DOI:
10.1039/D4TA05778C
(Paper)
J. Mater. Chem. A, 2024,
12, 31847-31860
Tailoring atomically dispersed Fe-induced oxygen vacancies for highly efficient gas-phase photocatalytic CO2 reduction and NO removal with diminished noxious byproducts†
Received
17th August 2024
, Accepted 6th October 2024
First published on 7th October 2024
Abstract
Single-atom-supported metal oxides have attracted extensive interest in energy catalysis, offering a promising avenue for mitigating greenhouse gas emissions and environmental pollution. This study presents a facile synthesis of single-atom Fe-modified Bi2WO6 photocatalysts. By carefully tuning the Fe ratios, the 1.5Fe-Bi2WO6 sample demonstrates exceptional photocatalytic efficiency in CO2 to CO reduction (36.78 μmol g−1). Additionally, an outstanding NO removal performance is also achieved through this photocatalyst with an impressively low conversion of toxic NO2 at just 0.37%. The reaction intermediates and mechanisms governing the photocatalytic reduction of CO2 into CO are elucidated using in situ DRIFTS and in situ XAS techniques. Regarding NO removal, the introduction of Fe single-atoms, along with induced oxygen vacancies, plays a pivotal role in facilitating the transformation of NO and NO2 into nitrate by stabilizing NO and NO2 species. Mechanistic insights into photocatalytic NO oxidation are garnered through scavenger trapping and EPR experiments employing DMPO. This study emphasizes single-atom-supported metal oxide's potential in sustainable chemistry and air purification, providing a promising solution for urgent environmental challenges.
1. Introduction
Carbon dioxide (CO2) and nitrogen oxides (NOx) are harmful atmospheric pollutants. CO2 intensifies global warming and climate change, while NOx contributes to air pollution and health problems.1–3 Photocatalysis has emerged as a promising solution, harnessing sunlight to transform CO2 and diminish NOx concentrations into valuable compounds.4–6 Extensive research efforts have been committed to developing and constructing diverse semiconductor systems for such purposes. Yet, when it comes to the photocatalytic oxidation of NO, it inevitably leads to the production of an intermediate compound, NO2, which is even more hazardous than the initial reactant. As a consequence, if catalytic reactions lead to the predominance of NO2, it could potentially compromise the overall effectiveness of nitric oxide (NO) removal.7 To refrain from such a situation, tailoring active sites is essential in designing the photocatalytic system.
Zhang et al. first proposed the terminology of the heterogeneous single-atom (SA) catalyst (Pt1/FeOx) for CO oxidation.8 This innovative approach involves catalysts composed of isolated or incorporated metal atoms with unique electronic structures and unsaturated coordination.9–12 However, the high surface energy of single atoms will lead to undesired aggregation; hence, the interaction between single atoms and support materials also plays a crucial role in stabilizing single atoms. Among the several types of supported SAs, metal oxide-supported single-atom catalysts have received much attention because of the specific and variable properties observed in metal oxides. The stability of single atoms within the lattice matrix is a key factor, as they tend to maintain thermodynamic stability due to their robust interactions with the surrounding lattice atoms.13,14 Determining the exact anchor location is quite a challenge, especially when metal oxide supports possess fewer surface defects, and thus their interaction with metal species is weak. In this study, we selected bismuth tungstate (Bi2WO6), a member of the Aurivillius oxides family, and an n-type semiconductor as the support material. Bi2WO6 with an orthorhombic structure is constructed by a sandwich substructure of alternating (Bi2O2)n2n+ layers and (WO4)n2n− that are held together by chemical bonds.15,16 Considering that single-atom alloying strategies involve the incorporation of host elements such as Fe, Co, Ni, Cu, Ag, and Au with a small number of individual guest atoms like Ru, Pd, Ir, Pt, and others,17 we have opted to designate transition metal Fe single atoms as the active sites. This choice is attributed to their promising characteristics and cost-effectiveness when compared to noble metals.18 Given the similarity in ionic radius of Fe3+ to that of W6+ and is significantly smaller than that of Bi3+,19 it is highly probable that Fe3+ ions can seamlessly substitute for W6+ ions within the Bi2WO6 lattice. In addition, creating controlled oxygen vacancies (OVs) on metal oxide surfaces represents an effective strategy to enhance the photocatalyst's performance by modifying their band structure, increasing active sites, and lowering activation energy barriers, thus improving reactant adsorption and facilitating challenging chemical reactions.20 It has been verified that the introduction of metal heteroatoms could induce the generation of OVs and sustain their presence during the photocatalytic process.21 This synergistic effect between metal single atoms and OVs stabilizes adsorbed species on the catalyst surface, ultimately resulting in enhanced catalytic performance.
To the best of our knowledge, numerous studies have explored the utilization of Bi2WO6 material in various photocatalytic applications.21–23 However, there is a notable lack of research dedicated to investigating gas-phase photocatalytic CO2 reduction and NO removal over Bi2WO6-based materials using the SAs approach. In this context, we have designed and employed Fe-Bi2WO6 for both applications. The coordination environment and distribution of atomically dispersed Fe were verified by X-ray absorption spectroscopy (XAS) and double spherical aberration-corrected scanning transmission electron microscopy (STEM), respectively. The influence of Fe introduction on the oxygen vacancies' formation was verified by electron paramagnetic resonance (EPR), X-ray photoelectron spectroscopy (XPS), and density functional theory (DFT) calculation results. Eventually, the mechanism of gas-phase photocatalytic CO2 reduction as well as NO oxidation was investigated. In situ XAS and in situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) were employed to study the CO2 reduction mechanism, while scavenger trapping tests and 5,5-dimethyl-1-pyrroline N-oxide (DMPO) spin trapping EPR were conducted for NO oxidation. This allocation of research efforts provides a valuable platform for gaining in-depth insights into the structure–property relationship between single atoms and their associated photocatalytic activity.
2. Experimental section
2.1. Chemicals and materials
Bismuth nitrate pentahydrate (Bi(NO3)3·5H2O), sodium tungstate dihydrate (Na2WO4·2H2O), ethylene glycol (EG), iron(III) nitrate nonahydrate (Fe(NO3)3·9H2O), potassium dichromate (K2Cr2O7), potassium iodide (KI), terephthalic acid, benzoquinone, and DI water. All chemicals were of analytical grade and used without further purification.
2.2. Synthesis procedures of Fe-Bi2WO6
Initially, 1.9403 g of Bi(NO3)3·5H2O and a desired amount of Fe(NO3)3·9H2O were mixed in 50 mL of EG and sonicated for 30 min. Subsequently, 0.6597 g of Na2WO4·2H2O was added to the above solution and stirred further for 30 min. The resulting mixture is then transferred to a 100 mL Teflon-lined stainless-steel autoclave and heated at 190 °C for 2 h. Finally, the precipitate was collected, washed with DI water several times, and dried at 60 °C for 12 h. The obtained samples with the expected molar ratio of Fe3+ to W6+ were set to be 1%, 1.5%, and 2%, denoted as 1Fe/Bi2WO6, 1.5Fe/Bi2WO6, and 2Fe/Bi2WO6, respectively.
For comparison, the pristine Bi2WO6 was prepared without the addition of Fe(NO3)3·9H2O by a similar process.
