Next Article in Journal
A Brønsted Acidic Deep Eutectic Solvent for N-Boc Deprotection
Next Article in Special Issue
Xanthan Gum-Mediated Silver Nanoparticles for Ultrasensitive Electrochemical Detection of Hg2+ Ions from Water
Previous Article in Journal
Methane-Assisted Iron Oxides Chemical Looping in a Solar Concentrator: A Real Case Study
Previous Article in Special Issue
Highly Active Ag-Cu Nanocrystal Catalyst-Coated Brewer’s Spent Grain Biochar for the Mineralization of Methyl Orange and Methylene Blue Dye Mixture
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Visible-Light-Driven CO2 Reduction into Methanol Utilizing Sol-Gel-Prepared CeO2-Coupled Bi2O3 Nanocomposite Heterojunctions

1
Chemistry Department, Faculty of Science, King Abdulaziz University, P.O. Box 80203, Jeddah 21589, Saudi Arabia
2
Nanomaterials and Nanotechnology Department, Advanced Materials Institute, Central Metallurgical R&D Institute (CMRDI), P.O. Box 87, Cairo 11421, Egypt
3
Chemical and Materials Engineering Department, King Abdulaziz University, Jeddah 21589, Saudi Arabia
4
King Abdulaziz City for Science and Technology, Riyadh 11442, Saudi Arabia
5
College of Chemistry and Molecular Sciences, Wuhan University, Wuhan 430072, China
*
Authors to whom correspondence should be addressed.
Catalysts 2022, 12(11), 1479; https://doi.org/10.3390/catal12111479
Submission received: 11 September 2022 / Revised: 16 November 2022 / Accepted: 16 November 2022 / Published: 19 November 2022
(This article belongs to the Special Issue Nanoparticles in the Catalysis)

Abstract

:
Carbon dioxide (CO2) photoreduction into renewable fuels over semiconductor photocatalysts has emerged as a green and sustainable alternative for energy production. Consequently, tremendous efforts are being performed to develop robust and sustainable photocatalysts. Therefore, visible-light active nanocomposite photocatalysts composed of 5.0–20.0 wt.% bismuth oxide (Bi2O3) and cerium oxide (CeO2) were synthesized by a sol-gel-based process. The prepared nanocomposites were evaluated for the promoted photocatalytic reduction of CO2 into methanol (CH3OH). Various characterizations of the obtained photocatalysts exposed an outstanding development of crystalline structure, morphology, and surface texture due to the presence of Bi2O3. Moreover, the absorbance of light in the visible regime was improved with enhanced charge separation, as revealed by the exploration of optical response, photoluminescence, and photocurrent measurements. The overall bandgap calculations revealed a reduction to 2.75 eV for 15% Bi2O3/CeO2 compared to 2.93 eV for pure CeO2. Moreover, the adjusted 2.8 g L−1 dose of 15% Bi2O3/CeO2 selectively produced 1300 μmol g−1 CH3OH after 9 h of visible light irradiation. This photocatalyst also exhibits bearable reusability five times. The improved progression of 15% Bi2O3/CeO2 is denoted by significant charge separation as well as enhanced mobility. This study suggests the application of metal oxide-based heterojunctions for renewable fuel production under visible light.

Graphical Abstract

1. Introduction

With the development of the industrial revolution in the last two centuries, several industries have evolved, such as cement [1], fertilizers [2], and aerosols [3]. In addition, the rapid growth of the population with not enough environmental awareness has led to the release of massive waste materials through several activities [4]. The release of greenhouse gases proportionally increased and contributed to the significant rise in the earth’s temperature, with a threat value close to 1.5 °C [5,6]. Thus, reducing, capturing, or recycling greenhouse gases is a matter of life or death for humanity. Carbon dioxide gas (CO2) is the primary source of global warming that is currently humped due to various human activities ranging from burning fossil fuels to the rapid development of several industries [7,8]. There are several approaches for reducing CO2 emissions through the use of renewable energy sources or recycling the released CO2 into value-added products [9,10,11,12,13]. The photoconversion of CO2 into valuable and renewable hydrocarbon fuels on semiconductor photocatalysts has caused substantial apprehension toward the clean invention of alternative energy resources [14,15,16,17]. The photoreduction of CO2 either in the gas or liquid phase is an ecological pathway owing to its simplicity and sustainability, particularly regarding conservational topics [18,19]. The catalytic conversion of CO2 into value-added products, such as methanol (CH3OH) or other organic materials, is an additional benefit to the global warming issue [17,20,21]. As a consequence, plentiful research is currently being performed to advance this approach by utilizing simulated and natural solar light [22,23,24]. In the meantime, abundant ceramic-based photocatalysts demonstrate effectual progress in several fields, such as pollutant degradation and fuel production [25,26,27,28,29]. Notably, TiO2 and ZnO represent famous oxides for fuel production and the photocatalytic reduction of CO2 owing to their suitable reduction potential balanced within their bandgap energies (Eg) [30,31,32,33]. However, the extended Eg, photostability in solutions, and the rapid recombination of the photoinduced charges are key issues in the realization of these oxides [34,35]. Therefore, there are several efforts regarding the coupling of these oxides with more stable ceramics to enhance their stability [31,36,37,38]. The use of other stable ceramic oxides with comparable Eg represents an alternative path [39,40].
Interestingly, the use of rare-earth metal oxides such as cerium oxide (CeO2) is promising in photocatalysis. In a recent study by Hezam et al., the nanostructured CeO2 with controlled oxygen vacancies could be efficient for the photoreduction of CO2 into CH3OH [41]. Moreover, CeO2 has narrower Eg (2.8~3.1 eV) than TiO2 and ZnO, exhibiting an n-type character and representing a more stable ceramic oxide with potential applications in various photocatalytic reactions, including water splitting [42], organic transformation reactions [43], and pollutant degradation [44]. Moreover, the coupling with p-type oxides such as bismuth oxides (Bi2O3) showed a significant photocatalytic performance in dye degradation [37,45], self-cleaning coatings [46], and NO oxidation [47]. In a recent study performed by Yang et al., the β-Bi2O3-loaded CeO2 heterostructure photocatalyst exhibited a significant redox for the oxidation of stable antibiotic molecules due to the enhanced charge separation of the added β-Bi2O3 in the heterojunction system [48]. Moreover, Masula et al. investigated the improved photocatalytic performance of Bi2O3/CeO2 nanocomposites for the mineralization of wastewater pollutants in a short time due to their enhanced visible-light absorbance [49]. However, there are no studies regarding the photoreduction of CO2 as far as we know. It could be of great scientific and industrial benefit regarding the ability of this nanocomposite in the renewable reforming of CO2 into valuable and usable fuels. In this regard, nanostructured CeO2 was simply prepared using the polymeric surfactant-assisted sol-gel method and subsequently coupled with 5.0~20.0 wt.% of Bi2O3. The prepared Bi2O3/CeO2 heterojunctions featured a mesoporous surface texture, amended light harvesting, and superior photogenerated charge separations as concluded from the detailed characterization. The photoreduction of CO2 into selective CH3OH production over the optimized 15% Bi2O3/CeO2 presented a generation rate of 144.4 μmol g−1 h−1 and demonstrated robust stability for regeneration.

