410 cm−1, 445 cm−1 and 530 cm−1 belong to the oxygen vibration modes [37]. Additionally, in the Raman spectra of CeO2/Bi2O3 composites with different ratios, it can be found that the broad peaks of CeO2 (445 cm−1) and Bi2O3 (460 cm−1) overlap. Meanwhile, it is observed that the characteristic peak intensity of Bi2O3 diminishes with the increase in CeO2 content, all of which indicate that CeO2/Bi2O3 composites are successful prepared.

2.3. SEM, TEM, HRTEM and EDS Analysis

The morphology and microstructure of the prepared photocatalyst are studied by FESEM, TEM, HRTEM and EDS. Figure 3a,b show the FESEM images of CeBi-0.5. It can be seen that CeBi-0.5 presents an oval nanosheet structure with uniform shape and size on a large scale. The sizes of the longitudinal axis and horizontal axis are about 500 nm and 1 μm, respectively, and the nanosheets self-assemble into flower-like nanostructures. Figure 3c,d are the TEM and HRTEM images of CeBi-0.5. It is observed that the prepared composite presents a two-dimensional lamellar structure, which is composed of about 5–20 nm nanoparticles. In addition, the lattice fringes with an interplanar spacing of 0.32 nm and 0.27 nm correspond to (111) crystal planes of cubic fluorite CeO2 and (200) crystal plane of monoclinic α-Bi2O3, respectively (Figure 3d). Figure 3e shows the EDS spectrum of CeBi-0.5 composite. Only Bi, Ce and O elements are detected, and no other elements, indicating that the CeO2 and Bi2O3 are successfully combined.

2.4. Nitrogen Adsorption–Desorption Isotherm and Pore Size Distribution

The specific surface area and pore-size distribution of the photocatalysts have a major influence on the enhancement of the photocatalytic performance. Figure 4 shows the N2 adsorption–desorption isotherm of CeO2, Bi2O3 CeBi-0.5 and the pore-size distribution isotherm of CeBi-0.5. The curve of pure CeO2 presents an inverse S shape and there is no hysteresis loop, which belongs to a type II adsorption isotherm, indicating that pure CeO2 has no pores. The Bi2O3 and CeBi-0.5 composite exhibit a typical IV like isotherm with H3 hysteresis loop, indicative of the characteristic of mesopores, according to the IUPAC classification [38]. It can be seen that the CeBi-0.5 composite has a relatively narrow pore size distribution, centered at about 5–20 nm (the inset of Figure 4). The BET-specific surface areas of CeO2, Bi2O3 and CeBi-0.5 composite are measured as 4.8 m2/g, 38.7 m2/g and 52.5 m2/g, respectively. In particular, two-dimensional CeBi-0.5 composite exhibit higher surface areas than pure CeO2 and Bi2O3, which may offer a more active site in the process of photocatalytic desulfurization, enhancing the photocatalytic performance.

2.5. XPS Analysis

The chemical states and bonding environment of the elements on the surface of the composites are analyzed by XPS. Figure 5a shows the high-resolution spectrum of the Ce 3d. The emission peaks of Ce 3d are divided into eight fitting peaks, where the v and u represent spin-orbit coupling 3d 5/2 and 3d3/2, respectively. The peaks of v1 and u1 are assigned to the Ce IV (3d9 4f2) O (2p4) state; the peaks of v3 and u3 are assigned to the Ce IV (3d9 4f1) O (2p5) state and the peaks of v4 and u4 are assigned to the Ce IV (3d9 4f0) O (2p6) state. These six fitting peaks belong to the Ce4+ species. The other two fitting peaks v2 and u2 are related to the Ce III (3d9 4f2) O (2p5) state, which belong to the Ce4+ species. Therefore, there are trivalent and tetravalent cerium in CeO2 and CeBi-0.5 [39]. The presence of trivalent cerium implies the presence of more oxygen vacancies. As shown in Figure 5b, there are characteristic peaks at 159 eV and 164.3 eV, which are attributed to Bi 4 f7/2 and Bi 4 f3/2 spin orbitals, respectively [40]. This indicates the presence of trivalent bismuth in Bi2O3 and CeBi-0.5. The fitting peaks located at 529.7 eV (529.5 eV) and 531.3 eV (531.2 eV) observed in Figure 5c correspond to the lattice oxygen and surface hydroxyl oxygen in CeO2 (Bi2O3), respectively. It is noted that, compared with pure CeO2 and Bi2O3, the binding energies of Ce 3d, Bi 4f and O 1s have slightly shifted, which indicates that there is a strong electron interaction between the CeO2 and Bi2O3 interfaces [41]. Therefore, the heterojunction structure between CeO2 and Bi2O3 may be formed.