2.3. Characterizations of materials
The crystal phase of the as-prepared catalysts was examined by X-ray diffraction (XRD) using Cu Kα radiation with a wavelength of 1.5406 Å (Bruker, D2 PHASER with XFalsh). The surface morphologies and microstructures of catalysts were observed by a field-emission scanning electron microscope (FESEM, JEOL, 6700F) and transmission electron microscope (TEM) (JEOL TEM 2100F) equipped with selected area electron diffraction (SAED). High-angle annular dark-field and annular bright-field scanning transmission electron microscopy (HAADF/ABF-STEM) were conducted by JEOL JEM-ARM300F2 equipped with double spherical aberration correctors and energy dispersive spectroscopy (EDS, JED EX-34400MNU). The content of Fe in Fe-Bi2WO6 samples was determined by Agilent 710 inductively coupled plasma optical emission spectroscopy (ICP-OES). The UV-visible diffuse reflectance spectra (UV-vis DRS) were obtained by using a Jasco V-670 spectrometer with an integrated method. Nitrogen adsorption isotherms and CO2 adsorption ability were measured using BELSORB MAX II. XPS was conducted by the Nexsa G2 X-ray Photoelectron Spectrometer System from Thermo Fisher Scientific. UPS measurement was performed using the ULVAC PHI 5000 Versa Probe, and a helium discharge lamp (Specs UVS 10/35) with an energy of 21.2 eV was utilized as a UPS excitation source. X-ray absorption near-edge structure (XANES) measurements of Fe K-edge of as-prepared samples and the samples after stability tests were carried out at the Taiwan Photon Source (TPS)-Beamline of 44A in the National Synchrotron Radiation Research Center (NSRRC, Hsinchu, Taiwan) and collected in fluorescence mode. The in situ XAS of the Fe K-edge was conducted under the conditions of CO2/H2O exposure and light irradiation at the Taiwan Light Source (TLS)-Beamline of 17C1 in the NSRRC. The sample chamber provided an airtight environment during the measurements. The Fe K-edge spectra of 1.5Fe-Bi2WO6 before and after CO2/H2O adsorption and light irradiation were collected by fluorescence mode, and the AM 1.5 solar simulator was used as a light source. The XANES and Fourier transform extended X-ray fine structure (FT-EXAFS) raw data were processed and fitted by Athena and Artemis software. Photoluminescence (PL) and time-resolved photoluminescence (TRPL) were recorded by the Horiba Jobin-Yvon instrument and time-correlated single-photon counting with an excitation wavelength of 405 nm and a frequency of 40 MHz. Electron paramagnetic resonance (EPR) spectroscopy was recorded by a Bruker EMX-Plus EPR spectrometer. The CO temperature-programmed desorption measurements were conducted using AutoChem II 2920 Micrometrics. In situ DRIFTS measurements were performed by a Thermo Nicolet 6700 FTIR spectrometer equipped with a HgCdTe detector and a three-window DRIFTS cell. The isotope labeling experiment was conducted by using 13CO2 to verify the carbon source of the CO product during the photocatalytic CO2 reduction reaction, and the formed 13CO product was monitored by gas chromatography/mass spectrometry (GC-MS, Thermo-Trace 1300 GC + ISQ MS).
2.4. The photocatalytic activity evaluation of catalysts
2.4.1 Photocatalytic CO2 reduction.
The gas-phase photocatalytic CO2 reduction was performed in a closed quartz reactor without the presence of any sacrificial reagents at room temperature (25 ± 5 °C). The AM 1.5 solar simulator was used as a light source and placed perpendicularly above the quartz window of the reactor. According to our previously published research,24 the photocatalytic CO2 reduction experiment was conducted as follows: Initially, 20 milligrams of the catalyst were evenly distributed onto a glass holder, which was then placed inside the reaction chamber and sealed with a quartz window. Subsequently, a high-flow rate of 50 sccm of humidified CO2 was introduced into the chamber and allowed to purge for 15 minutes. Following this, a low-flow rate of 5 sccm of humidified CO2 was introduced into the reactor for an additional 15 minutes before switching on the light source. After 4 hours of continuous light exposure, the products generated during the photocatalytic reaction were analyzed using SRI 8610C gas chromatography (GG) with a glass column Porapak Q equipped with a flame ionization detector (FID) and a helium ionization detector (HID). For the photocatalytic CO2 reduction measurement, the humidity was maintained by flowing CO2 gas through a water bubbler. The stability of photocatalytic CO2 reduction was tested by performing the same protocol on the same sample every 4 hours for a total of 5 cycles.
2.4.2 The photocatalytic NO removal experiment.
The photocatalytic NO removal activity of catalysts was assessed by detecting the NO at the ppb level in a continuous flow reactor at room temperature (25 ± 5 °C), as depicted in Fig. S1.† In detail, 0.2 g of catalyst was ultrasonically dispersed in a certain amount of DI water. The suspension was coated on a Petri dish of 12 cm diameter, dried at 60 °C in an oven, and subsequently placed in a chamber with dimensions of 30 cm × 15 cm × 10 cm (length × width × height). A 300 W Osram light source (230 V) equipped with a 420 nm cut-off filter was positioned vertically, 20 cm above the quartz window. An airstream consisting of 500 ppb NO (humidity level: 70%) flowed through the chamber at a controlled rate of 3 L min−1. The adsorption process was conducted in the dark for 1 h to ensure adsorption/desorption equilibrium before turning on the light source. The concentration of NO was continuously recorded by a chemiluminescence NOx analyzer (Thermal Environmental Instruments Inc., Franklin, MA, model 42c).
The NO removal efficiency (η), NO2 conversion efficiency (ε), NO degradation efficiency into NO3− green product (ρ), and NO3− selectivity (δ) were calculated using eqn (1)–(4), respectively:
| | (1) |
| | (2) |
| | (4) |
where
CNOini and
CiniNO2 represent the concentrations of NO and NO
2 at adsorption–desorption equilibrium in the dark, respectively.
CNOfin and
CfinNO2 represent the final concentrations of NO and NO
2 after turning on the light, respectively.
For the durability test of the catalyst, the catalyst was washed with DI water after every run, dried at 60 °C in a vacuum oven, and used for the next run under the same protocol as the above photocatalytic NO removal measurement.
The trapping measurements were also performed to determine the photocatalytic NO removal mechanism. KI, K2Cr2O7, terephthalic acid, and benzoquinone were utilized as scavengers for trapping holes (h+), electrons (e−), hydroxyl radicals (HO˙), and superoxide radicals (O2˙−), in which 1% of the scavenger was mixed with the catalyst before proceeding with the photocatalytic activity test measurement.
2.5. Computational details
In this study, all DFT calculations were carried out based on the Vienna ab initio simulation package (VASP).25,26 The exchange and correlation potentials were performed through the generalized gradient approximation (GGA)27,28 within the Perdew–Burke–Ernzerhof (PBE)29,30 exchange–correlation functional. The energy cutoff was set to 500 eV for all the calculations. The parameters of the self-consistent field (SCF) force and energy conversion were set to 10−6 eV and 0.02 eV A−1, respectively. The k-point meshes were set to 2 × 5 × 2 for geometry optimization and 4 × 9 × 4 for electronic investigated calculations. An orthorhombic unit cell of Bi2WO6 is built along with the P21 ab space group (No. 29).31,32 The computed lattice parameters of Bi2WO6 as a = 5.437 Å, b = 5.458 Å, and c = 10.872 Å are well agreed with the previous theoretical works33,34 and other experimental studies.31,32 The computed average Bi–O and W–O bong lengths are 2.295 Å and 1.864 Å, respectively. These values are well consistent with the experimental31,32 and the theoretical references,33–35 which demonstrate our reasonable methodology in this study. A 72-atom 2 × 1 × 1 supercell of bulk Bi2WO6 was first relaxed, then the defects were involved in systems with Fe introduction. The DFT + U method36 is applied along with the Hubbard U = 4 for Fe to capture the effects of electronic correlation and Coulomb-driven localization on the structural properties.