2. Results and Discussion

2.1. Structural, Surface, and Optical Investigations of the Produced Photocatalysts

The formed materials were identified by XRD graphs as described in Figure 1 for the pure sample in a compared to the Bi3+-added samples in b, c, d, and e for 5, 10, 15, and 20 wt.%. The pure sample a was identically identified as pure CeO2 by the coincident 2θ values of 28.5°, 32.8°, 47.5°, 56.0°, and 59.1° to the (1 1 1), (2 0 0), (2 2 0), (3 1 1), and (2 2 2) planes of cubic CeO2 as described in JCPDS card No. 34–0394, respectively [50]. Similarly, the 5% Bi3+ precursor addition (represented by b) did not alter the overall pattern due to the limited detection of minor contamination of the XRD instrument. However, the further addition of the Bi3+ precursor in c~e resulted in the appearance of new diffraction peaks, mainly existing at 27.6°, 35.2°, 39.1°, 45.1°, and 45.4° assigned to (201), (200), (321), (222), and (421) of the β-phase of Bi2O3 (JCPDS No. 76–0147) [20]. Thus, the Bi2O3/CeO2 nanocomposite formation was revealed by XRD depiction [51].
TEM images describing the morphology of the pure CeO2 compared with various Bi2O3-added CeO2 nanostructures can be seen in Figure 2. The pure CeO2 indicated the formation of irregular-shaped nanocrystals with well-defined edges (Figure 2a). The particle size of the CeO2 nanocrystals ranged between 20 and 40 nm. The magnified elected area at a high resolution (Figure 2b) confirmed the face-centered cubic phase of CeO2 owing to the d-spacing measurement of 0.31 nm representing the (111) plane [46]. The Bi2O3-loading of CeO2 revealed certain changes to the structure and morphology as seen in Figure 2c–h. In Figure 2c, representing 5% Bi2O3/CeO2, small spherical nanoparticles were seen side-by-side with the CeO2 nanocrystals of <10 nm as represented by black–white arrows [51]. The high-resolution TEM image of the selected part of this revealed the formation of Bi2O3 nanocrystals due to the measured d-spacing (0.2 nm) of the (222) plane as shown in Figure 2d [48]. On the other hand, the higher content of Bi2O3 at 10 wt.% (Figure 2e) and 15 wt.% (Figure 2g) showed an elongation of its Bi2O3 structure assigned by the representative black–white arrows. The Bi2O3 was grown in a rod-like morphology but maintained the same crystalline phase as confirmed by the d-spacing measurements in the high-resolution images (Figure 2f,h) without any effect on the CeO2 nanocrystals. This refers to the fact that the solution-based β-Bi2O3 is usually grown in a rod-like structure [20,40,49]. These findings indicate the successful fabrication of the Bi2O3/CeO2 heterojunction. The close contact between the Bi2O3 and CeO2 could enhance the charge transfer amongst them and explain the efficient photocatalytic activity toward the production of CH3OH [48,49,51].
To clarify the surface texture of the obtained Bi2O3/CeO2 heterojunctions, the N2 adsorption–desorption isotherms comparing pure CeO2 and 15% Bi2O3/CeO2 photocatalysts are shown in Figure 3. Both samples exhibited minor adsorption at a relative pressure (P/P0) of 0.65 and an H2-hysteresis loop with type IV until P/P0 = 0.94 [45]. This finding designates that the prepared samples parade a mesoporous structure. The inclusion of Bi2O3 has no significant effect on the superficial structure but slightly changes the SBET and pore diameters (inset of Figure 3). The determined SBET is comparatively decreased by including Bi2O3 to CeO2 from 191 to 169 m2 g−1 (for 15% Bi2O3/CeO2, Table 1) owing to the spread of smaller-sized Bi2O3 on the surface of CeO2. However, the pore size is almost the same (8.92 nm for CeO2 and 8.87 nm for 15% Bi2O3/CeO2) as seen in the inset of Figure 3.
The surface composition of the 15% Bi2O3/CeO2 sample was analyzed via XPS as seen in Figure 4. The Bi 4f the high-resolution spectrum (Figure 4a) was deconvoluted into two oxidation states for the majority of Bi3+ at binding energies of 164.2 and 159.2 eV for 4f5/2 and 4f7/2, respectively, and a minority of Bi5+ at 165.0 and 160.5 eV for the corresponding mentioned orbitals [49]. The Ce 3d spectrum (Figure 4b) was deconvoluted into the co-existing minority Ce3+ and the majority of Ce 4+ [44]. The Ce3+ is located at 902.1 and 886.5 for 3d3/2 and 3d5/2 orbitals [46]. On the other hand, the major peaks located at 917.2, 908.2, and 898.7 eV are attributed to the 3d3/2 binding of Ce4+ while the binding energies positioned at 890.2 and 883.1 eV correspond to the 3d5/2 of Ce4+ in CeO2. In Figure 4c, the O1s spectrum is deconvoluted into three peaks at 532.4, 530.1, and 529.0 eV assigned to the adsorbed oxygen species, the lattice bonding Bi–O of β-Bi2O3, and the lattice bonding Ce–O in CeO2, respectively [47]. Thus, the confirmation of the formed Bi2O3/CeO2 heterojunction is established by the XPS study. The quantitative analysis using the instrumental software showed the content of Bi2O3 in this sample at 13.9 wt.%.
The light harvesting by the synthesized photocatalysts was also investigated as displayed in Figure 5. The pristine CeO2 disclosed a minor visible-light absorbance with an absorption edge at 423 nm and an estimated Eg of 2.93 eV [43]. However, the synthesized Bi2O3/CeO2 showed an extension of the visible-light harvesting as the absorption edges shifted toward longer wavelengths due to the low-bandgap Bi2O3 [48]. The presence of 5, 10, 15, and 20 wt.% Bi2O3 in CeO2 redshifted the absorption edges to 427, 442, 453, and 454 nm, with corresponding Eg values of 2.89, 2.82, 2.75, and 2.74 eV, respectively (Table 1, inset of Figure 5). Thus, the presence of Bi2O3 could enhance light harvesting and associated photocatalytic performance [37,49].