2.6. DRS Analysis

The UV-Vis DRS of CeO2, Bi2O3 and CeO2/Bi2O3 composites are shown in Figure 6a. Compared with the pure CeO2 and Bi2O3, it is observed that the absorption edge of CeO2/Bi2O3 composites shifted significantly to the visible light region. In addition, the composites have higher absorption than pure CeO2 in the visible light region, indicating that the heterojunction formed between CeO2 and Bi2O3 could improve the light absorption ability and expand the light response region. The band gap energy of a semiconductor can be calculated by the following formula [42,43]:
(α h υ) = A(h υEg)n/2
where α, h, n, Eg and A are absorption coefficient, Planck constant, light frequency, band gap energy and a constant, respectively, and n is 1 for a direct transition. Plotting (αh υ)2 versus energy () based on the spectral response in Figure 6a gives the extrapolated intercept corresponding to the Eg value (Figure 6b). The band gap values (Eg) of pure CeO2 and Bi2O3 are about 2.95 eV and 2.85 eV, respectively. Simultaneously, the Eg of the CeO2/Bi2O3 composites decreases gradually with the increase in the Bi2O3 molar ratio. Therefore, the prepared composites could make better use of visible light.

2.7. Photocatalytic Activity

The removal experiment for DBT in model oil is carried out to study the photocatalytic activity of different photocatalysts under visible light. Figure 7a shows the photocatalytic desulfurization rate of CeO2, Bi2O3 and CeO2/Bi2O3 composites. It was found that all photocatalysts have a much lower adsorption ability for DBT in the dark. In the absence of any photocatalyst, the desulfurization rate of H2O2 is about 7.5% in 3 h, indicating that H2O2 can oxidize the DBT molecular to some extent, but the oxidation ability of H2O2 is extremely limited. When there is a photocatalyst in the reaction system, it can be found that the desulfurization rate is effectively improved. It is worth noting that the Bi2O3/CeO2 composites show a significant desulfurization trend compared with the pure CeO2 and Bi2O3. When the molar ratio of Ce to Bi is 1:2, the composite shows the best photocatalytic efficiency, and its desulfurization rate reaches 90.5 % in 3 h. Meanwhile, the corresponding photocatalytic desulfurization kinetic curves over the prepared photocatalysts are shown in Figure 7b. The reaction data are fitted by a first-order model as depicted by the formula [44]:
ln(C0/Ct) = kt
where k is the pseudo-first-order rate constant, and the relationship between ln(C0/C) and catalytic reaction time t is considered as linear. It is worth noting that CeBi-0.5 shows the maximum kinetic constant (k) value of 0.81802 h−1, which is approximately 5.19-flod and 2.45-flod of pure CeO2 and Bi2O3, respectively. This indicates that CeBi-0.5 composite can effectively improve the photocatalytic activity. As shown in Figure 7c, the desulfurization rate of the CeBi-0.5 composite after three photodegradation cycles is 70%, which may be due to the loss of active components and the easy filling of the oxygen vacancy in the photocatalytic process [45]. Consequently, the excellent photocatalytic performance of the composite is attributed to its unique two-dimensional lamellar structure, which is conducive to the enrichment of organic sulfur molecules. Moreover, the formation of heterojunction between CeO2 and Bi2O3 in the composite can promote the effective separation of photogenerated electrons and holes, and can significantly improve the photocatalytic activity.

2.8. Photoelectrochemical and PL Analysis

In general, the transient photocurrent responses could evaluate the separation and migration efficiency of photogenerated electron-hole pairs in the semiconductor materials The transient photocurrent response curve is shown in Figure 8a. The photocurrent intensity of CeBi-0.5 composite is larger than those of pure CeO2 and Bi2O3, indicating that CeBi-0.5 composite exhibits the higher separation efficiency for electron-hole pairs [46]. Moreover, EIS is used to study the charge transfer resistance of the photogenerated carriers (Figure 8b). Generally speaking, the smaller the arc in the EIS Nyquist curve, the lower the charge transfer resistance on the electrode surface, and the recombination of photo generated charge carriers would be inhibited. The arc radius of CeBi-0.5 composite is significantly smaller than those of pure CeO2 and Bi2O3, indicating that the CeBi-0.5 composite is beneficial for improving the transfer efficiency of photogenerated electron-hole pairs. Fluorescence spectroscopy is also an effective approach to study the separation efficiency of photogenerated electrons and holes. From Figure 8c, it can be found that the PL intensity of CeBi-0.5 composite is lower than that of pure CeO2 and Bi2O3, indicating that the recombination of photogenerated electrons and holes can be greatly inhibited in CeBi-0.5 composite [42]. Therefore, the formation of heterojunction structures between CeO2 and Bi2O3 can effectively promote photo generation and electron separation, thereby improving photocatalytic activity.