The NO oxidation mechanism is determined through the Gibbs free energy (ΔG) valuation computed as follows37
Where ΔZPE and Δ
S are the zero-point energy taken out from the vibrational frequency calculations and the entropy change obtained from the standard thermodynamic tables,
38 respectively;
T is the temperature at 298.15 K; and Δ
E is the computed DFT-reaction energy difference between the reactant and product molecules adsorbed on the surface. Overall, the photocatalytic NO oxidation could be generally described by using the following equations:
| | (6) |
| | (8) |
| | (9) |
where * represents the catalyst's surface, and O*,
, and
denote the adsorbed species on the catalyst's surface, while O
2, NO
2, and NO
3 signify the gaseous molecules.
3. Results and discussion
3.1. Exploration of material properties
The synthesis of Fe-Bi2WO6 photocatalysts with atomic dispersion of Fe is achieved following the procedure depicted in Fig. 1a. The actual amounts of Fe in 1, 1.5, and 2Fe-Bi2WO6 samples were determined to be 0.06, 0.11, and 0.18 wt% by ICP-OES (Table S1†). The crystal structure of the as-prepared sample was examined using X-ray diffraction. Fig. 1b indicates the orthorhombic structure of Bi2WO6 with diffraction peaks at 28.30, 32.91, 47.14, 55.82, 58.54, 68.78, 76.08, and 78.54° corresponding to (131), (200), (202), (133), (262), (400), (102), and (204) crystal planes (PDF 00-039-0256), respectively. The characteristic peaks of iron and iron oxides are not observable in the XRD patterns of Fe-Bi2WO6 samples, which implies the influence of iron introduction into the crystal structure of Bi2WO6 is negligible. The similarity in ionic radii between Fe (63 pm) and W (60 pm), along with the significantly smaller radius of W compared to that of Bi (108 pm), suggests the tendency of Fe ions to replace the W positions within the Bi2WO6 lattice. This observation aligns well with the results obtained from computational calculations (Fig. S2†), and the computed average bond distances of Bi, W, and Fe with O are summarized in Table S2†. Surface morphology and microstructure analyses were conducted using SEM and TEM techniques. As shown in Fig. S3a,† pristine Bi2WO6 has a hierarchical structure with a flower-like shape composed of many thin sheets. For xFe-Bi2WO6 (x = 1, 1.5, 2), no obvious differences could be observed, indicating that the original morphology of Bi2WO6 is unaffected by the Fe modification (Fig. S3b–d†). The mesoporosity of these materials has been confirmed through nitrogen adsorption–desorption measurements (Fig. S4 and Table S3†). All four samples exhibit type IV isotherms with the H3-type hysteresis loop, which is characteristic of mesoporous structures (Fig. S4a†). The BET surface area of 1.5Fe-Bi2WO6 shows a slight increase in comparison with Bi2WO6, which indicates that the introduction of Fe enlarges the surface area of samples (Table S3†). The Barret–Joyner–Halenda (BJH) model was used for pore size analysis (Fig. S4b†), which indicates a broad distribution of mesopores ranging from approximately 2 to 50 nm.
|
| Fig. 1 (a) Synthesis process of Fe-Bi2WO6, (b) XRD patterns of Bi2WO6 and xFe-Bi2WO6 (x = 1, 1.5, 2); HRTEM images of (c) Bi2WO6, and (d) 1.5Fe-Bi2WO6 and inset is respective SAED patterns; (e) HAADF-STEM, (f) ABF-STEM images of 1.5Fe-Bi2WO6; (g) intensity profiles of corresponding D1, D2, D3 points in HAADF-STEM image and B1, B2, B3 points in ABF-STEM image, respectively; (h) STEM images and corresponding EDS mappings of Bi, W, O, and Fe elements for 1.5Fe-Bi2WO6 sample. | |
High-resolution TEM (HRTEM) images of Bi2WO6 and 1.5Fe-Bi2WO6 (Fig. 1c and d) reveal lattice spacings of 0.314 nm and 0.271 nm corresponding to the (131) and (200) crystallographic planes, respectively. Additionally, the SAED patterns, shown in the inset figures within the HRTEM images (Fig. 1c and d), confirm a polycrystalline structure for both Bi2WO6 and 1.5Fe-Bi2WO6. The SAED patterns exhibit indexed reflections corresponding to the (131), (200), (202), and (133) planes of the orthorhombic structure, which are consistent with the XRD results. To visualize the distribution of atomically dispersed Fe, aberration-corrected HAADF/ABF-STEM was carried out and exhibited in Fig. 1e and f. Based on the Z-contrast, the lighter elements in the Z number will be displayed darker and brighter in the Z-contrast images of HAADF- and ABF-STEM, respectively. As seen in Fig. 1e and f, the dark spots in the HAADF-STEM image and the brighter points in the ABF-STEM image of 1.5Fe-Bi2WO6 correspond to isolated Fe atoms. This observation is further confirmed by comparing the relative intensity of atoms at points D1, D2, and D3 in the dark-field image (Fig. 1e) and their corresponding points B1, B2, and B3 in the bright-field image (Fig. 1f), as depicted in Fig. 1g. The STEM-EDS elemental mappings (Fig. 1h) provide additional confirmation of the presence of Bi, W, O, and Fe elements. Furthermore, it clearly illustrates the uniform distribution of Fe throughout the 1.5Fe-Bi2WO6 sample, thereby confirming the successful incorporation of Fe into the crystal structure of Bi2WO6.
The electronic structures and local coordination environment of Fe in Fe-Bi2WO6 samples are corroborated by XANES and EXAFS analysis (Fig. 2a and b). The Fe K-edge rising-edge profiles of Fe-Bi2WO6 samples closely resemble the one of Fe2O3, suggesting the presence of atomically dispersed Fe in a positive valence state, specifically near Fe3+ (Fig. 2a), which aligns well with the findings from EPR results (Fig. 2d). The FT-EXAFS curves (Fig. 2b) of Fe-Bi2WO6 samples have a dominant peak at around 1.45 Å correlated with the Fe–O bond, whereas the peaks related to the Fe–Fe bond in both Fe foil and Fe2O3 are not detectable. Moreover, wavelet transform (WT) EXAFS analysis was conducted to investigate the Fe species. Fig. 2c clearly displays a single intensity peak at approximately 4.9 Å−1 for the Fe-Bi2WO6 samples, corresponding to the presence of the Fe–O from the first nearest-neighbor coordination shell. Notably, the characteristic intensity at 7.7 Å−1 associated with Fe–Fe bonding, as exemplified in the Fe foil, remains unobservable in these Fe-Bi2WO6 samples. This result provides evidence that Fe single atoms are firmly anchored to Bi2WO6 through Fe–O bonding. Besides, the FT-EXAFS fittings (Fig. S5 and S6†) were conducted, and fitting parameters were listed in Table S4,† showing that the coordination numbers of Fe are 4.06, 4.11, and 4.18 for 1, 1.5, and 2Fe-Bi2WO6, respectively. The k-space spectra of Fe-Bi2WO6 samples exhibit substantially different oscillations in comparison with those of Fe foil and Fe2O3 references, magnifying the unconventional coordination of atomically dispersed Fe in Fe-Bi2WO6. The lower coordination numbers of Fe in Fe-Bi2WO6 samples in comparison with the coordination number of 6 for Fe2O3 imply the existence of Fe on the surface instead of a bulk lattice and the formation of oxygen vacancies. The presence of oxygen vacancies and Fe after the modification with Fe is further confirmed by EPR (Fig. 2e). A strong peak at around g = 4.25 in 1.5Fe-Bi2WO6 specifies the presence of Fe3+ species in the sample.39 The enhanced EPR signal at g = 2.007 in 1.5Fe-Bi2WO6, compared to pristine Bi2WO6, confirms a greater presence of oxygen vacancies resulting from the introduction of Fe as a charge compensation mechanism.40 This observation confirms that the introduction of single atoms induces oxygen vacancy formation. Notably, in pristine Bi2WO6, a signal at g = 2.007 is detected, attributed to oxygen vacancies formed during the synthesis process.
