DOI:
10.1039/D2QI01155G
(Research Article)
Inorg. Chem. Front., 2022,
9, 4753-4767
Photocatalytic CO2 reduction on Cu single atoms incorporated in ordered macroporous TiO2 toward tunable products†
Received
29th May 2022
, Accepted 19th July 2022
First published on 20th July 2022
Abstract
The photocatalytic conversion of CO2 to hydrocarbons is a fascinating strategy to defuse the growing energy and environmental dilemmas, but there are great challenges in improving photocatalytic efficiency and tuning product selectivity. Both are of equal importance to increase the yield of the desired product and clarify the photocatalytic CO2 reduction mechanism. Herein, a Cu single-atom-incorporated three-dimensional-ordered macroporous TiO2 (Cu0.01/3DOM-TiO2) photocatalyst was synthesized using a template-assisted in situ pyrolysis method. The Cu single atoms are uniformly anchored in a 3DOM TiO2 matrix, which not only broadens the light absorption range but also provides specific active sites for the adsorption and transformation of CO2 molecules. The photocatalytic CO2 reduction reaction was conducted in gas–solid and liquid–solid systems, respectively, to explore the effects of reaction modes on the CO2 conversion efficiency and product selectivity. The results indicate that the photocatalytic CO2 reduction reaction in gas–solid system mainly produces methane (CH4), with a high selectivity of 83.3% and a formation rate of 43.15 μmol g−1 h−1. In contrast, the main product in the liquid–solid system is ethylene (C2H4), with a selectivity of 58.4% and a formation rate of 6.99 μmol g−1 h−1. The Cu0.01/3DOM-TiO2 photocatalyst shows superior activity and selectivity in the gas–solid system while favorably producing C2H4 in the liquid–solid system. The possible photocatalytic mechanisms of CO2 reduction in the two different reaction systems are discussed according to in situ infrared spectroscopy. This work provides some new information on promoting photocatalytic CO2 reduction to desirable products by the rational design of photocatalysts as well as tuning the reaction conditions.
1. Introduction
Photosynthetic organisms harness solar energy to convert CO2 and H2O into carbohydrates and oxygen, which achieves energy transformation in nature and maintains the carbon–oxygen balance. Enlightened by natural photosynthesis, artificial photosynthesis for CO2 reduction with H2O powered by solar irradiation has attracted extensive interest, which is regarded as a double-benefit approach that converts greenhouse gas to chemical fuels and valuable chemicals by harvesting solar energy.1 However, the reactivity and product selectivity of photocatalytic CO2 reduction are always restricted by three main obstacles: (1) the difficult adsorption of reactants on photocatalysts in heterogeneous catalytic systems, (2) the high energy requirement for C–O bond activation up to 750 kJ mol−1, and (3) the various intermediates and complex pathways of the CO2 transformation process.2,3 Hence, the rational construction of heterogeneous photocatalysts is imperative for improving the reactivity and product selectivity of CO2 reduction, which requires the meticulous design and masterly integration of a light-harvesting component and catalytically active sites.4
In recent years, single-atom (SA) catalysts have emerged as new options for the design and preparation of cost-effective and high-efficiency photocatalysts, which provide 100% active metal dispersion and thus maximize metal utilization.5 The SA catalysts can be designed on the atomic scale by rationally embellishing the isolated specific metal-atom sites on the support materials.6 Based on the inherent functions of light absorption and photo-generated carriers supplying the support materials, SA active sites further offer catalytic centers for promoting charge separation/transfer and reactant adsorption/activation.7 Moreover, the definite active centers of SA photocatalysts enable improved understanding of the fundamental mechanism of photocatalysis, and allow us to draw more precise structure–performance correlations. Therefore, various metal cocatalysts including noble metals, transition metals and rare Earth metals, have been prepared in SA sites to increase the photocatalytic efficiency.8–12 Among them, copper (Cu) has been identified as one of the most promising cocatalysts to convert CO2 into various hydrocarbons owing to its optimum binding ability with CO2 and reaction intermediates.13–16 The individual Cu atoms immobilized on light-harvesting semiconductors can act as excellent active sites for activating CO2 molecules and stabilizing reaction intermediates in the photocatalytic CO2 reduction reaction.17,18 Moreover, the introduction of highly dispersed Cu heteroatoms in light-harvesting semiconductors can broaden the light-adsorption range and enhance the photoexcited charge separation efficiency.19 Although numerous studies have highlighted the excellent catalytic activity of Cu SAs in electrocatalytic/photocatalytic CO2 reduction,16,20,21 the controllable and large-scale production of heterogeneous photocatalysts with Cu SAs is still challenging. Most strategies to anchor Cu SAs on light-harvesting semiconductors are explored as post-treatment methods (e.g. impregnation and photodeposition), where the Cu SAs tend to aggregate into nanoparticles or leach from supports during photocatalytic reactions because of the weak combination.22,23 Hence, the in situ incorporation of highly stable Cu SAs in semiconductors is desirable. Meanwhile, the morphological features of the semiconductor support are a crucial factor to create stable Cu single atoms as well as heterogeneous photocatalysts with high performance, where the support not only determines the availability and stability of Cu SA sites but also supplies photogenerated electrons for CO2 reduction. Three-dimensional ordered macroporous (3DOM) materials with an interpenetrating porous structure and high porosity can greatly facilitate CO2 diffusion and enhance the light utilization simultaneously.24,25 Moreover, the large specific surface area of the 3DOM support can afford more accessible Cu SA sites for CO2 photoreduction.26 Although some metal-modified 3DOM TiO2 photocatalysts have been reported, the formation of metal nanoparticles with large sizes generally results in limited photocatalysis.27,28 Our recent study indicated that Pd single atoms could be in situ anchored in 3DOM CeO2via a template-assisted pyrolysis method owing to the strong metal–support interaction.29 Inspired by this, the masterly integration of catalytically active sites (Cu SAs) and a light-harvesting semiconductor (3DOM TiO2) is expected to achieve.