2.2. Photoreduction of CO2 over Bi2O3/CeO2 Heterojunctions

The photocatalytic progression of the produced Bi2O3/CeO2 is evaluated for the reduction of CO2 as described in Figure 6 compared to the pure CeO2. In Figure 6a, the photocatalytic generation of CH3OH of Bi2O3/CeO2 is much higher than the pure CeO2 after 9 h of light irradiation as seen in Table 1. The production rate of CH3OH reached 120.0 and 122.2 μmol g−1 h−1 by 15 and 20% Bi2O3/CeO2, respectively, which is almost 15.4 and 15.7 times higher than the pure CeO2 at the same 2.0 gL−1 dose (7.8 μmol g−1 h−1, Figure 6b). As a result, the photoreduction of CO2 into CH3OH can be surged by governing the amount of Bi2O3. The 15 wt.% of Bi2O3 is chosen as the optimal included amount in the Bi2O3/CeO2 heterojunction based on the above results.
Photoluminescence (PL) spectral emission and photocurrent intensity are regarded as powerful means for understanding the photocatalytic activity promotion of the 15% Bi2O3/CeO2 as seen in Figure 7 [41,52,53]. In Figure 7a, we can see that the pure CeO2 exhibited an intensive PL peak positioned at 406.3 nm due to the large recombination of the photogenerated charges [44]. The introduction of 5, 10, 15, and 20 wt.% Bi2O3 amounts clearly quenched the PL intensity by 12, 21, 32, and 32.1%, respectively. These results indicate the suppression of the charge carrier’s recombination rates. So, the catalytic activity could be increased by the long-life charge carriers as proved in previous PL quenching due to the precise Bi2O3 occurrence in the 15% Bi2O3/CeO2 heterojunction. Moreover, transient photocurrent measurements have also established the best photocatalytic performance of the same 15% Bi2O3/CeO2 nanocomposite as presented in Figure 7b and Table 1. The 15% Bi2O3/CeO2 exhibited the largest photocurrent value (51.54 µA cm−2), which is almost 4.5 times higher than the photocurrent generated from pure CeO2. This finding reveals the promoted charge mobility due to the superior separation of the photoinduced charge carriers. Accordingly, the photoreduction rate of CO2 could be enhanced using this optimal photocatalyst.
The dose tuning plays a key role in the photoactivity of nanocomposite photocatalysts due to the upsurging of photoactive sites that are balanced with the adsorbed target molecules and light photons [54,55,56,57,58]. Figure 8a displays the impact of dose alteration (0.7~3.5 g L−1) of 15% Bi2O3/CeO2 on the photocatalytic generation of CH3OH. The dose-dependent generation of CH3OH parades the comparative upsurge from 608 (±19.5) to a maximum of 1300 (±41.6) µM g−1 for 0.7 and 2.8 g L−1, correspondingly, after 9 h of visible-light illumination. This outcome exhibits the improvement of the photocatalytic active spots bared to CO2 molecules in the case of 2.8 g L−1 compared to other doses. Nevertheless, the extra dose of the 15% Bi2O3/CeO2 at 3.5 g L−1 decreased the CH3OH generation to 1230 (±39.4) μmol g−1. Accordingly, the dosage accumulation could moderately impede the light photons from reaching the photocatalyst’s surface, resulting in a reduction of the overall activity [59,60]. Thus, 2.8 g L−1 is the optimized dose for the 15% Bi2O3/CeO2 that exhibits the highest generation rate of CH3OH as described in Figure 8b and proven by chromatograms in Figure S2 (supplementary materials). The spent 15% Bi2O3/CeO2 photocatalyst was also tested for recyclability as presented in Figure 8c. The outcomes showed that the optimal 15% Bi2O3/CeO2 validates five-time reusability for the CO2 reduction under visible light to CH3OH at 1274 μmol g−1, which is 98% of its initial value in the first run. The XRD and TEM investigations for the spent material did not show any significant structural changes (Figure S3, supplemental data). Therefore, the 15% Bi2O3/CeO2 exhibits a workable photocatalytic application for CO2 reduction beneath visible light.
Finally, the photoreduction pathway could be proposed as seen in Figure 9 based on previous discussions. The represented band diagrams of p-type Bi2O3 and n-type CeO2 can be calculated by the following formula (supplementary materials S-4):
ECB = χ − Ee − 0.5Eg
EVB = ECB + Eg
where EVB and ECB are the valence (VB) and conduction (CB) band potentials versus the standard hydrogen electrode (NHE), respectively. Ee is the energy of free electrons and χ is the electronegativity of the semiconductor photocatalyst, which is determined in supplementary materials [61,62,63]. Thus, upon illumination of both semiconductors by visible light, the photoinduced charge carriers, namely electrons (es), are generated at −0.405 and +0.535 eV corresponding to the CB of Bi2O3 and CeO2, respectively, while the holes (h+s) were located at +2.525 and +2.93, corresponding to VB of Bi2O3 and CeO2, respectively. Since the p-n heterojunction is formed between p-type Bi2O3 and n-type CeO2, an electric field (signed as E) is designed at the interface of the formed heterojunction and band-bending occurs due to the unified Fermi levels [64,65,66]. Thus, the excited es at the CB of p-type Bi2O3 could recombine with the h+s left at the VB of n-type CeO2 by the driving electric field effect [67]. On the other hand, the high redox potential es on the CB of CeO2 and h+s on the VB of Bi2O3 can contribute to the photoreduction of CO2 as follows: The H2 molecules could react with the h+s resulting in the splitting reaction to protons, oxygen, and free es as shown in Equation (3) [68,69].
3 H 2 O + h VB + ( Bi 2 O 3 ) 6 H + + 3 2 O 2 + 6 e
At last, at the CB of CeO2, which is very close to the reduction potential of CO2 into CH3OH [68,70], the produced protons and es in (3) can selectively convert the CO2 to CH3OH as in (4)
6 e + 6 H + + CO 2   ( CB   of   CeO 2 ) CH 3 OH + H 2 O
Therefore, the improved, recyclable, and efficient 15% Bi2O3/CeO2 heterojunctions reveal the promoted photoconversion of CO2 to an alternative renewable fuel under visible-light illumination.

3. Experiments

3.1. Chemicals

All materials were purchased from Merck in this work without subsequent treatment. The Ce and Bi precursors were Ce(NO3)3·6H2O (99.999% trace metals basis) and Bi(NO3)3·5H2O (99.999% trace metals basis). The polymeric non-ionic surfactant, namely Poly(ethylene glycol)-block-poly(propylene glycol)-block-poly(ethylene glycol) (Pluronic F-108, MW~14,600), was utilized as the growth-limiting reagent and micelle template [71]. Acetic (CH3COOH, ≥99.7%) and hydrochloric (HCl, ACS reagent, 37%) acids were used in the sol-gel process for hydrolysis and esterification procedures. Ethanol absolute (CH3CH2OH, ≥99.8%) and deionized water (DI-H2O, >18.2 MΩ) were also used for hydrolysis and washing purposes. Sodium carbonate (Na2CO3, ≥99.5%, ACS reagent) was used as the CO2 source in the photoreaction system.

3.2. Synthesis of CeO2 and Bi2O3/CeO2 Photocatalysts

Nanocrystalline CeO2 particles were synthesized by a modified sol-gel method in the presence of Pluronic F-108. In detail, 2.0 g of F-108 was disseminated in 40 mL of CH3CH2OH under vigorous stirring. After that, 12 mmol of HCl and 40 mmol CH3COOH were progressively poured under the same conditions and then left for 1h. Then, 1.63 g of Ce(NO3)3·6H2O was included in the above viscous solution and kept for another 2 h. The cloudy suspended mixture was then transferred into a clean Petri dish for 12 h of polymerization in a humidity chamber operated at 40 °C before drying at 65 °C for another 24 h. The obtained hard gel was crystallized for 5 h at 650 °C to obtain mesostructured CeO2 nanoparticles. The Bi2O3/CeO2 heterojunctions were prepared by impregnating the Bi2O3 into the CeO2 matrix. In brief, 1.0 g of synthesized CeO2 was sonicated in 100 mL of CH3CH2OH for 10 min. Next, 52 mg of Bi(NO3)3·5H2O was familiarized into the solution while stirring and maintained for 1 h. Then, CH3CH2OH was evaporated by increasing the temperature to 80 °C. After approximately 5 h, the calcination of the dried powder was performed at 600 °C for 3 h to crystallize the 5% Bi2O3/CeO2 heterojunction. The above procedure was repeated by increasing the Bi3+ precursor amount to 104, 156, and 208 mg to obtain 10, 15, and 20% Bi2O3/CeO2 nanocomposites.