|
| Fig. 2 (a) Normalized Fe K-edge XANES spectra, (b) FT-EXAFS spectra, (c) WT-EXAFS spectra of xFe-Bi2WO6 (x = 1, 1.5, 2), Fe2O3 and Fe foil references; (d) EPR spectra, and (e) enlarged spectra providing information about oxygen vacancies of Bi2WO6 and 1.5Fe-Bi2WO6. | |
The chemical compositions and oxidation states of elements in samples were further studied by X-ray photoelectron spectroscopy (Fig. S7 and S8†). The high-resolution Bi 4f spectrum (Fig. S8a†) of Bi2WO6 shows two peaks at 164.42 and 159.09 eV, assigned to Bi 4f5/2 and Bi 4f7/2, respectively.41 The Bi 4f peaks of Fe-Bi2WO6 samples exhibit redshifts because the introduction of Fe changes the electronic density around the Bi element. For W 4f (Fig. S8b†), the two peaks at 37.32 and 35.24 eV correspond to W 4f5/2 and W 4f7/2.42 The O 1s spectrum of Bi2WO6 (Fig. S8c†) could be deconvoluted into three peaks at 530.09, 531.30, and 532.45 eV assigned to lattice oxygen, oxygen vacancies, and adsorbed oxygen, respectively.43 The area of the oxygen vacancy peak increases with higher Fe concentration, confirming the oxygen vacancy formation upon Fe modification. Additionally, redshifts in the binding energy of the W 4f and O 1 s spectra are observed in Fe-Bi2WO6. Moreover, two distinct peaks are observed in the XPS of Fe 2p (Fig. S8d†), corresponding to Fe 2p1/2 and Fe 2p3/2, further verifying the presence of Fe species in Fe-Bi2WO6 samples.44
The photoabsorption properties and band gaps of all samples were determined using UV-vis diffuse reflectance spectroscopy (Fig. 3a). It's observed that there are redshifts in the absorption edges of xFe-Bi2WO6 (x = 1, 1.5, 2). These shifts occur because the modification of Bi2WO6 with Fe creates defect levels near the conduction band, resulting in a broadening of the band tails and an extension of the light absorption range. The optical band gap values of the catalysts were calculated using the Kubelka–Munk equation. Bi2WO6 is known to have an indirect bandgap, and the determined band gap (Eg) values for Bi2WO6, 1Fe-Bi2WO6, 1.5Fe-Bi2WO6, and 2Fe-Bi2WO6 are 2.88 eV, 2.77 eV, 2.74 eV, and 2.69 eV, respectively (Fig. S9†).
|
| Fig. 3 (a) UV-vis DRS, (b) band edge position schematic, (c) PL spectra, (d) TRPL spectra of Bi2WO6, and xFe-Bi2WO6 (x = 1, 1.5, 2). | |
To further understand the change in the band structure of catalysts, the UPS measurement was used, as shown in Fig. S10 and Table S5†. The valence band maximum values of Bi2WO6 and Fe-Bi2WO6 samples are obtained to be 1.58, 1.71, 1.73, and 1.80 vs. NHE, respectively. Combining with the acquired optical band gap values from UV-vis DRS (Fig. S9†), the conduction band minimum values of Bi2WO6 and Fe-Bi2WO6 samples are calculated to be −1.30, −1.06, −1.01, and −0.89 eV, respectively. The band structures of Bi2WO6 and Fe-Bi2WO6 samples are illustrated in Fig. 3b, confirming the suitable band positions of as-prepared catalysts for photocatalytic CO2 reduction reactions. Effective charge separation is crucial for photocatalytic activity. To evaluate this characteristic of the prepared materials, PL and TRPL measurements were conducted. Fig. 3c exhibits the lowest PL emission intensity of 1.5Fe-Bi2WO6 than the ones of pristine Bi2WO6, 1Fe-Bi2WO6, and 2Fe-Bi2WO6, indicating better charge separation after modifying with an appropriate concentration of Fe. Furthermore, the TRPL spectra and TRPL fitting results (Fig. 3d and Table S6†) specify the shorter lifetime of 1.5Fe-Bi2WO6 (1.07 ns) in comparison to Bi2WO6 (1.46 ns), attributed to the suppressed radiative recombination pathway and more efficient charge dissociation and migration in 1.5Fe-Bi2WO6.45 In addition, shorter TRPL lifetimes often indicate the presence of defects or surface states, which can enhance photocatalytic activity by promoting reactant adsorption and interaction with charge carriers. The enhanced charge separation and transfer increase the probability of interaction with oxygen molecules, thereby facilitating the generation of essential reactive oxygen species for the photocatalytic NO removal process.
3.2. Evaluation of photocatalytic reduction and oxidation activity along with mechanistic insights
3.2.1 Photocatalytic CO2 reduction.
The gas-phase photocatalytic CO2 reduction was evaluated through a 4 hours batch-type measurement. As depicted in Fig. 4a, the modification with atomically dispersed Fe significantly enhances the photocatalytic reduction of CO2 to CO. Among the xFe-Bi2WO6 samples, 1.5Fe-Bi2WO6 demonstrates the highest CO production rate at 36.78 μmol g−1, surpassing the pristine Bi2WO6 by 2.7 times. In addition, the CO production rate decreases for 2Fe-Bi2WO6, which is due to inefficient charge separation and transfer, as proven in the PL and TRPL results (Fig. 3c, d and Table S6†). To explain the higher CO2 reduction activity of 1.5Fe-Bi2WO6, the CO2 adsorption ability is one of the important steps. Fig. S11† shows the CO2 adsorption measurement at an adsorption temperature of 298 K. The presence of Fe significantly boosts the adsorption capacity, and notably, 1.5Fe-Bi2WO6 exhibits the highest CO2 uptake capacity among the samples. This improvement suggests that when the Fe single atom is incorporated into the Bi2WO6 surface, it acts as the active site and facilitates electron transfer between the Fe single atom and CO2 molecule on the catalyst surface, ultimately promoting the activation of CO2. Meanwhile, the induced surface oxygen vacancies alter the catalyst's electronic structure and expand the light absorption range, as evidenced in the UV-vis spectra (Fig. 3a). Additionally, as reported in previous studies, these oxygen vacancies reduce the energy barrier for dissociating adsorbed COOH intermediates into CO and OH on the catalyst's surface.16 Efficient product desorption is also a key factor influencing the efficiency and selectivity of the photocatalytic CO2 reduction process. The CO-TPD experiments, illustrated in Fig. 4b, reveal that CO product desorption over 1.5Fe-Bi2WO6 is more effective compared to other catalysts. This improved desorption capability is one of the factors contributing to the superior photocatalytic CO2 reduction performance of 1.5Fe-Bi2WO6. Overall, the exceptional CO2 adsorption capacity, high surface area, efficient light absorption, effective charge separation, and efficient CO desorption all collectively enhance the activity of 1.5Fe-Bi2WO6. Given the significance of CO as a syngas, achieving high selectivity for target products is paramount to reducing post-treatment costs. To gain deeper insights into the catalytically active sites in the photocatalytic CO2 reduction mechanism, we conducted in situ XAS to examine the valence state and charge transfer between active sites and reactants under the working reaction conditions. Fig. 4c presents the in situ XANES Fe K-edge spectra of the 1.5Fe-Bi2WO6 photocatalyst during the reaction test. The non-obvious change in the absorption edge indicates the stable valence state of Fe during the photocatalytic CO2 reduction process. However, there is a decrease in white line intensity after CO2/H2O adsorption, indicating interaction and charge transfer from CO2/H2O to Fe on the catalyst surface, leading to the formation of intermediates. Notably, the white light intensity of the Fe K-edge XANES spectra increases after light irradiation, implying an increase in unoccupied d states of Fe or the transfer of photogenerated electrons from Fe to the reactants.46–48 These in situ XAS results further provide crucial insights into the role of atomically dispersed Fe as the active site in photocatalytic CO2 reduction. The durability of the catalyst for photocatalytic CO2 reduction was assessed and illustrated in Fig. 4d, demonstrating the robustness of 1.5Fe-Bi2WO6 without structural changes even after undergoing five cycles of CO2 photoreduction, as confirmed by XAS measurements (Fig. S12 and Table S7†). We also conducted a comparison of gas-phase CO2 reduction efficiency among Bi2WO6-based catalysts (as summarized in Table S8†), revealing that Fe single atoms-modified Bi2WO6 stands out as a promising and efficient photocatalyst for this reaction. To validate the obtained hydrocarbon products and eliminate the possibility of carbon contamination, a control photocatalytic experiment was carried out under inert gas and visible-light irradiation. No CO was detected in the blank experiment. Furthermore, to confirm the origin of CO, an isotope-tracer experiment was performed using 13CO2 under the same photocatalytic reaction conditions (Fig. S13†).