2.9. Photocatalysis Mechanism

In general, photoinduced holes (h+), superoxide radicals (·O2−) and hydroxyl radicals (·OH) are considered to be the main active species in the photocatalytic degradation of semiconductors [19]. In order to further determine the main active species of CeO2/Bi2O3 composites in photocatalytic reactions, the trapping capture experiment for active species is carried out. Three different kinds of captors, triethanolamine (TEOA; h+ scavenger), tert-butyl alcohol (TBA; a ·OH radical scavenger) and benzoquinone (BQ; a ·O2− radical scavenger), are used [47]. As can been seen in Figure 9, the desulfurization rate is slightly decreased using TEOA and BQ as capture agents, indicating that neither h+ nor ·O2− is the main reactive species. However, the addition of TBA causes obvious deactivation of the photocatalyst, reducing the photocatalytic activity for the desulfurization rate from 90.5% to 35%, as shown in the experimental results. This clearly demonstrates that active ·OH radicals are the dominant reactive species responsible for the photocatalytic desulfurization system over these photocatalysts.
To further determine the mechanism of photocatalytic desulfurization, the positions of conductive band (CB) and valence band (VB) need to be determined, according to the following formula [48,49]:
ECB = χ − E0 − 0.5Eg
EVB = ECB + Eg
where χ is the absolute electronegativity of the semiconductor (the χ values of Bi2O3 and CeO2 are 5.99 eV and 5.56 eV, respectively.), E0 is the potential energy of the standard hydrogen electrode (4.5 eV), and Eg is the band gap of the semiconductor. Combined with Eg values of Bi2O3 and CeO2 are 2.85 eV and 2.95 eV, and the CB and VB values of Bi2O3 are calculated to be 0.07 eV and 2.92 eV, and −0.42 eV and 2.53 eV for Bi2O3.
The mechanism of photocatalytic desulfurization is shown in Figure 10. When the composite is irradiated under visible light, both Bi2O3 and CeO2 can be activated and generate electron-hole pairs, due to the CB of CeO2 being more negative than that of Bi2O3. The photogenerated electrons on the CB of CeO2 tend to migrate to that of Bi2O3 through the interface, and then the electrons will react with H2O2 to form ·OH, whereas the holes in the VB of Bi2O3 are spontaneously transferred to CeO2, where -OH can be oxidized to ·OH by the holes. Hydroxyl radicals have strong oxidation ability, which can oxidize non-polar DBT molecules adsorbed on the surface of two-dimensional materials into polar dibenzothiophene sulfone (DBTO2). Then DBTO2 can be extracted and removed, due to its strong polarity, and finally the organic sulfur in the model oil is effectively removed [50]. Based on the above discussion, we propose the following reaction equations:
CeO2/Bi2O3+ hν → e + h+
e + H2O2 → •OH + OH
h+ + OH → •OH
DBT + •OH → DBTO2

3. Experimental Section

3.1. Materials

Cerium nitrate hexahydrate (Ce(NO3)3·6H2O, 99%),Bismuth nitrate pentahydrate (Bi(NO3)3·5H2O, 99%), Acetonitrile (C2H3N, 99%), Anhydrous ethanol (CH3CH2OH,) and Tert-butyl alcohol ((CH3)3COH, TBA) were purchased from China National Medicines Corporation Ltd. Benzo-thiophene (C12H8S, DBT) and n-octane (C8H18) were purchased from Aladdin Bio-Chem Technology Co., Ltd. (Shanghai, China). Ethylene glycol ((CH2OH)2), Triethanolamine (HOCH2CH2)3N, TEOA) and Benzoquinone (C6H4O2, BQ) were purchased from Shanghai Maclean Biochemical Technology Co. Ltd. All of the reagents in this experiment were analytical grade and directly used without further purification. Distilled water was used throughout this study.

3.2. Preparation of Samples

The typical synthesis process is described as follows: 0.01 mol Bi(NO3)3·5H2O was dissolved in a mixed solution of 60 mL ethylene glycol and ethanol (VEG: VEthanol = 1:2) marked as A, and then a certain stoichiometric ratio of Ce(NO3)3·6H2O was dissolved in 10 mL deionized water, marked as B. Subsequently, solution B was slowly dropped into solution A, After the two solutions were fully mixed, the solution was relocated into a 100 mL Teflon liner and then placed in a microwave synthesizer at 160 °C for 30 min, then the precipitate was washed with distilled water and anhydrous ethanol three times, and dried in an oven at 60 °C for 12 h. The composites with molar ratios between CeO2 and Bi2O3 (2:1, 1:1, 1:2) were labeled as CeBi-2, CeBi-1 and CeBi-0.5, respectively. Pure CeO2 and Bi2O3 photocatalysts were also synthesized under similar experimental conditions.

3.3. Characterization

X-ray diffraction (XRD) measurement was performed by a Rigaku, D/max-RB instrument (Tokyo, Japan) between 20 and 80° at a scan rate of 5°/min with Cu Ka radiation; Raman spectra were measured by a Thermo Fisher Scientific DXR spectrometer (Waltham, MA, USA) and the excitation laser wavelength was 532 nm. The morphology was observed by field emission scanning electron microscopy (FESEM, Zeiss Supra55, Jena, Germany) and transition electron microscopy (TEM, JEOL JEM-2100, Tokyo, Japan), equipped with an EDS spectrometer (EDS, Oxford, UK). N2 adsorption-desorption isotherms were determined using a Micromeritics ASAP 2010 analyzer (Norcross, GA, USA), the surface areas were calculated according to the Brunauer-Emmett-Teller (BET) method, and the pore size distribution was obtained by the Barrett-Joyner-Halenda (BJH) model. X-ray photoelectron spectroscopy (XPS) analysis was carried out by a Thermo Fisher Scientific ESCALAB 250 (Waltham, MA, USA) spectrometer with mono Al Ka radiation (1486.6 eV). Photoluminescence (PL) spectra were collected on a PerkinElmer LS45 fluorescence spectrometer (Waltham, MA, USA). Ultraviolet-visible diffuse reflectance spectra (UV-Vis DRS) were measured by a Shimadzu UV-2450 spectrophotometer (Kyoto, Japan) equipped with an integrating sphere.