3.3. Depictive Characterization of the Obtained CeO2 and Bi2O3/CeO2

Numerous instruments have been employed to characterize the produced sample. The Bruker AXS D8 X-ray diffraction (XRD, Billerica, MA, USA) unit explored the crystalline phases of the ground powders. The surface composition of the 15% Bi2O3/CeO2 was investigated by the Thermo VG Multilab 2000 X-ray photoelectron spectrometer (XPS, East Grinstead, UK). Regarding the morphology study, a high-resolving-power JEOL-JEM-1230 transmission electron microscope (TEM, Tokyo, Japan) examined the samples dispersed on carbon-coated grids. The Nova Quantachrome analyzer generated N2 adsorption–desorption (Ostfildern-Scharnhausen, Germany) isotherms to describe the surface texture of the degassed samples at 150 °C. The specific surface area (SBET) was calculated based on the Brunauer–Emmett–Teller (BET) formula, which estimates the specific surface area of the prepared structures. The light harvesting of CeO2 and Bi2O3/CeO2 was studied via a Jasco spectrophotometer (V-570, Tokyo, Japan) in the 200–800 nm spectral range. The bandgap energy (Eg) was determined based on the modified Tauc formula through their diffusive reflectance. Photoluminescence emission (PL) was investigated by an RF-5301 fluorescence spectrometer (Shimadzu, Kyoto, Tapan) in ambience. Lastly, the electrochemical workstation (Zahner Zennium, Ostfildern-Scharnhausen, Germany) measured the transient photocurrents of the 0.3 × 0.3 cm2 working electrodes composed of each sample on Ti foil at zero bias during the ON and OFF states of 1.0 mW cm−2 light illumination.

3.4. Setup for Photoreduction of CO2

The photoreduction experiments were performed in a homemade photoreactor containing a 250 mL two-neck quartz flask. The simulated visible light was provided by a 300 W Xe lamp (Newport, CA, USA, 45.2 mW cm−2) equipped with a 420 nm cut-off filter. In detail, a certain dose of the synthesized photocatalyst was dispersed in 125 mL of H2O before being heated to 80 °C. Then, 135 mg of Na2CO3 was inserted into the above suspension. Moreover, nitrogen gas was initially introduced for 30 min to ensure an inert atmosphere at 100 sccm. After that, 0.315 mL of 4 M HCl was gently injected to generate CO2. The whole system was kept in murky conditions for 1 h to settle the equilibrium. The photoreactor was then irradiated by visible light. At fixed one-hour intervals, a gas chromatograph (GC-FID, Shimadzu) equipped with a flame ionization detector was employed to sense the formed products through the photoreaction [68]. The produced spectra generated by the detection unit were compared with the standard GC spectra of 2.5 mM L−1 CH3OH solution (supplementary Figure S1) [21]. The produced CH3OH was recognized in the above-mentioned spectra under the same experimental conditions using the GC. Generally, neither CH3OH nor other products are detected before light illumination. The reusability of the spent photocatalyst for five independent cycles was performed via simple drying under ambient conditions for 1 h at 120 °C.

4. Conclusions

A facile sol-gel-based method was applied for the fabrication of the Bi2O3/CeO2 heterojunction photocatalyst. The precise addition of p-type Bi2O3 amended the absorbance in the visible light regime through the reduction of the overall bandgap of the produced Bi2O3/CeO2 heterojunction down to 2.74 eV. The various Bi2O3-included CeO2 were applied for the selective photocatalytic reduction of CO2 into methanol beneath visible light. The 2.8 g L−1 dose of 15% Bi2O3/CeO2 generated 1300 μmol/g of CH3OH within 9 h of illumination, nearly 18.6-fold higher than pure CeO2. This optimized heterojunction verified the five-time reusability at 98% of its initial performance. The substantial photoactivity of Bi2O3/CeO2 is attributed to the significant photogenerated charge separation in addition to the amended harvesting of light. This work provides plenty of room for the future of renewable fuel production from CO2 utilizing metal oxide-based heterostructure photocatalysts.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal12111479/s1, Figure S1: Gas chromatograph of standard 2.5 mM/L CH3OH solution showing a single peak at 2.46 min retention. The spectra is used as a reference to compare the selective photoreduction of CO2 over the prepared Bi2O3/CeO2 nanocomposites in the main manuscript; Figure S2: Gas chromatograph of the produced product over the optimized dose 2.8 g/L of 15% Bi2O3/CeO2 showing the accumulated production of CH3OH over time of irradiation; Figure S3: XRD patterns (a) and TEM images (b) of the spent nanocomposite after the first and the fifth runs displaying no substantial change in the crystalline phase or morphological structure after recycling for five times.

Author Contributions

Conceptualization, M.M.M.M., A.S and A.D.C.; methodology, M.A.S. and A.A.A.; software, M.M.M.M., A.S., S.F.Z. and S.A.-F.; validation, N.H.K.; formal analysis, S.F.Z., A.S. and K.N.; investigation, A.S., K.N. and A.A.A.; resources, S.A.-F.; data curation, A.S., M.A.S. and A.D.C.; writing—original draft, A.S. and A.D.C.; writing—review and editing, M.M.M.M. and A.S.; visualization, N.H.K.; supervision, A.S.; project administration, M.M.M.M.; funding acquisition, M.M.M.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Deanship of Scientific Research (DSR), King Abdulaziz University, Jeddah, Saudi Arabia (NO. IFPRC-108-130-2020).

Data Availability Statement

Data is contained within the article or supplementary material.