|
| Fig. 4 (a) Photocatalytic CO2 reduction activity, (b) CO-TPD curves of Bi2WO6 and xFe-Bi2WO6 (x = 1, 1.5, 2) samples, respectively; (c) Fe K-edge XANES spectra of 1.5Fe-Bi2WO6 before/after CO2/H2O adsorption and under light irradiation; (d) photocatalytic CO2 reduction stability test over 1.5Fe-Bi2WO6. | |
The adsorption behaviors and formation of intermediates during photocatalytic CO2 reduction were further investigated by in situ DRIFTS under dark and light conditions. The background spectra were collected before purging CO2/H2O and light irradiation. Fig. 5a and b illustrates the adsorption of CO2 and H2O vapor on the surface of 1.5Fe-Bi2WO6. The peaks around 1654 cm−1 and 1632 cm−1 are apparently observable after purging CO2 and H2O vapor without light illumination, assigned to the surface chemisorbed CO2 and bending mode of adsorbed water (δ-H2O), respectively.1,49 The others at about 1384 cm−1 and 1185 cm−1 are attributed to monodentate carbonate (m-CO32−) and bicarbonate (HCO3−) species.50 After light irradiation (Fig. 5c and d), there is a clear reduction in the adsorbed species, suggesting the consumption of absorbed intermediates during the photoreduction reaction. The evolution of a new peak around 1290 cm−1 related to monodentate carbonate (m-CO32−) species.51 Furthermore, the peak at 1533 cm−1 gradually evolved, corresponding to COOH species, which is a crucial intermediate for the formation of CO product.52,53 As reported in previous studies, the formation of COOH intermediate is the rate-determining step during the photocatalytic CO2 reduction reaction.54,55 The possible photocatalytic CO2 reduction mechanism is proposed and illustrated in Fig. 5e. First, the photogenerated carriers will be separated and migrated into the catalyst's active sites upon light irradiation. Subsequently, the CO2 and H2O molecules adsorbed on the catalyst's surface become activated, forming m-CO32− and HCO3− intermediates. These intermediates are then further reduced to CO through the formation and dissociation of the COOH intermediate in the presence of H+ and e− under light irradiation. Finally, the generated CO is released from the catalyst's surface. The following mechanism is proposed for CO2 reduction into CO over 1.5Fe-Bi2WO6:
| 1.5Fe-Bi2WO6 + hν → 1.5Fe-Bi2WO6 (e− + h+) | (10) |
| | (11) |
| | (12) |
| | (13) |
| COOH* + H+ + e− → CO* + H2O | (14) |
|
| Fig. 5
In situ DRIFTS spectra for photocatalytic CO2 reduction over 1.5Fe-Bi2WO6 recorded (a and b) in the dark for 1 hour and (c and d) under light irradiation conditions for 3 hours; (e) proposed photocatalytic CO2 to CO reduction pathway. | |
3.2.2 Photocatalytic NO removal.
Photocatalytic NO removal performance was evaluated in a flow reaction system under visible light illumination. Upon reaching the adsorption–desorption equilibrium of NO on the catalyst surface, the lamp was switched on to initiate the reaction. The outcomes of the photocatalytic NO removal process are presented in Fig. 6. It's worth noting that a blank test conducted without catalysts highlights the remarkable photostability of NO, as demonstrated in Fig. 6a. In the case of Bi2WO6, only 36.28% of NO is removed after 30 minutes of light irradiation (Fig. 6a), while it exhibits a substantial conversion efficiency of 17.34% for the harmful NO2 byproduct (Fig. 6b and c). Given the undesirable formation of a toxic NO2 byproduct, it is crucial to mitigate its formation during photocatalytic NO oxidation. As illustrated in Fig. 6a, the introduction of atomically dispersed Fe led to the enhancement of the NO removal efficiency, achieving 42.61 and 50.37% for 1Fe- and 1.5Fe-Bi2WO6, respectively. The NO2 conversion performances are significantly inhibited in the presence of Fe, resulting in rates of 1.75% and 0.37% for 1Fe- and 1.5Fe-Bi2WO6, respectively, while the selectivity towards the NO3− product increases from 82.66% (Bi2WO6) to 99.63% (1.5Fe-Bi2WO6) (Fig. 6c). However, when the Fe dopant amount is further increased to 2Fe-Bi2WO6, there is a decrease in NO removal efficiency and an increase in NO2 conversion. This can be attributed to the fact that a higher concentration of Fe dopants can act as recombination reservoirs for photogenerated electrons and holes, as indicated by the PL and TRPL results (Fig. 3c and d).56 Hence, the availability of charge carriers needed for the generation of reactive species essential for the oxidation of NO and NO2 to form NO3− product is reduced. The accumulation of adsorbed reactants and intermediates onto the catalyst surface without its efficient conversion to final products can result in surface saturation and poisoning, further hindering the exposure of active sites. Similar catalyst behavior linked to surface passivation has been reported in recent studies.57–59 Moreover, the higher Fe introduction into Bi2WO6 causes a gradual downward shift in the conduction band, rendering it thermodynamically unfavorable for reduction reactions, particularly the generation of superoxide radicals. In addition, as shown in UPS results (Fig. S10†), the work function decreases upon the introduction of Fe, indicating improved charge transfer, which could be one of the factors that affect photocatalytic activity. For a more in-depth understanding of the photocatalytic NO removal rate, we have fitted the kinetic curves, which are displayed in Fig. 6d. The calculated rate constants are 11.04 × 10−2 min−1 for 1.5Fe-Bi2WO6, which is 1.5, 1.2, and 1.1 times higher than Bi2WO6 (7.53 × 10−2 min−1), 1Fe-Bi2WO6 (9.14 × 10−2 min−1), and 2Fe-Bi2WO6 Bi2WO6 (9.97 × 10−2 min−1), respectively.
|
| Fig. 6 (a) Photocatalytic NO removal efficiency, (b) evolution of NO2, (c) NO2, NO3− conversion efficiency and NO3− selectivity, (d) the kinetic constants of photocatalytic NO removal reaction for Bi2WO6 and xFe-Bi2WO6 (x = 1, 1.5, and 2) under visible light; (e) stability test and (f) the selectivity toward NO3− of 1.5Fe-Bi2WO6 for photocatalytic NO removal. | |
The long-term stability of catalysts is another critical factor to consider in practical applications of the photocatalytic NO removal process. As demonstrated in Fig. 6e, 1.5Fe-Bi2WO6 exhibits remarkable stability, with only a 4.34% decrease in NO removal efficiency after 5 runs. Moreover, the selectivity of NO3 only vanishes from 99.63% to 97.94%, and the NO2 conversion increases from 0.37% to 2.06%, respectively (Fig. 6f). Furthermore, the structural analysis of Fe single atoms after photocatalytic NO removal was also examined using XAS. The XANES spectrum (Fig. S14a†) of 1.5Fe-Bi2WO6 revealed no alternations in the oxidation states of Fe species. The coordination environment around Fe single atoms was verified through EXAFS fitting (Fig. S14b–d and Table S9†), confirming their existence in the single atom form after the photocatalytic NO removal stability test. The coordination number remained lower than that of the Fe2O3 reference, indicating the durability of induced oxygen vacancies. These results highlight that 1.5Fe-Bi2WO6 exhibits not only excellent stability but also a preference for green product selectivity. In comparison with previously reported studies, as summarized in Table S10†, the atomically dispersed 1.5Fe-Bi2WO6 stands out as a potential and promising candidate for effective photocatalytic NO removal, characterized by low NO2 conversion and good stability.