3.4. Photocatalytic Desulfurization Measurement

The model oil with sulfur content of 100 ppm was prepared by dissolving 0.08 g DBT into 200 mL of n-octane. Then, 0.1 g of photocatalysts were added to the above solution with constant stirring; subsequently, an appropriate 30wt% hydrogen peroxide aqueous solution (molar ratio of (H2O2/DBT = 4:1) was added. After dark adsorption for 30 min, the suspension was transferred into a photochemical reaction instrument containing a 300 W xenon lamp with an ultraviolet cut-off filter (GHX-2, Yangzhou Science and Technology City Instrument co., Ltd., China). The dispersion was collected every 30 min and extracted with acetonitrile, then the sulfur content was measured using an ultraviolet fluorescence sulfur analyzer (THA2000S, Taizhou Jinhang Analysis Instrument Co., Ltd., Taizhou, China). The desulfurization rate D (%) is obtained according to the following formula:
D = (1 − Ct/C0) × 100%
where C0 is the initial sulfur content and Ct is the sulfur content of the solution at reaction time t.

3.5. Photoelectrochemical Measurements

Photoelectrochemical analysis was performed in a standard three electrode system, with Pt foil as the counter electrode, Ag/AgCl electrode as the reference electrode and the prepared sample deposited on FTO glass as the working electrode. Photocurrent was measured on a photoelectrochemical workstation (LK5600, Tianjin Lanlike Chemical Electronics High Technology Co., Ltd., Tianjin, China) with 0.1 M Na2SO4 aqueous solution as electrolyte and 300 W xenon lamp as light source. EIS test also used the three-electrode system. The electrolyte solution was 0.5 M KCl containing 0.01 M K3Fe(CN)6/K4Fe(CN)6 (molar ratio 1:1), performed at bias voltages 0.5 V, in the frequency range of 0.1 Hz-100 kHz, with oscillation potential amplitudes of 0.01 V.

4. Conclusions

In summary, a two-dimensional layered CeO2/Bi2O3 photocatalyst was successfully synthesized by microwave solvothermal method and used for photocatalytic desulfurization. The prepared photocatalysis significantly enhances the performance of photocatalytic removal of DBT under visible light irradiation. When the molar ratio of Bi2O3 to CeO2 is 2:1, the prepared sample exhibits the best photocatalytic activity, with removal rate of DBT reaching 90.5% in3 h and the reaction rate constant at 0.81802 h−1, which is approximately 5.19-flod and 2.45-flod of pure CeO2 and Bi2O3, respectively. Free radical scavenging experiments demonstrate that hydroxyl radical is the main active species in the photocatalytic process. The enhanced photocatalytic activity is attributed to the unique two-dimensional structure of the heterojunction formed between cerium oxide and bismuth oxide. This study indicates that the construction of two-dimensional heterostructure photocatalyst is a promising method to improve photocatalytic desulfurization.

Author Contributions

Conceptualization, X.L. and Q.Z.; methodology, X.L. and J.Q.; validation, X.L. and H.H.; formal analysis, X.L., W.C. and H.H.; investigation, X.L. and W.C.; resources, X.L.; data curation, H.H. and W.C.; writing-original draft preparation, X.L. and W.C.; writing-review and editing, H.H. and J.Q.; visualization, X.L. and H.H.; supervision, X.L.; project administration, Q.Z.; funding acquisition, X.L. and Q.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by the National Natural Science Foundation of China (No. 12274361), the Natural Science Foundation of Jiangsu Province (BK20211361), College Natural Science Research Project of Jiangsu Province (20KJA430004) and the Funding for school-level research projects of Yancheng Institute of Technology (xjr2019026).