In addition, tuning the selectivity of CO2 reduction reaction to the desired product is of great significance but remains a huge challenge.30,31 Tan and coworkers found that the ethylene selectivity of electrocatalytic CO2 reduction with the same catalyst could be tuned by modulating the local CO2 concentration, where the kinetics of CO2 reduction was substantially influenced by the modulation of CO2 mass transport.32 Wang et al. reported that the introduction of N2 into CO2 could control the *CO coverage on Cu–Au bimetallic catalysts and thus facilitate *CO protonation for methane formation.33 However, the effects of reaction conditions on product selectivity of CO2 photoreduction are rarely studied. The photocatalytic CO2 reduction reaction is commonly conducted in either a liquid–solid or a gas–solid system. Actually, the reaction system is a non-negligible factor influencing the product selectivity of CO2 photoreduction. Therefore, exploring the effect of reaction systems on the product selectivity of photocatalytic CO2 reduction is as important as the development of efficient photocatalysts. In this work, a heterogeneous photocatalyst of 3DOM TiO2 doped with Cu single atoms (Cu0.01/3DOM-TiO2) is controllably synthesized via a facile template-assisted pyrolysis method. Taking this catalyst as a research model, the photocatalytic CO2 reduction performance is studied in gas–solid and liquid–solid reaction systems, respectively, to explore the effects of reaction modes on the catalytic activity and product selectivity. Interestingly, the Cu0.01/3DOM-TiO2 photocatalyst exhibits higher activity and selectivity for methane (CH4) production in the gas–solid system, while possessing a favorable capability to convert CO2 into ethylene (C2H4) in the liquid–solid system. The possible photocatalytic mechanisms of CO2 reduction in both catalytic systems are discussed according to in situ diffused reflection infrared Fourier transform spectroscopy (DRIFTS) measurements.
2. Experimental
2.1. Chemicals
Titanium butoxide (C16H36O4Ti, ≥99.0%), anhydrous cupric chloride (CuCl2, 98%), styrene (>99.5%, GC), potassium persulfate (K2S2O8, AR, 99.5%), sodium hydroxide (NaOH, 97%), acetic acid glacial (CH3COOH, ≥99.8%), ethanol (CH3CH2OH, ≥99.7%), methanol (CH3OH, 99.5%), sodium sulfate anhydrous (Na2SO4, 99%), potassium iodide (KI, 99%), potassium hydrogen phthalate (C8H5KO4, 99.8%), N2 (≥99.99%), CO2 (≥99.99%) and argon gas (≥99.99%) were obtained commercially and used without further purification.