To gain further insight into the photocatalytic NO removal mechanism, trapping tests with various scavengers and spin-trapping EPR were conducted, as shown in Fig. 7a–d. K2Cr2O7, KI, benzoquinone, and terephthalic acid were utilized to trap correspondingly the photogenerated electrons, holes, superoxide (O2˙−), and hydroxyl radicals (OH˙). As shown in Fig. 7a, the NO removal performance vanishes significantly with the addition of K2Cr2O7, KI, and benzoquinone, which implies the important roles of photogenerated electrons and holes in the photocatalytic NO removal process. The rate constant values of the NO removal reaction in the absence and presence of terephthalic acid, benzoquinone, KI, and K2Cr2O7 were calculated to be 11.04, 10.51, 8.93, 5.94, and 0.34 × 10−2 min−1, respectively. Furthermore, the presence of hole scavengers significantly enhances NO2 conversion, highlighting the crucial role of holes in the deep oxidation process of NO2 into NO3− (Fig. 7b). EPR measurements were conducted to further confirm the formation of reactive radicals during the photocatalytic NO removal reaction. As illustrated in Fig. 7c and d, no signals of O2˙− and OH˙ radicals were detected under the dark condition, suggesting that light irradiation is crucial for the formation of reactive free radicals. Under light illumination, the signals of DMPO-O2˙− (intensity of 1:1:1:1) and DMPO-OH˙ (intensity of 1:2:2:1) adducts were observable for both Bi2WO6 and 1.5 Fe-Bi2WO6, indicating the generation of these radicals during the photocatalytic process and being consistent with the trapping tests. In addition, 1.5Fe-Bi2WO6 shows a higher intensity in the DMPO-O2˙− and DMPO-OH˙ spectra demonstrating that more reactive radicals are generated after Fe modification.
|
| Fig. 7 (a) Trapping tests and the right panel represent the kinetic rate of photocatalytic NO oxidation over 1.5Fe-Bi2WO6 with/without scavengers; (b) NO2 conversion during photocatalytic NO removal over 1.5Fe-Bi2WO6 with/without scavengers; (c and d) DMPO spin-trapping EPR spectra of Bi2WO6 and 1.5Fe-Bi2WO6 for O2−˙ and HO˙ radicals in the aqueous dispersion, respectively; (e) schematic illustration of photocatalytic NO removal over 1.5Fe-Bi2WO6; (f) Gibbs free energy for the formation of intermediates during photocatalytic NO removal over Bi2WO6 and Fe-Bi2WO6. | |
Based on the obtained band structure, trapping tests, and EPR results, the mechanism for the photocatalytic NO removal process was proposed, following eqn (16)–(27) and illustrated in Fig. 7e. Firstly, Fe-Bi2WO6 absorbs the incident photons and generates electron–hole pairs (eqn (16)). These pairs are then separated, migrate to the surface of the catalyst, and participate in the redox reactions. Notably, the conduction band potential is more negative than the redox potential of O2/O2˙−, enabling photoinduced electrons to participate in the reduction of adsorbed oxygens, yielding O2˙− radicals (eqn (17)). These active O2˙− species will subsequently attend the generation of active OH˙ radicals through the reduction of H2O2 (eqn (18) and (19). This explains the detection of OH˙ radicals in EPR spectra, although the valence band potential of 1.5Fe-Bi2WO6 is less positive than the redox potential of H2O/OH˙ (1.99 V). Subsequently, these reactive radicals (O2˙−, OH˙) oxidize the adsorbed NO to generate NO3− (eqn (20) and (21)). In addition, NO could be oxidized to form NO2 through eqn (22) and (23) and the desorption of NO2 from the catalyst's surface induces the rising production of the detrimental by-product. Therefore, it becomes essential to stabilize NO2 on the catalyst's surface to further transform it into NO3− (eqn (24)–(27)). In this process, Fe atoms serve as active sites for adsorption and reduction of NO2 to NO2−. Previous studies have shown that Fe3+ species could act as active sites for the adsorption and stabilization of NO2.18,60,61 Moreover, more generated oxygen vacancies could boost the surface adsorption and activation of O2 molecules into active O2˙− and OH˙ species.62 Therefore, the synergistic effect between atomically dispersed Fe and oxygen vacancies is advantageous for the subsequent deep oxidation of stabilized NO2 by photogenerated holes, O2˙− and OH˙ active species to produce NO3− and vanish the evolution of noxious NO2 byproduct.63 The DFT calculations of NO, NO2, and NO3 adsorption energy over pristine Bi2WO6 and Fe-Bi2WO6 were conducted and shown in Fig. S15.† The adsorption energy of NO and NO2 on Fe-Bi2WO6 is higher than the pristine one, suggesting that NO and NO2 are strongly stabilized on the surface of Fe-Bi2WO6, facilitating subsequent deep oxidation processes. Additionally, Fig. 7f displays Gibbs free energy diagrams for the conversion of NO to NO3. The formation energies of intermediates over Fe-Bi2WO6 are lower than the ones of pristine Bi2WO6, indicating the modification of Bi2WO6 with Fe decreases the formation energy barrier and is favorable for the formation of NO3−.
| 1.5Fe-Bi2WO6 + hν → e− + h+ | (16) |
| O2˙− + e− + 2H+ → H2O2 | (18) |
| H2O2 + e− → HO˙ + HO− | (19) |
| NO + 2HO˙ + HO− → NO3− + H2O | (21) |
| | (22) |
| NO + 2HO˙ → NO2 + H2O | (23) |
| NO2 + 2HO− + h+ → NO3− + H2O | (24) |
| NO2 + HO˙ → NO3− + H+ | (25) |
| 2NO2− + O2˙− + h+ → 2NO3− | (27) |
4. Conclusions
In summary, we have demonstrated an efficient strategy for engineering single atoms of the low-cost transition metal (Fe)-supported metal oxide Bi2WO6. This strategy has yielded remarkable improvements in the photocatalytic performance, particularly the selective reduction of CO2 and oxidation of NO. The introduction of Fe and induced oxygen vacancies provides a synergistic effect that stabilizes NO/NO2 and generates more reactive radicals, resulting in a substantial reduction in the production of harmful NO2 byproducts. Our comprehensive experimental and theoretical investigations have unveiled the pivotal roles of Fe SAs and oxygen vacancies in inhibiting NO2 formation, ultimately achieving a remarkable 99.63% green product selectivity. Additionally, the atomically dispersed Fe-Bi2WO6 photocatalyst exhibits the highest gas-phase photocatalytic CO2 to CO reduction activity (36.78 μmol g−1) compared to other bismuth tungstate-based photocatalysts. This study not only offers a straightforward strategy for designing functional photocatalysts for efficient solar energy conversion but also opens new avenues for engineering non-precious active centers in metal oxide catalysts for various applications in heterogeneous catalysis.