Acknowledgments

The authors extend their appreciation to the Deputyship for Research & Innovation, Ministry of Education in Saudi Arabia for funding this research work through project number IFPRC-108-130-2020 and the Deanship of Scientific Research (DSR), King Abdulaziz University, Jeddah, Saudi Arabia.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Plaza, M.G.; Mart, S.; Rubiera, F. CO2 Capture, Use, and Storage in the Cement Industry: State of the Art and Expectations. Energies 2020, 13, 5692. [Google Scholar] [CrossRef]
  2. Kang, S.M.; Adhikari, A.; Bhatta, D.; Gam, H.J.; Gim, M.J.; Son, J.I.; Shin, J.Y.; Lee, I.J. Comparison of Effects of Chemical and Food Waste-Derived Fertilizers on the Growth and Nutrient Content of Lettuce (Lactuca Sativa L.). Resources 2022, 11, 21. [Google Scholar] [CrossRef]
  3. Gao, Y.; Zhuang, B.; Wang, T.; Chen, H.; Li, S.; Wei, W.; Lin, H.; Li, M. Climatic–Environmental Effects of Aerosols and Their Sensitivity to Aerosol Mixing States in East Asia in Winter. Remote Sens. 2022, 14, 3539. [Google Scholar] [CrossRef]
  4. Modarress Fathi, B.; Ansari, A.; Ansari, A. Threats of Internet-of-Thing on Environmental Sustainability by E-Waste. Sustainability 2022, 14, 161. [Google Scholar] [CrossRef]
  5. Valipour, M.; Bateni, S.M.; Jun, C. Global Surface Temperature: A New Insight. Climate 2021, 9, 81. [Google Scholar] [CrossRef]
  6. Siddique, R.; Mejia, A.; Mizukami, N.; Palmer, R.N. Impacts of Globalwarming of 1.5, 2.0 and 3.0 °C on Hydrologic Regimes in the Northeastern U.S. Climate 2021, 9, 9. [Google Scholar] [CrossRef]
  7. Ata, B.; Pakrooh, P.; Barkat, A.; Benhizia, R.; Pénzes, J. Inequalities in Regional Level Domestic CO2 Emissions and Energy Use: A Case Study of Iran. Energies 2022, 15, 3902. [Google Scholar] [CrossRef]
  8. Zhang, C.; Zhang, W.; Luo, W.; Gao, X.; Zhang, B. Analysis of Influencing Factors of Carbon Emissions in China’s Logistics Industry: A Gdim-Based Indicator Decomposition. Energies 2021, 14, 5742. [Google Scholar] [CrossRef]
  9. Pires, L.; Vaz, M.; Carolina, A.; De Oliveira, C.; Falcon, L.; Stella, M.; Pimenta, S.; Bessa, I.G.; Wouters, J.; Andrade, H.S.; et al. Capture and Reuse of Carbon Dioxide (CO2) for a Plastics Circular Economy: A Review. Processes 2021, 9, 759. [Google Scholar] [CrossRef]
  10. Catizzone, E.; Bonura, G.; Migliori, M.; Frusteri, F.; Giordano, G. CO2 Recycling to Dimethyl Ether: State-of-the-Art and Perspectives. Molecules 2018, 23, 31. [Google Scholar] [CrossRef]
  11. Nebel, A.; Cantor, J.; Salim, S.; Salih, A.; Patel, D. The Role of Renewable Energies, Storage and Sector-Coupling Technologies in the German Energy Sector under Different CO2 Emission Restrictions. Sustainability 2022, 14, 10379. [Google Scholar] [CrossRef]
  12. Sarkar, M.; Chung, B. Do Effect of Renewable Energy to Reduce Carbon Emissions under a Flexible Production System: A Step toward Sustainability. Energies 2021, 14, 215. [Google Scholar] [CrossRef]
  13. Materials, M. Supercritical CO2 Curing of Resource-Recycling Secondary Main Materials. Materials 2022, 15, 4581. [Google Scholar]
  14. Hussain, S.; Wang, Y.; Guo, L.; He, T. Theoretical Insights into the Mechanism of Photocatalytic Reduction of CO2 over Semiconductor Catalysts. J. Photochem. Photobiol. C Photochem. Rev. 2022, 52, 100538. [Google Scholar] [CrossRef]
  15. Jerome, M.P.; Alahmad, F.A.; Salem, M.T.; Tahir, M. Layered Double Hydroxide (LDH) Nanomaterials with Engineering Aspects for Photocatalytic CO2conversion to Energy Efficient Fuels: Fundamentals, Recent Advances, and Challenges. J. Environ. Chem. Eng. 2022, 10, 108151. [Google Scholar] [CrossRef]
  16. Sundar, D.; Karuppasamy, L.; Gurusamy, L.; Liu, C.H.; Wu, J.J. Perovskites-like Composites for CO2 Photoreduction into Hydrocarbon Fuels. Curr. Opin. Green Sustain. Chem. 2022, 33, 100563. [Google Scholar] [CrossRef]
  17. Kumar, A.; Singh, P.; Khan, A.A.P.; Van Le, Q.; Nguyen, V.H.; Thakur, S.; Raizada, P. CO2 Photoreduction into Solar Fuels via Vacancy Engineered Bismuth-Based Photocatalysts: Selectivity and Mechanistic Insights. Chem. Eng. J. 2022, 439, 135563. [Google Scholar] [CrossRef]
  18. Zhang, Z.; Wang, Y.; Cui, G.; Liu, H.; Abanades, S.; Lu, H. Improvement of CO2 Photoreduction Efficiency by Process Intensification. Catalysts 2021, 11, 912. [Google Scholar] [CrossRef]
  19. Ali, S.; Flores, M.C.; Razzaq, A.; Sorcar, S.; Hiragond, C.B.; Kim, H.R.; Park, Y.H.; Hwang, Y.; Kim, H.S.; Kim, H.; et al. Gas Phase Photocatalytic CO2 Reduction, “a Brief Overview for Benchmarking”. Catalysts 2019, 9, 727. [Google Scholar] [CrossRef] [Green Version]
  20. Vega-Mendoza, M.S.; Luévano-Hipólito, E.; Torres-Martínez, L.M. Design and Fabrication of Photocatalytic Coatings with α/β-Bi2O3 and Recycled-Fly Ash for Environmental Remediation and Solar Fuel Generation. Ceram. Int. 2021, 47, 26907–26918. [Google Scholar] [CrossRef]
  21. Liu, Y.; Ji, G.; Dastageer, M.A.; Zhu, L.; Wang, J.; Zhang, B.; Chang, X.; Gondal, M.A. Highly-Active Direct Z-Scheme Si/TiO2 Photocatalyst for Boosted CO2 Reduction into Value-Added Methanol. RSC Adv. 2014, 4, 56961–56969. [Google Scholar] [CrossRef]
  22. Low, J.; Cheng, B.; Yu, J. Surface Modification and Enhanced Photocatalytic CO2 Reduction Performance of TiO2: A Review. Appl. Surf. Sci. 2017, 392, 658–686. [Google Scholar] [CrossRef]
  23. Tang, T.; Yin, Z.; Chen, J.; Zhang, S.; Sheng, W.; Wei, W.; Xiao, Y.; Shi, Q.; Cao, S. Novel P-n Heterojunction Bi2O3/Ti3+-TiO2 Photocatalyst Enables the Complete Removal of Tetracyclines under Visible Light. Chem. Eng. J. 2021, 417, 128058. [Google Scholar] [CrossRef]