Data Availability Statement

Data available on request from the authors.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Shindell, D.; Smith, C.J. Climate and air-quality benefits of a realistic phase-out of fossil fuels. Nature 2019, 573, 408–411. [Google Scholar] [CrossRef]
  2. Shen, N.; Wang, Y.F.; Peng, H.; Hou, Z.P. Renewable energy green innovation, fossil energy consumption, and air pollution-spatial empirical analysis based on China. Sustainability 2020, 12, 6397. [Google Scholar] [CrossRef]
  3. Tanimu, A.; Alhooshani, K. Advanced Hydrodesulfurization Catalysts: A Review of Design and Synthesis. Energy Fuels 2019, 33, 2810–2838. [Google Scholar] [CrossRef]
  4. Dedual, G.; MacDonald, M.J.; Alshareef, A.; Wu, Z.; Tsang, D.C.; Yip, A.C. Requirements for effective photocatalytic oxidative desulfurization of a thiophene-containing solution using TiO2. J. Environ. Chem. Eng. 2014, 2, 1947–1955. [Google Scholar] [CrossRef]
  5. Bhadra, B.N.; Jhung, S.H. Oxidative desulfurization and denitrogenation of fuels using metal-organic framework-based/-derived catalysts. Appl. Catal. B 2019, 259, 118021. [Google Scholar] [CrossRef]
  6. Yaseen, M.; Ullah, S.; Ahmad, W.; Subhan, S.; Subhan, F. Fabrication of Zn and Mn loaded activated carbon derived from corn cobs for the adsorptive desulfurization of model and real fuel oils. Fuel 2021, 284, 119102. [Google Scholar] [CrossRef]
  7. Makoś, P.; Boczkaj, G. Deep eutectic solvents based highly efficient extractive desulfurization of fuels—Eco-friendly approach. J. Mol. Liq. 2019, 296, 111916. [Google Scholar] [CrossRef]
  8. Manan, N.S.A.; Aldous, L.; Alias, Y.; Murray, P.; Yellowlees, L.J.; Lagunas, M.C.; Hardacre, C. Electrochemistry of Sulfur and Polysulfides in Ionic Liquids. J. Phys. Chem. B 2011, 115, 13873–13879. [Google Scholar] [CrossRef] [PubMed]
  9. Capecchi, E.; Piccinino, D.; Bizzarri, B.M.; Botta, L.; Crucianelli, M.; Saladino, R. Oxidative Bio-Desulfurization by Nanostructured Peroxidase Mediator System. Catalysts 2020, 10, 313. [Google Scholar] [CrossRef]
  10. Shafiq, I.; Shafique, S.; Akhter, P.; Ishaq, M.; Yang, W.; Hussain, M. Recent breakthroughs in deep aerobic oxidative desulfurization of petroleum refinery products. J. Clean. Prod. 2021, 294, 125731. [Google Scholar] [CrossRef]
  11. Zhou, X.; Wang, T.; Liu, H.; Gao, X.; Wang, C.; Wang, G. Desulfurization through Photocatalytic Oxidation: A Critical Review. ChemSusChem 2021, 14, 492–511. [Google Scholar] [CrossRef] [PubMed]
  12. Shafiq; Shafique, S.; Akhter, P.; Abbas, G.; Qurashi, A.; Hussain, M. Efficient catalyst development for deep aerobic photocatalytic oxidative desulfurization: Recent advances, confines, and outlooks. Catal. Rev. Sci. Eng. 2022, 64, 789–834. [Google Scholar] [CrossRef]
  13. Gao, X.-M.; Fu, F.; Zhang, L.-P.; Li, W.-H. The preparation of Ag–BiVO4 metal composite oxides and its application in efficient photocatalytic oxidative thiophene. Phys. B Condens. Matter 2013, 419, 80–85. [Google Scholar] [CrossRef]
  14. Zarrabi, M.; Entezari, M.H.; Goharshadi, E.K. Photocatalytic oxidative desulfurization of dibenzothiophene by C/TiO2@MCM-41 nanoparticles under visible light and mild conditions. RSC Adv. 2015, 5, 34652–34662. [Google Scholar] [CrossRef]
  15. Boshagh, F.; Rahmani, M.; Rostami, K.; Yousefifar, M. Key Factors Affecting the Development of Oxidative Desulfurization of Liquid Fuels: A Critical Review. Energy Fuels 2021, 36, 98–132. [Google Scholar] [CrossRef]
  16. Kalantari, K.; Kalbasi, M.; Sohrabi, M.; Royaee, S.J. Synthesis and characterization of N-doped TiO2 nanoparticles and their application in photocatalytic oxidation of dibenzothiophene under visible light. Ceram. Int. 2016, 42, 14834–14842. [Google Scholar] [CrossRef]
  17. Shafiq, I.; Hussain, M.; Shafique, S.; Rashid, R.; Akhter, P.; Ahmed, A.; Jeon, J.-K.; Park, Y.-K. Oxidative desulfurization of refinery diesel pool fractions using LaVO4 photocatalyst. J. Ind. Eng. Chem. 2021, 98, 283–288. [Google Scholar] [CrossRef]