Data availability
The data supporting this article have been included as part of the ESI.†
Author contributions
Nguyen Quoc Thang: conceptualization, methodology, investigation, formal analysis, writing – original draft. Amr Sabbah: conceptualization, methodology, investigation, validation, writing – review and editing. Chih-Yang Huang: methodology, investigation. Nguyen Hoang Phuong: methodology, investigation, Tsai-Yu Lin: methodology, investigation. Mahmoud Kamal Hussien: methodology, investigation, Heng-Liang Wu: methodology, investigation, resources. Chih-I Wu: methodology, investigation, resources. Nguyet. N. T. Pham: methodology, investigation. Pham Van Viet: investigation, resources. Chih-Hao Lee: investigation, resources. Li-Chyong Chen: conceptualization, validation, writing – review and editing, resources, supervision, project administration, funding acquisition. Kuei-Hsien Chen: conceptualization, validation, writing – review and editing, resources, supervision, project administration, funding acquisition.
Conflicts of interest
The authors declare that they have no competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.
Acknowledgements
The authors acknowledge the technical support provided by Mr Tzu-Hsien Tseng at the Institutional Center of National Chung Hsing University and Ms. Payal Maharathi at National Taiwan University. This work was financially supported by the National Science and Technology Council (NSTC) of Taiwan, Science Vanguard Project (NSTC 111-2123-M-002-009), Academic Summit Project (NSTC 112-2639-M-002-005-ASP), and NSTC 112-2113-M-001-023. Additional financial supports from the Center of Atomic Initiative for New Materials, National Taiwan University, from the Featured Areas Research Center Program within the framework of the Higher Education Sprout Project by the Ministry of Education in Taiwan (111L9008 and 112L9008), and Academia Sinica (AS-SS-112-01) are also acknowledged.
References
- A. Sabbah, I. Shown, M. Qorbani, F.-Y. Fu, T.-Y. Lin, H.-L. Wu, P.-W. Chung, C.-I. Wu, S. R. M. Santiago, J.-L. Shen, K.-H. Chen and L.-C. Chen, Nano Energy, 2022, 93, 106809 CrossRef CAS.
- T. Billo, I. Shown, A. k. Anbalagan, T. A. Effendi, A. Sabbah, F.-Y. Fu, C.-M. Chu, W.-Y. Woon, R.-S. Chen, C.-H. Lee, K.-H. Chen and L.-C. Chen, Nano Energy, 2020, 72, 104717 CrossRef CAS.
- R. Chen, J. Li, H. Wang, P. Chen, X. a. Dong, Y. Sun, Y. Zhou and F. Dong, J. Mater. Chem. A, 2021, 9, 20184–20210 RSC.
- I. Shown, S. Samireddi, Y. C. Chang, R. Putikam, P. H. Chang, A. Sabbah, F. Y. Fu, W. F. Chen, C. I. Wu, T. Y. Yu, P. W. Chung, M. C. Lin, L. C. Chen and K. H. Chen, Nat. Commun., 2018, 9, 169 CrossRef PubMed.
- M. Qorbani, A. Sabbah, Y. R. Lai, S. Kholimatussadiah, S. Quadir, C. Y. Huang, I. Shown, Y. F. Huang, M. Hayashi, K. H. Chen and L. C. Chen, Nat. Commun., 2022, 13, 1256 CrossRef CAS PubMed.
- C. Zhang, Y. Xu, H. Bai, D. Li, L. Wei, C. Feng, Y. Huang, Z. Wang, X. Li, X. Cui, C. Hu and F. Wang, Nano Energy, 2024, 121, 109197 CrossRef CAS.
- C. Wu, Q. Tang, S. Zhang, K. Lv, X. Fuku and J. Wang, ACS Appl. Mater. Interfaces, 2023, 15, 30127–30138 CrossRef CAS PubMed.
- B. Qiao, A. Wang, X. Yang, L. F. Allard, Z. Jiang, Y. Cui, J. Liu, J. Li and T. Zhang, Nat. Chem., 2011, 3, 634–641 CrossRef CAS PubMed.
- B. Li, C. Guo, X. Wang, W. Dong, B. Xu, X. Xing, D. Zhou, X. Xue, Q. Luan, W. Tang and C. Hou, Mater. Today Nano, 2023, 21, 100281 CrossRef CAS.
- G. Ren, M. Shi, S. Liu, Z. Li, Z. Zhang and X. Meng, Chem. Eng. J., 2023, 454, 140158 CrossRef CAS.
- Y. Shi, C. Zhao, H. Wei, J. Guo, S. Liang, A. Wang, T. Zhang, J. Liu and T. Ma, Adv. Mater., 2014, 26, 8147–8153 CrossRef CAS PubMed.
- Q. Yang, Y. Jiang, H. Zhuo, E. M. Mitchell and Q. Yu, Nano Energy, 2023, 111, 108404 CrossRef CAS.
- X. Xiong, C. Mao, Z. Yang, Q. Zhang, G. I. N. Waterhouse, L. Gu and T. Zhang, Adv. Energy Mater., 2020, 10, 2002928 CrossRef CAS.
- C. Gao, J. Low, R. Long, T. Kong, J. Zhu and Y. Xiong, Chem. Rev., 2020, 120, 12175–12216 CrossRef CAS PubMed.
- P. Yang, J. Zhang, C. Chen, X. Guo, J. Zhang, X. Zhang and Z. Wu, ACS Appl. Nano Mater., 2022, 5, 5128–5139 CrossRef CAS.
- S. Xiong, S. Bao, W. Wang, J. Hao, Y. Mao, P. Liu, Y. Huang, Z. Duan, Y. Lv and D. Ouyang, Appl. Catal., B, 2022, 305, 121026 CrossRef CAS.
- Y. Wang, Y. Zhang, W. Yu, F. Chen, T. Ma and H. Huang, J. Mater. Chem. A, 2023, 11, 2568–2594 RSC.
- G. Cheng, X. Liu, X. Song, X. Chen, W. Dai, R. Yuan and X. Fu, Appl. Catal., B, 2020, 277, 119196 CrossRef CAS.
- L. Li, J. Yang, L. Yang, F. Fu, H. Xu and X. Fan, J. Environ. Chem. Eng., 2022, 10, 107708 CrossRef CAS.
- M. Xiao, L. Zhang, B. Luo, M. Lyu, Z. Wang, H. Huang, S. Wang, A. Du and L. Wang, Angew Chem. Int. Ed. Engl., 2020, 59, 7230–7234 CrossRef CAS PubMed.
- L. Liu, J. Liu, K. Sun, J. Wan, F. Fu and J. Fan, Chem. Eng. J., 2021, 411, 128629 CrossRef CAS.
- Y. Qiu, J. Lu, Y. Yan and J. Niu, J. Hazard. Mater., 2022, 422, 126920 CrossRef CAS PubMed.
- T. Hu, H. Li, N. Du and W. Hou, ChemCatChem, 2018, 10, 3040–3048 CrossRef CAS.
- N. Q. Thang, A. Sabbah, L.-C. Chen, K.-H. Chen, L. V. Hai, C. M. Thi and P. V. Viet, Chem. Eng. Sci., 2021, 229, 116049 CrossRef CAS.
- G. Kresse and J. Furthmüller, Comput. Mater. Sci., 1996, 6, 15–50 CrossRef CAS.
- G. Kresse and J. Furthmuller, Phys. Rev. B: Condens. Matter Mater. Phys., 1996, 54, 11169–11186 CrossRef CAS PubMed.
- J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865–3868 CrossRef CAS PubMed.
- J. P. Perdew, J. A. Chevary, S. H. Vosko, K. A. Jackson, M. R. Pederson, D. J. Singh and C. Fiolhais, Phys. Rev. B: Condens. Matter Mater. Phys., 1992, 46, 6671–6687 CrossRef CAS PubMed.