  24. Zhang, X.; Kim, D.; Yan, J.; Lee, L.Y.S. Photocatalytic CO2 Reduction Enabled by Interfacial S-Scheme Heterojunction between Ultrasmall Copper Phosphosulfide and g-C3N4. ACS Appl. Mater. Interfaces 2021, 13, 9762–9770. [Google Scholar] [CrossRef] [PubMed]
  25. Mohamed, R.M.; Harraz, F.A.; Mkhalid, I.A. Hydrothermal Synthesis of Size-Controllable Yttrium Orthovanadate (YVO4) Nanoparticles and Its Application in Photocatalytic Degradation of Direct Blue Dye. J. Alloys Compd. 2012, 532, 55–60. [Google Scholar] [CrossRef]
  26. Mohamed, R.M.; Mkhalid, I.A. Visible Light Photocatalytic Degradation of Cyanide Using Au-TiO2/Multi-Walled Carbon Nanotube Nanocomposites. J. Ind. Eng. Chem. 2015, 22, 390–395. [Google Scholar] [CrossRef]
  27. Ismail, A.A.; Ibrahim, I.A.; Mohamed, R.M. Sol-Gel Synthesis of Vanadia-Silica for Photocatalytic Degradation of Cyanide. Appl. Catal. B Environ. 2003, 45, 161–166. [Google Scholar] [CrossRef]
  28. Wang, Z.; Cheng, B.; Zhang, L.; Yu, J.; Tan, H. BiOBr/NiO S-Scheme Heterojunction Photocatalyst for CO2 Photoreduction. Sol. RRL 2022, 6, 2100587. [Google Scholar] [CrossRef]
  29. Mokhtar, M.; Shawky, A. Enhanced Visible-Light-Driven H2 Evolution over Sol-Gel Prepared Nd2O3 Supported with PtO Nanoparticles. Ceram. Int. 2022, 48, 36670–36677. [Google Scholar] [CrossRef]
  30. Sharma, A.; Hosseini-bandegharaei, A.; Kumar, N.; Kumar, S. Insight into ZnO/Carbon Hybrid Materials for Photocatalytic Reduction of CO2: An in-Depth Review. J. CO2 Util. 2022, 65, 102205. [Google Scholar] [CrossRef]
  31. Bathla, A.; Lee, J.; Younis, S.A.; Kim, K.H. Recent Advances in Photocatalytic Reduction of CO2 by TiO2– and MOF–Based Nanocomposites Impregnated with Metal Nanoparticles. Mater. Today Chem. 2022, 24, 100870. [Google Scholar] [CrossRef]
  32. Kadi, M.W.; Mohamed, R.M. Promoted Photocatalytic Hydrogen Evolution over Core-Shellstructured CO3O4/TiO2 under Visible Light. Int. J. Mater. Technol. Innov. 2021, 1, 1–13. [Google Scholar]
  33. Allam, N.; El-Shazly, A.; Hegazy, A.; Hamza, M.; El Shenawy, E. The Synergetic Effect of Cobalt Content on Enhancing the Photoelectrochemical Hydrogen Production Performance of In-Situ-Doped TiO2 Photocatalysts. Int. J. Mater. Technol. Innov. 2022, 2, 20–28. [Google Scholar] [CrossRef]
  34. Jiang, Z.; Cheng, B.; Zhang, Y.; Wageh, S.; Al-Ghamdi, A.A.; Yu, J.; Wang, L. S-Scheme ZnO/WO3 Heterojunction Photocatalyst for Efficient H2O2 Production. J. Mater. Sci. Technol. 2022, 124, 193–201. [Google Scholar] [CrossRef]
  35. Li, J.; Zhang, M.; Guan, Z.; Li, Q.; He, C.; Yang, J. Synergistic Effect of Surface and Bulk Single-Electron-Trapped Oxygen Vacancy of TiO2 in the Photocatalytic Reduction of CO2. Appl. Catal. B Environ. 2017, 206, 300–307. [Google Scholar] [CrossRef]
  36. Zou, Z.; Xie, C.; Zhang, S.; Yu, X.; Zou, T.; Li, J. Preparation and Photocatalytic Activity of TiO2/CeO2/Bi2O3 Composite for Rhodamine B Degradation under Visible Light Irradiation. J. Alloys Compd. 2013, 581, 385–391. [Google Scholar] [CrossRef]
  37. Hezam, A.; Namratha, K.; Drmosh, Q.A.; Yamani, Z.H.; Byrappa, K. Synthesis of Heterostructured Bi2O3–CeO2–ZnO Photocatalyst with Enhanced Sunlight Photocatalytic Activity. Ceram. Int. 2017, 43, 5292–5301. [Google Scholar] [CrossRef]
  38. Sabzehmeidani, M.M.; Karimi, H.; Ghaedi, M. Nanofibers Based Quaternary CeO2/CO3O4/Ag/Ag3PO4 S-Scheme Heterojunction Photocatalyst with Enhanced Degradation of Organic Dyes. Mater. Res. Bull. 2022, 147, 111629. [Google Scholar] [CrossRef]
  39. Zhao, Y.; Lin, C.; Zhang, L.; Meng, X.; Tang, H. Bi2O3 Induced Tremella-like Bi2WO6 with Visible Light Catalytic Performance. Ceram. Int. 2022, 48, 4584–4594. [Google Scholar] [CrossRef]
  40. He, W.; Wei, Y.; Xiong, J.; Tang, Z.; Wang, Y.; Wang, X.; Deng, J.; Yu, X.; Zhang, X.; Zhao, Z. Boosting Selective Photocatalytic CO2 Reduction to CO over Dual-Core@shell Structured Bi2O3/Bi2WO6@g-C3N4 Catalysts with Strong Interaction Interface. Sep. Purif. Technol. 2022, 300, 121850. [Google Scholar] [CrossRef]
  41. Hezam, A.; Namratha, K.; Drmosh, Q.A.; Ponnamma, D.; Wang, J.; Prasad, S.; Ahamed, M.; Cheng, C.; Byrappa, K. CeO2 Nanostructures Enriched with Oxygen Vacancies for Photocatalytic CO2 Reduction. ACS Appl. Nano Mater. 2020, 3, 138–148. [Google Scholar] [CrossRef] [Green Version]
  42. Xu, L.; Zeng, J.; Li, Q.; Luo, X.; Chen, T.; Liu, J.; Wang, L.L. Multifunctional Silicene/CeO2 Heterojunctions: Desirable Electronic Material and Promising Water-Splitting Photocatalyst. Chin. Chem. Lett. 2022, 33, 3947–3950. [Google Scholar] [CrossRef]
  43. Cui, Z.; Zhang, D.; Hu, J.; Fang, C. CdS/CeO2 Heterostructures as Visible-Light Photocatalysts for the Reduction of Nitro to Amine Organics. J. Alloys Compd. 2021, 885, 160961. [Google Scholar] [CrossRef]
  44. Wu, D.; Zhang, X.; Liu, S.; Ren, Z.; Xing, Y.; Jin, X.; Ni, G. Fabrication of a Z-Scheme CeO2/Bi2O4 Heterojunction Photocatalyst with Superior Visible-Light Responsive Photocatalytic Performance. J. Alloys Compd. 2022, 909, 164671. [Google Scholar] [CrossRef]
  45. Hsieh, S.H.; Manivel, A.; Lee, G.J.; Wu, J.J. Synthesis of Mesoporous Bi2O3/CeO2 Microsphere for Photocatalytic Degradation of Orange II Dye. Mater. Res. Bull. 2013, 48, 4174–4180. [Google Scholar] [CrossRef]
  46. Cui, S.; Ma, Q.; Chen, H.; Zhang, Y.; Sun, F. Heterojunction G-C3N4/CeO2/Bi2O3 Composite for the Photocatalytic Purification of Exhaust Gas. Mater. Chem. Phys. 2022, 285, 126081. [Google Scholar] [CrossRef]
  47. Nie, J.; Zhu, G.; Zhang, W.; Gao, J.; Zhong, P.; Xie, X.; Huang, Y.; Hojamberdiev, M. Oxygen Vacancy Defects-Boosted Deep Oxidation of NO by β-Bi2O3/CeO2-δ p-n Heterojunction Photocatalyst in Situ Synthesized from Bi/Ce(CO3)(OH) Precursor. Chem. Eng. J. 2021, 424, 130327. [Google Scholar] [CrossRef]