  18. Shafiq, I.; Hussain, M.; Rashid, R.; Shafique, S.; Akhter, P.; Yang, W.; Nawaz, Z.; Ahmed, A.; Park, Y.-K. Development of hierarchically porous LaVO4 for efficient visible-light-driven photocatalytic desulfurization of diesel. Chem. Eng. J. 2021, 420, 130529. [Google Scholar] [CrossRef]
  19. Mahboob, I.; Shafiq, I.; Shafique, S.; Akhter, P.; Hussain, M.; Park, Y.-K. Effect of active species scavengers in photocatalytic desulfurization of hydrocracker diesel using mesoporous Ag3VO4. Chem. Eng. J. 2022, 441, 136063. [Google Scholar] [CrossRef]
  20. Shafiq, I.; Hussain, M.; Shafique, S.; Akhter, P.; Ahmed, A.; Ashraf, R.S.; Khan, M.A.; Jeon, B.-H.; Park, Y.-K. Systematic Assessment of Visible-Light-Driven Microspherical V2O5 Photocatalyst for the Removal of Hazardous Organosulfur Compounds from Diesel. Nanomaterials 2021, 11, 2908. [Google Scholar] [CrossRef]
  21. Zhou, X.; Wang, T.; Zhang, L.; Che, S.; Liu, H.; Liu, S.; Wang, C.; Su, D.; Teng, Z. Highly efficient Ag2O/Na-g-C3N4 heterojunction for photocatalytic desulfurization of thiophene in fuel under ambient air conditions. Appl. Catal. B 2022, 316, 121614. [Google Scholar] [CrossRef]
  22. Belousov, A.S.; Suleimanov, E.V.; Parkhacheva, A.A.; Fukina, D.G.; Koryagin, A.V.; Titaev, D.N.; Lazarev, M.A. Synthesis and Characterization of Bi2WxMo1-xO6 Solid Solutions and Their Application in Photocatalytic Desulfurization under Visible Light. Processes 2022, 10, 789. [Google Scholar] [CrossRef]
  23. Lu, X.W.; Chen, F.; Qian, J.Q.; Fu, M.; Jiang, Q.; Zhang, Q.F. Facile fabrication of CeF3/g-C3N4 heterojunction photocatalysts with upconversion properties for enhanced photocatalytic desulfurization performance. J. Rare Earths 2021, 39, 1204–1210. [Google Scholar] [CrossRef]
  24. Tran, D.P.; Pham, M.-T.; Bui, X.-T.; Wang, Y.-F.; You, S.-J. CeO2 as a photocatalytic material for CO2 conversion: A review. Sol. Energy 2022, 240, 443–466. [Google Scholar] [CrossRef]
  25. Mousavi-Kamazani, M.; Ashrafi, S. Single-step sonochemical synthesis of Cu2O-CeO2 nanocomposites with enhanced photocatalytic oxidative desulfurization. Ultrason. Sonochem. 2020, 63, 104948. [Google Scholar] [CrossRef]
  26. Radwan, M.S.; Aboutaleb, W.A.; El Naggar, A.M.; El Sayed, H.A.; Shehata, M.R.; Medany, S.S. Photo-oxidative extractive desulfurization of dibenzothiofene over Fe2O3-CeO2 nanocomposites at visible light irradiation. J. Photochem. Photobiol. A Chem. 2022, 433, 114137. [Google Scholar] [CrossRef]
  27. Lu, X.W.; Li, X.Z.; Qian, J.C.; Miao, N.M.; Yao, C.; Chen, Z.G. Synthesis and characterization of CeO2/TiO2 nanotube arrays and enhanced photocatalytic oxidative desulfurization performance. J. Alloys Compd. 2016, 661, 363–371. [Google Scholar] [CrossRef]
  28. He, B.W.; Wang, Z.L.; Xiao, P.; Chen, T.; Yu, J.G.; Zhang, L.Y. Cooperative Coupling of H2O2 Production and Organic Synthesis over a Floatable Polystyrene-Sphere-Supported TiO2/Bi2O3 S-Scheme Photocatalyst. Adv. Mater. 2022, 34, 2203225. [Google Scholar] [CrossRef]
  29. Dong, Y.; Ma, A.; Zhang, D.; Gao, Y.; Li, H. Preparation of high-performance α-Bi2O3 photocatalysts and their photocatalytic activity. Surf. Innov. 2020, 8, 295–303. [Google Scholar] [CrossRef]
  30. Gupta, G.; Kaur, M.; Kansal, S.K.; Umar, A.; Ibrahim, A.A. α-Bi2O3 nanosheets: An efficient material for sunlight-driven photocatalytic degradation of Rhodamine B. Ceram. Int. 2022, 48, 29580–29588. [Google Scholar] [CrossRef]
  31. Zhao, Y.; Zhang, S.; Shi, R.; Waterhouse, G.I.; Tang, J.; Zhang, T. Two-dimensional photocatalyst design: A critical review of recent experimental and computational advances. Mater. Today 2020, 34, 78–91. [Google Scholar] [CrossRef]
  32. Liu, C.; Zhang, Q.; Zou, Z. Recent advances in designing ZnIn2S4-based heterostructured photocatalysts for hydrogen evolution. J. Mater. Sci. Technol. 2023, 139, 167–188. [Google Scholar] [CrossRef]