- B. Hammer, L. B. Hansen and J. K. Nørskov, Phys. Rev. B: Condens. Matter Mater. Phys., 1999, 59, 7413–7421 CrossRef.
- M. Ernzerhof and G. E. Scuseria, J. Chem. Phys., 1999, 110, 5029–5036 CrossRef CAS.
- J. G. Thompson, A. D. Rae, R. L. Withers and D. C. Craig, Acta Crystallogr., Sect. B: Struct. Sci., 1991, 47, 174–180 CrossRef.
- N. A. McDowell, K. S. Knight and P. Lightfoot, Chemistry, 2006, 12, 1493–1499 CrossRef CAS PubMed.
- H. Djani, P. Hermet and P. Ghosez, J. Phys. Chem. C, 2014, 118, 13514–13524 CrossRef CAS.
- V. Koteski, J. Belošević-Čavor, V. Ivanovski, A. Umićević and D. Toprek, Appl. Surf. Sci., 2020, 515, 146036 CrossRef CAS.
- Y. Wei, X. Wei, S. Guo, Y. Huang, G. Zhu and J. Zhang, J. Mater. Sci. Eng. B, 2016, 206, 79–84 CrossRef CAS.
- B. Himmetoglu, A. Floris, S. de Gironcoli and M. Cococcioni, Int. J. Quantum Chem., 2014, 114, 14–49 CrossRef CAS.
- J. K. Nørskov, J. Rossmeisl, A. Logadottir, L. Lindqvist, J. R. Kitchin, T. Bligaard and H. Jónsson, J. Phys. Chem. B, 2004, 108, 17886–17892 CrossRef.
-
W. M. Haynes, CRC Handbook of Chemistry and Physics, CRC press, 2016 Search PubMed.
- J. Ma, H. He and F. Liu, Appl. Catal., B, 2015, 179, 21–28 CrossRef CAS.
- Q. Li, X. Zhu, J. Yang, Q. Yu, X. Zhu, J. Chu, Y. Du, C. Wang, Y. Hua, H. Li and H. Xu, Inorg. Chem. Front., 2020, 7, 597–602 RSC.
- Y. Liu, D. Shen, Q. Zhang, Y. Lin and F. Peng, Appl. Catal., B, 2021, 283, 119630 CrossRef CAS.
- Y. Jiang, H. Y. Chen, J. Y. Li, J. F. Liao, H. H. Zhang, X. D. Wang and D. B. Kuang, Adv. Funct. Mater., 2020, 30, 2004293 CrossRef CAS.
- H. Ma, W. Yang, S. Gao, W. Geng, Y. Lu, C. Zhou, J. K. Shang, T. Shi and Q. Li, Chem. Eng. J., 2023, 455, 140471 CrossRef CAS.
- Y. Xiong, H. Li, C. Liu, L. Zheng, C. Liu, J. O. Wang, S. Liu, Y. Han, L. Gu, J. Qian and D. Wang, Adv. Mater., 2022, 34, e2110653 CrossRef PubMed.
- G. Zhao, W. Li, H. Zhang, W. Wang and Y. Ren, Chem. Eng. J., 2022, 430, 132937 CrossRef CAS.
- M. Zhou, Z. Wang, A. Mei, Z. Yang, W. Chen, S. Ou, S. Wang, K. Chen, P. Reiss, K. Qi, J. Ma and Y. Liu, Nat. Commun., 2023, 14, 2473 CrossRef CAS PubMed.
- J. Fan, Y. Zhao, H. Du, L. Zheng, M. Gao, D. Li and J. Feng, ACS Appl. Mater. Interfaces, 2022, 14, 26752–26765 CrossRef CAS PubMed.
- S. Ji, Y. Qu, T. Wang, Y. Chen, G. Wang, X. Li, J. Dong, Q. Chen, W. Zhang, Z. Zhang, S. Liang, R. Yu, Y. Wang, D. Wang and Y. Li, Angew Chem. Int. Ed. Engl., 2020, 59, 10651–10657 CrossRef CAS PubMed.
- H. Huang, X. Liu, F. Li, Q. He, H. Ji and C. Yu, Sustainable Energy Fuels, 2022, 6, 4903–4915 RSC.
- M. Wang, M. Shen, X. Jin, J. Tian, Y. Zhou, Y. Shao, L. Zhang, Y. Li and J. Shi, Nanoscale, 2020, 12, 12374–12382 RSC.
- Y. Bo, P. Du, H. Li, R. Liu, C. Wang, H. Liu, D. Liu, T. Kong, Z. Lu, C. Gao and Y. Xiong, Appl. Catal., B, 2023, 330, 122667 CrossRef CAS.
- Z. Jiang, H. Sun, T. Wang, B. Wang, W. Wei, H. Li, S. Yuan, T. An, H. Zhao, J. Yu and P. K. Wong, Energy Environ. Sci., 2018, 11, 2382–2389 RSC.
- M. K. Hussien, A. Sabbah, M. Qorbani, R. Putikam, S. Kholimatussadiah, D. M. Tzou, M. H. Elsayed, Y. J. Lu, Y. Y. Wang, X. H. Lee, T. Y. Lin, N. Q. Thang, H. L. Wu, S. C. Haw, K. C. Wu, M. C. Lin, K. H. Chen and L. C. Chen, Small, 2024, 20, 2400724 CrossRef CAS PubMed.
- M. Kamal Hussien, A. Sabbah, M. Qorbani, M. Hammad Elsayed, P. Raghunath, T.-Y. Lin, S. Quadir, H.-Y. Wang, H.-L. Wu, D.-L. M. Tzou, M.-C. Lin, P.-W. Chung, H.-H. Chou, L.-C. Chen and K.-H. Chen, Chem. Eng. J., 2022, 430, 132853 CrossRef CAS.
- X. Jiao, X. Li, X. Jin, Y. Sun, J. Xu, L. Liang, H. Ju, J. Zhu, Y. Pan, W. Yan, Y. Lin and Y. Xie, J. Am. Chem. Soc., 2017, 139, 18044–18051 CrossRef CAS PubMed.
- L. Wen, B. Liu, X. Zhao, K. Nakata, T. Murakami and A. Fujishima, Int. J. Photoenergy, 2012, 2012, 1–10 Search PubMed.
- X. Wu, J. Zhou, Q. Tan, K. Li, Q. Li, S. A. Correia Carabineiro and K. Lv, ACS Appl. Mater. Interfaces, 2024, 16, 11479–11488 CrossRef CAS PubMed.
- R. Hailili, Z. Li, X. Lu, H. Sheng, D. W. Bahnemann and J. Zhao, Environ. Sci.: Nano, 2024, 11, 3301–3316 RSC.
- T. Shen, X. Shi, J. Guo, J. Li and S. Yuan, Chem. Eng. J., 2021, 408, 128014 CrossRef CAS.
- X. Song, W. Jiang, Z. Cai, X. Yue, X. Chen, W. Dai and X. Fu, Chem. Eng. J., 2022, 444, 136709 CrossRef CAS.
- Q. Wu and R. van de Krol, J. Am. Chem. Soc., 2012, 134, 9369–9375 CrossRef CAS PubMed.
- Y. Xin, Q. Zhu, T. Gao, X. Li, W. Zhang, H. Wang, D. Ji, Y. Huang, M. Padervand, F. Yu and C. Wang, Appl. Catal., B, 2023, 324, 122238 CrossRef CAS.
- L. Liu, P. Ouyang, Y. Li, Y. Duan, F. Dong and K. Lv, J. Hazard. Mater., 2022, 439, 129637 CrossRef CAS PubMed.
|
This journal is © The Royal Society of Chemistry 2024 |
Click here to see how this site uses Cookies. View our privacy policy here.