  48. Yang, X.; Zhang, Y.; Wang, Y.; Xin, C.; Zhang, P.; Liu, D.; Mamba, B.B.; Kefeni, K.K.; Kuvarega, A.T.; Gui, J. Hollow β-Bi2O3@CeO2 Heterostructure Microsphere with Controllable Crystal Phase for Efficient Photocatalysis. Chem. Eng. J. 2020, 387, 124100. [Google Scholar] [CrossRef]
  49. Masula, K.; Bhongiri, Y.; Raghav Rao, G.; Vijay Kumar, P.; Pola, S.; Basude, M. Evolution of Photocatalytic Activity of CeO2–Bi2O3 Composite Material for Wastewater Degradation under Visible-Light Irradiation. Opt. Mater. 2022, 126, 112201. [Google Scholar] [CrossRef]
  50. Li, X.; Guan, J.; Jiang, H.; Song, X.; Huo, P.; Wang, H. RGO Modified R-CeO2/g-C3N4 Multi-Interface Contact S-Scheme Photocatalyst for Efficient CO2 Photoreduction. Appl. Surf. Sci. 2021, 563, 150042. [Google Scholar] [CrossRef]
  51. Li, L.; Yan, B. CeO2-Bi2O3 Nanocomposite: Two Step Synthesis, Microstructure and Photocatalytic Activity. J. Non Cryst. Solids 2009, 355, 776–779. [Google Scholar] [CrossRef]
  52. Albukhari, S.M.; Shawky, A. Ag/Ag2O-Decorated Sol-Gel-Processed TeO2 Nanojunctions for Enhanced H2 Production under Visible Light. J. Mol. Liq. 2021, 336, 116870. [Google Scholar] [CrossRef]
  53. Prajapati, P.K.; Malik, A.; Nandal, N.; Pandita, S.; Singh, R.; Bhandari, S.; Saran, S.; Jain, S.L. Morphology Controlled Fe and Ni-Doped CeO2 Nanorods as an Excellent Heterojunction Photocatalyst for CO2 Reduction. Appl. Surf. Sci. 2022, 588, 152912. [Google Scholar] [CrossRef]
  54. Kadi, M.W.; Mohamed, R.M.; Ismail, A.A.; Bahnemann, D.W. Soft and Hard Templates Assisted Synthesis Mesoporous CuO/g-C3N4 Heterostructures for Highly Enhanced and Accelerated Hg(II) Photoreduction under Visible Light. J. Colloid Interface Sci. 2020, 580, 223–233. [Google Scholar] [CrossRef]
  55. Kadi, M.W.; Mohamed, R.M.; Ismail, A.A.; Bahnemann, D.W. Performance of Mesoporous α-Fe2O3/g-C3N4 Heterojunction for Photoreduction of Hg(II) under Visible Light Illumination. Ceram. Int. 2020, 46, 23098–23106. [Google Scholar] [CrossRef]
  56. Kadi, M.W.; Mohamed, R.M.; Ismail, A.A.; Bahnemann, D.W. Decoration of G-C3N4 Nanosheets by Mesoporous CoFe2O4 Nanoparticles for Promoting Visible-Light Photocatalytic Hg(II) Reduction. Colloids Surf. A Physicochem. Eng. Asp. 2020, 603, 125206. [Google Scholar] [CrossRef]
  57. Shawky, A.; Alhaddad, M.; Al-Namshah, K.S.; Mohamed, R.M.; Awwad, N.S. Synthesis of Pt-Decorated CaTiO3 Nanocrystals for Efficient Photoconversion of Nitrobenzene to Aniline under Visible Light. J. Mol. Liq. 2020, 304, 112704. [Google Scholar] [CrossRef]
  58. Mohamed, R.M.; Shawky, A. CNT Supported Mn-Doped ZnO Nanoparticles: Simple Synthesis and Improved Photocatalytic Activity for Degradation of Malachite Green Dye under Visible Light. Appl. Nanosci. 2018, 8, 1179–1188. [Google Scholar] [CrossRef]
  59. Mohamed, R.M.; McKinney, D.; Kadi, M.W.; Mkhalid, I.A.; Sigmund, W. Platinum/Zinc Oxide Nanoparticles: Enhanced Photocatalysts Degrade Malachite Green Dye under Visible Light Conditions. Ceram. Int. 2016, 42, 9375–9381. [Google Scholar] [CrossRef]
  60. Kadi, M.W.; McKinney, D.; Mohamed, R.M.; Mkhalid, I.A.; Sigmund, W. Fluorine Doped Zinc Oxide Nanowires: Enhanced Photocatalysts Degrade Malachite Green Dye under Visible Light Conditions. Ceram. Int. 2016, 42, 4672–4678. [Google Scholar] [CrossRef]
  61. Rajendran, S.; Khan, M.M.; Gracia, F.; Qin, J.; Gupta, V.K.; Arumainathan, S. Ce3+-Ion-Induced Visible-Light Photocatalytic Degradation and Electrochemical Activity of ZnO/CeO2 Nanocomposite. Sci. Rep. 2016, 6, 1–11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Channei, D.; Chansaenpak, K.; Phanichphant, S.; Jannoey, P.; Khanitchaidecha, W.; Nakaruk, A. Synthesis and Characterization of WO3/CeO2 Heterostructured Nanoparticles for Photodegradation of Indigo Carmine Dye. ACS Omega 2021, 6, 19771–19777. [Google Scholar] [CrossRef] [PubMed]
  63. Basaleh, A.S.; Shawky, A.; Zaki, Z.I. Visible Light-Driven Photodegradation of Ciprofloxacin over Sol-Gel Prepared Bi2O3-Modified La-Doped NaTaO3 Nanostructures. Ceram. Int. 2021, 47, 19205–19212. [Google Scholar] [CrossRef]
  64. Shawky, A.; Mohamed, R.M. S-Scheme Heterojunctions: Emerging Designed Photocatalysts toward Green Energy and Environmental Remediation Redox Reactions. J. Environ. Chem. Eng. 2022, 10, 108249. [Google Scholar] [CrossRef]
  65. Zhang, L.; Zhang, J.; Yu, H.; Yu, J. Emerging S-Scheme Photocatalyst. Adv. Mater. 2022, 34, 2107668. [Google Scholar] [CrossRef]
  66. Xu, Q.; Zhang, L.; Cheng, B.; Fan, J.; Yu, J. S-Scheme Heterojunction Photocatalyst. Chem 2020, 6, 1543–1559. [Google Scholar] [CrossRef]
  67. Shawky, A.; Mohamed, R.M.; Alahmadi, N.; Zaki, Z.I. Enhanced Photocatalytic Reduction of Hexavalent Chromium Ions over S-Scheme Based 2D MoS2-Supported TiO2 Heterojunctions under Visible Light. Colloids Surf. A Physicochem. Eng. Asp. 2022, 641, 128564. [Google Scholar] [CrossRef]
  68. Alhaddad, M.; Shawky, A. Pt-Decorated ZnMn2O4 Nanorods for Effective Photocatalytic Reduction of CO2 into Methanol under Visible Light. Ceram. Int. 2021, 47, 9763–9770. [Google Scholar] [CrossRef]
  69. Lais, A.; Gondal, M.A.; Dastageer, M.A. Semiconducting Oxide Photocatalysts for Reduction of CO2 to Methanol. Environ. Chem. Lett. 2018, 16, 183–210. [Google Scholar] [CrossRef]
  70. Pocoví-Martínez, S.; Zumeta-Dube, I.; Diaz, D. Production of Methanol from Aqueous CO2 by Using Co3O4 Nanostructures as Photocatalysts. J. Nanomater. 2019, 2019, 1–10. [Google Scholar] [CrossRef] [Green Version]
  71. Pal, N.; Lee, J.H.; Cho, E.B. Recent Trends in Morphology-Controlled Synthesis and Application of Mesoporous Silica Nanoparticles. Nanomaterials 2020, 10, 2122. [Google Scholar] [CrossRef] [PubMed]