  33. Lu, X.; Li, X.; Chen, F.; Chen, Z.; Qian, J.; Zhang, Q. Biotemplating synthesis of N-doped two-dimensional CeO2–TiO2 nanosheets with enhanced visible light photocatalytic desulfurization performance. J. Alloys Compd. 2020, 815, 152326. [Google Scholar] [CrossRef]
  34. Li, X.Z.; Zhang, Z.S.; Yao, C.; Lu, X.W.; Zhao, X.B.; Ni, C.Y. Attapulgite-CeO2/MoS2 ternary nanocomposite for photocatalytic oxidative desulfurization. Appl. Surf. Sci. 2016, 364, 589–596. [Google Scholar] [CrossRef]
  35. Shen, C.H.; Wen, X.J.; Fei, Z.H.; Liu, Z.T.; Mu, Q.M. Novel Z-scheme W18O49/CeO2 heterojunction for improved photocatalytic hydrogen evolution. J. Colloid Interface Sci. 2020, 579, 297–306. [Google Scholar] [CrossRef]
  36. Wu, Z.L.; Li, M.J.; Howe, J.; Meyer III, H.M.; Overbury, S.H. Probing defect sites on CeO2 nanocrystals with well-defined surface planes by Raman spectroscopy and O2 adsorption. Langmuir 2010, 2, 16595–16606. [Google Scholar] [CrossRef]
  37. Schmidt, S.; Kubaski, E.T.; Volanti, D.P.; Sequinel, T.; Bezzon, V.D.N.; Beltran, A.; Tebcherani, S.M.; Varela, J.A. Effect of pressure-assisted heat treatment on photoluminescence emission of α Bi2O3 needles. Inorg. Chem. 2015, 54, 10184–10191. [Google Scholar] [CrossRef]
  38. Ravikovitch, P.I.; Neimark, A.V. Characterization of nanoporous materials from adsorption and desorption isotherms. Colloids Surf. A 2001, 187–188, 11–21. [Google Scholar] [CrossRef]
  39. Anandan, C.; Bera, P. XPS studies on the interaction of CeO2 with silicon in magnetron sputtered CeO2 thin films on Si and Si3N4 substrates. Appl. Surf. Sci. 2013, 283, 297–303. [Google Scholar] [CrossRef]
  40. Maheshwaran, S.; Balaji, R.; Chen, S.-M.; Chang, Y.-S.; Tamilalagan, E.; Chandrasekar, N.; Ethiraj, S.; Kumar, M. Ultrasensitive electrochemical detection of furazolidone in biological samples using 1D-2D BiVO4@MoS2 hierarchical nano-heterojunction composites armed electrodes. Environ. Res. 2022, 205, 112515. [Google Scholar] [CrossRef]
  41. Shen, C.H.; Chen, Y.; Xu, X.J.; Li, X.Y.; Wen, X.J.; Liu, Z.T.; Xing, R.; Guo, H.; Fei, Z.H. Efficient photocatalytic H2 evolution and Cr (VI) reduction under visible light using a novel Z-scheme SnIn4S8/CeO2 heterojunction photocatalysts. J. Hazard. Mater. 2021, 416, 126217. [Google Scholar] [CrossRef] [PubMed]
  42. Han, N.; Xu, Q.; Beyene, G.; Zhang, Q.F. Enhanced photocatalytic activity over g-C3N4/(BiO)2(OH)xCl2−x Z-scheme heterojunction. Appl. Surf. Sci. 2020, 521, 146464. [Google Scholar] [CrossRef]
  43. Liu, C.; Xiao, W.; Yu, G.Y.; Wang, Q.; Hu, J.W.; Xu, C.B.; Du, X.Y.; Xu, J.G.; Zhang, Q.F.; Zou, Z.G. Interfacial engineering of Ti3C2 MXene/CdIn2S4 Schottky heterojunctions for boosting visible-light H2 evolution and Cr (VI) reduction. J. Colloid Interface Sci. 2023, 640, 851–863. [Google Scholar] [CrossRef] [PubMed]
  44. Yusuff, A.S.; Popoola, L.T.; Aderibigbe, E.I. Solar photocatalytic degradation of organic pollutants in textile industry wastewater by ZnO/pumice composite photocatalyst. J. Environ. Chem. Eng. 2020, 8, 103907. [Google Scholar] [CrossRef]
  45. Li, L.; Guo, C.; Ning, J.; Zhong, Y.; Chen, D.; Hu, Y. Oxygen-vacancy-assisted construction of FeOOH/CdS heterostructure as an efficient bifunctional photocatalyst for CO2 conversion and water oxidation. Appl. Catal. B Environ. 2022, 613, 764–774. [Google Scholar] [CrossRef]
  46. Banerjee, R.; Pal, A.; Ghosh, D.; Ghosh, A.B.; Nandi, M.; Biswas, P. Improved photocurrent response, photostability and photocatalytic hydrogen generation ability of CdS nanoparticles in presence of mesoporous carbon. Mater. Res. Bull. 2021, 134, 111085. [Google Scholar] [CrossRef]
  47. Li, X.Z.; Zhu, W.; Lu, X.W.; Zuo, S.X.; Yao, C.; Ni, C.Y. Integrated nanostructures of CeO2/attapulgite/g-C3N4 as efficient catalyst for photocatalytic desulfurization: Mechanism, kinetics and influencing factors. Chem. Eng. J. 2017, 326, 87–98. [Google Scholar] [CrossRef]
  48. Li, Y.; Wang, J.; Yao, H.; Dang, L.; Li, Z. Chemical etching preparation of BiOI/Bi2O3 heterostructures with enhanced photocatalytic activities. Catal. Commun. 2011, 12, 660–664. [Google Scholar] [CrossRef]
  49. Channei, D.; Inceesungvorn, B.; Wetchakun, N.; Ukritnukun, S.; Nattestad, A.; Chen, J.; Phanichphant, S.J.S.R. Photocatalytic Degradation of Methyl Orange by CeO2 and Fe–doped CeO2 Films under Visible Light Irradiation. Sci. Rep. 2014, 4, srep05757. [Google Scholar] [CrossRef]
  50. Wang, C.; Zhu, W.S.; Xu, Y.H.; Xu, H.; Zhang, M.; Chao, Y.H.; Yin, S.; Li, H.M.; Wang, J.G. Preparation of TiO2/g-C3N4 composites and their application in photocatalytic oxidative desulfurization. Ceram. Int. 2014, 40, 11627–11635. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of CeO2, Bi2O3 and CeO2/Bi2O3 composites.
Figure 1. XRD patterns of CeO2, Bi2O3 and CeO2/Bi2O3 composites.
Catalysts 13 00821 g001
Figure 2. Raman spectra of CeO2, Bi2O3 and CeO2/Bi2O3 composites.
Figure 2. Raman spectra of CeO2, Bi2O3 and CeO2/Bi2O3 composites.
Catalysts 13 00821 g002
Figure 3. FESEM of images of CeBi-0.5 (a,b); TEM (c) and HRTEM (d) images of CeBi-0.5; EDS spectrum of CeBi-0.5 (e).
Figure 3. FESEM of images of CeBi-0.5 (a,b); TEM (c) and HRTEM (d) images of CeBi-0.5; EDS spectrum of CeBi-0.5 (e).
Catalysts 13 00821 g003
Figure 4. N2 adsorption–desorption isotherm of CeO2, Bi2O3 CeBi-0.5 and the pore-size distribution isotherm of CeBi-0.5.
Figure 4. N2 adsorption–desorption isotherm of CeO2, Bi2O3 CeBi-0.5 and the pore-size distribution isotherm of CeBi-0.5.
Catalysts 13 00821 g004
Figure 5. High-resolution XPS spectrum of (a) Ce 3d, (b) Bi 4f, (c) O 1s.
Figure 5. High-resolution XPS spectrum of (a) Ce 3d, (b) Bi 4f, (c) O 1s.
Catalysts 13 00821 g005
Figure 6. (a) UV-Vis DRS of CeO2, Bi2O3 and CeO2/Bi2O3 composites; (b) Tauc plots of (αhv)2 vs photon energy (hv).
Figure 6. (a) UV-Vis DRS of CeO2, Bi2O3 and CeO2/Bi2O3 composites; (b) Tauc plots of (αhv)2 vs photon energy (hv).
Catalysts 13 00821 g006
Figure 7. (a) Photocatalytic desulfurization rate of different catalysts; (b) Kinetic fitting curve; (c) Photocatalytic desulfurization cycle over the CeBi-0.5 composite.
Figure 7. (a) Photocatalytic desulfurization rate of different catalysts; (b) Kinetic fitting curve; (c) Photocatalytic desulfurization cycle over the CeBi-0.5 composite.
Catalysts 13 00821 g007
Figure 8. (a) Transient photocurrent responses curve, (b) Nyquist impedance plots, (c) PL spectra of CeO2, Bi2O3 and CeBi-0.5 composite.
Figure 8. (a) Transient photocurrent responses curve, (b) Nyquist impedance plots, (c) PL spectra of CeO2, Bi2O3 and CeBi-0.5 composite.
Catalysts 13 00821 g008
Figure 9. Photocatalytic desulfurization by different scavengers over CeBi-0.5.
Figure 9. Photocatalytic desulfurization by different scavengers over CeBi-0.5.
Catalysts 13 00821 g009
Figure 10. Photocatalytic desulfurization mechanism of CeO2/Bi2O3 nanocomposite.
Figure 10. Photocatalytic desulfurization mechanism of CeO2/Bi2O3 nanocomposite.
Catalysts 13 00821 g010
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

Lu, X.; Chen, W.; Hou, H.; Qian, J.; Zhang, Q. Preparation of Two-Dimensional Layered CeO2/Bi2O3 Composites for Efficient Photocatalytic Desulfurization. Catalysts 2023, 13, 821. https://doi.org/10.3390/catal13050821

AMA Style

Lu X, Chen W, Hou H, Qian J, Zhang Q. Preparation of Two-Dimensional Layered CeO2/Bi2O3 Composites for Efficient Photocatalytic Desulfurization. Catalysts. 2023; 13(5):821. https://doi.org/10.3390/catal13050821

Chicago/Turabian Style

Lu, Xiaowang, Wenxuan Chen, Haijun Hou, Junchao Qian, and Qinfang Zhang. 2023. "Preparation of Two-Dimensional Layered CeO2/Bi2O3 Composites for Efficient Photocatalytic Desulfurization" Catalysts 13, no. 5: 821. https://doi.org/10.3390/catal13050821

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