Figure 1. XRD patterns of the produced powders representing the crystallization of pure CeO2 (as labeled in a compared with 5, 10, 15, and 20% Bi2O3/CeO2 as represented by b, c, d, and e, respectively.
Figure 1. XRD patterns of the produced powders representing the crystallization of pure CeO2 (as labeled in a compared with 5, 10, 15, and 20% Bi2O3/CeO2 as represented by b, c, d, and e, respectively.
Catalysts 12 01479 g001
Figure 2. TEM investigations (up) and the corresponding high-resolution imaging for the selected area (down) of pure CeO2 (a,b), 5% Bi2O3/CeO2 (c,d), 10% Bi2O3/CeO2 (e,f), and 15% Bi2O3/CeO2 (g,h).
Figure 2. TEM investigations (up) and the corresponding high-resolution imaging for the selected area (down) of pure CeO2 (a,b), 5% Bi2O3/CeO2 (c,d), 10% Bi2O3/CeO2 (e,f), and 15% Bi2O3/CeO2 (g,h).
Catalysts 12 01479 g002
Figure 3. N2 adsorption/desorption isotherms and corresponding pore size distribution (inset) for prepared CeO2 compared to 15% Bi2O3/CeO2 heterojunctions as indicated.
Figure 3. N2 adsorption/desorption isotherms and corresponding pore size distribution (inset) for prepared CeO2 compared to 15% Bi2O3/CeO2 heterojunctions as indicated.
Catalysts 12 01479 g003
Figure 4. XPS analysis at high resolution for Bi 4f (a), Ce 3d (b), and O 1s (c) of the 15% Bi2O3/CeO2 heterojunction.
Figure 4. XPS analysis at high resolution for Bi 4f (a), Ce 3d (b), and O 1s (c) of the 15% Bi2O3/CeO2 heterojunction.
Catalysts 12 01479 g004
Figure 5. Light absorption spectra of the prepared CeO2 compared with Bi2O3/CeO2 heterojunctions and their corresponding calculated energy gaps as seen in the inset.
Figure 5. Light absorption spectra of the prepared CeO2 compared with Bi2O3/CeO2 heterojunctions and their corresponding calculated energy gaps as seen in the inset.
Catalysts 12 01479 g005
Figure 6. Photocatalytic production of CH3OH utilizing the produced CeO2 and Bi2O3/CeO2 heterojunctions as in (a). The influence of adding Bi2O3 on the generation rate of produced CH3OH is illustrated in (b).
Figure 6. Photocatalytic production of CH3OH utilizing the produced CeO2 and Bi2O3/CeO2 heterojunctions as in (a). The influence of adding Bi2O3 on the generation rate of produced CH3OH is illustrated in (b).
Catalysts 12 01479 g006
Figure 7. PL spectra (a) and transient photocurrents (b) of pure CeO2 and Bi2O3/CeO2 heterojunctions as stated.
Figure 7. PL spectra (a) and transient photocurrents (b) of pure CeO2 and Bi2O3/CeO2 heterojunctions as stated.
Catalysts 12 01479 g007
Figure 8. Photocatalytic production of CH3OH over different doses (in g L−1) of 15% Bi2O3/CeO2 heterojunctions as in (a). The influence of dosage on the generation rate of produced CH3OH is illustrated in (b). The reusability of the best-performing spent photocatalyst dose (2.8 g L−1) is illustrated in (c).
Figure 8. Photocatalytic production of CH3OH over different doses (in g L−1) of 15% Bi2O3/CeO2 heterojunctions as in (a). The influence of dosage on the generation rate of produced CH3OH is illustrated in (b). The reusability of the best-performing spent photocatalyst dose (2.8 g L−1) is illustrated in (c).
Catalysts 12 01479 g008
Figure 9. Estimated band diagrams, charge transfer, and photocatalytic pathway for CO2 conversion into methanol over optimized Bi2O3/CeO2 heterojunctions.
Figure 9. Estimated band diagrams, charge transfer, and photocatalytic pathway for CO2 conversion into methanol over optimized Bi2O3/CeO2 heterojunctions.
Catalysts 12 01479 g009
Table 1. Consequence of Bi2O3 inclusion on physicochemical characteristics of CeO2, namely, surface area (SBET), absorption edge (Abs. edge), and determined bandgap energies (Eg). The consistent photoreduction of CO2 to CH3OH after 9 h irradiation is listed. The corresponding photocurrent intensities are also stated in the last column.
Table 1. Consequence of Bi2O3 inclusion on physicochemical characteristics of CeO2, namely, surface area (SBET), absorption edge (Abs. edge), and determined bandgap energies (Eg). The consistent photoreduction of CO2 to CH3OH after 9 h irradiation is listed. The corresponding photocurrent intensities are also stated in the last column.
SampleSBET
(m2 g−1)
Abs. Edge
(nm)
Eg
(eV)
CH3OH Generation
(μm g−1)
Photocurrent
(µA cm−2)
CeO21914232.9370 (±2.8)11.56
5% Bi2O3/CeO21834272.89360 (±12.6)28.92
10% Bi2O3/CeO21744422.82720 (±21.6)43.52
15% Bi2O3/CeO21694532.751080 (±24.5)51.54
20% Bi2O3/CeO21634542.741100 (±38.4)51.81
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mostafa, M.M.M.; Shawky, A.; Zaman, S.F.; Narasimharao, K.; Abdel Salam, M.; Alshehri, A.A.; Khdary, N.H.; Al-Faifi, S.; Chowdhury, A.D. Visible-Light-Driven CO2 Reduction into Methanol Utilizing Sol-Gel-Prepared CeO2-Coupled Bi2O3 Nanocomposite Heterojunctions. Catalysts 2022, 12, 1479. https://doi.org/10.3390/catal12111479

AMA Style

Mostafa MMM, Shawky A, Zaman SF, Narasimharao K, Abdel Salam M, Alshehri AA, Khdary NH, Al-Faifi S, Chowdhury AD. Visible-Light-Driven CO2 Reduction into Methanol Utilizing Sol-Gel-Prepared CeO2-Coupled Bi2O3 Nanocomposite Heterojunctions. Catalysts. 2022; 12(11):1479. https://doi.org/10.3390/catal12111479

Chicago/Turabian Style

Mostafa, Mohamed Mokhtar Mohamed, Ahmed Shawky, Sharif Fakhruz Zaman, Katabathini Narasimharao, Mohamed Abdel Salam, Abdulmohsen Ali Alshehri, Nezar H. Khdary, Sulaiman Al-Faifi, and Abhishek Dutta Chowdhury. 2022. "Visible-Light-Driven CO2 Reduction into Methanol Utilizing Sol-Gel-Prepared CeO2-Coupled Bi2O3 Nanocomposite Heterojunctions" Catalysts 12, no. 11: 1479. https://doi.org/10.3390/catal12111479

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