2.2. Synthesis of Cu/3DOM-TiO2 photocatalysts
Firstly, the monodisperse polystyrene (PS) sub-microspheres with diameters of about 260 nm were prepared using an emulsifier-free emulsion polymerization process and assembled into the PS colloidal crystal templates (CCTs) by centrifugation.29 The as-prepared CCTs were immersed in methanol for 30 min and then dried for a moment to remove the superfluous methanol. Subsequently, the Cu/3DOM-TiO2 photocatalysts were prepared using a template-assisted pyrolysis method. Specifically, methanol (0.5 mL), acetic acid glacial (0.5 mL), titanium butoxide (2.72 g) and a certain amount of CuCl2 were mixed by vigorous stirring for 2 h to form a homogeneous precursor solution. The pretreated CCTs were immersed in the precursor solution under vacuum treatment. After being subjected to a vacuum for 1 h, the CCTs impregnated with precursor solution were collected, dried at room temperature, and calcined in air at 550 °C for 3 h (heating rate 1 °C min−1) to obtain the Cu/3DOM-TiO2 photocatalysts. The Cu content in the Cu/3DOM-TiO2 photocatalysts was adjusted by changing the dosage of CuCl2, and the as-synthesized photocatalyst was denoted as Cux/3DOM-TiO2, where x represents the molar ratio of Cu to Ti based on the dosages of CuCl2 and titanium butoxide. For example, the Cu0.01/3DOM-TiO2 photocatalyst was prepared by adding 10.8 mg CuCl2 to the precursor solution. In addition, the other Cux/3DOM-TiO2 samples (x = 0.0075, 0.0125, 0.02, 0.03, 0.04 and 0.05) were also synthesized by adding the corresponding dosage of CuCl2. The actual Cu contents of Cu0.0075/3DOM-TiO2, Cu0.01/3DOM-TiO2, Cu0.0125/3DOM-TiO2, Cu0.02/3DOM-TiO2, Cu0.03/3DOM-TiO2, Cu0.04/3DOM-TiO2 and Cu0.05/3DOM-TiO2 were measured by atomic absorption spectroscopy (AAS) as 0.51 wt%, 0.69 wt%, 1.03 wt%, 1.33 wt%, 1.98 wt%, 2.75 wt% and 3.38 wt%, respectively.
2.3. Synthesis of control samples
Several control samples including 3DOM-TiO2, Cu0.01/TiO2 and TiO2 were prepared using a similar procedure to that for preparing the Cu0.01/3DOM-TiO2 photocatalyst, without adding CuCl2, CCTs, CuCl2 and CCTs, respectively.
2.4. Characterization
The crystalline components of photocatalysts were analyzed using a powder X-ray diffractometer (XRD, DX-2700) with Cu Kα radiation. The actual Cu contents of the Cux/3DOM-TiO2 photocatalysts were measured on an atomic absorption spectrophotometer (Shimadzu AA-7000). The microstructures were observed using field emission scanning electron microscopy (FESEM, ZEISS Ultra-55) operated at 3.0 kV and transmission electron microscopy (TEM, JEOL-2010) working at 200 kV. Nitrogen adsorption/desorption isotherms were recorded on a Micromeritics ASAP 2460 apparatus at 77 K, to determine the specific surface areas and the pore size distributions of the photocatalysts. Thermogravimetric (TG) analysis was performed on a thermal gravimetric analyzer (TG209F3 Tarsus, Netzsch). Raman spectra were acquired on a Horiba Scientific LabRAM HR Evolution apparatus with a 514 nm laser as the excitation source. X-ray photoelectron spectra (XPS) were measured using a Thermo Scientific K-Alpha X-ray photoelectron spectrometer, and the binding energies were calibrated with a C 1s peak at 284.8 eV as the reference. UV–vis diffuse reflectance spectra (DRS) were taken on a Shimadzu UV-2600 spectrophotometer with BaSO4 as the reference. Photoluminescence (PL) spectra were measured on a HORIBA Fluoromax-4 spectrophotometer at an excitation wavelength of 350 nm. The radical intermediates generated during photocatalytic CO2 reduction were detected using in situ electron spin resonance (ESR) on a Bruker model JEOL JES-FA200 spectrometer, with 5,5-dimethyl-1-pyrroline-N-oxide (DMPO) as a radical-trapping reagent. The X-ray absorption of fine structures (XAFS) measurements of the Cu K-edge were conducted at the BL14W1 station at the Shanghai Synchrotron Radiation Facility (SSRF), China. The spectra were recorded in fluorescent mode using a Si (311) double-crystal monochromator and a 32-element Ge solid-state detector. The electron storage ring was operated at 3.5 GeV. Cu foil and CuO were employed as the reference samples. The XAFS data were processed using Athena and Artemis software, according to the standard procedures.
2.5. Photocatalytic CO2 reduction tests
The photocatalytic CO2 reduction reactions were carried out in a top-irradiated reactor with a volume of 170 mL. The simulated sunlight with a main irradiation range of 320–780 nm was provided by a Xe lamp (PLS-SXE300, Beijing Perfectlight) equipped with a specific reflector, and the optical power density irradiated on the reaction solution was 200 mW cm−2. The photocatalytic CO2 reduction reaction was conducted in gas–solid and liquid–solid systems, respectively. For the gas–solid reaction mode, the slurry prepared by dispersing 5 mg photocatalyst in 1 mL ethanol was coated on a glass plate (2 × 2 cm) and then dried at 60 °C. The glass-supported photocatalyst was placed at the bottom of the reactor, and 1 mL of deionized water was dripped beside the glass plate. After CO2 gas was blown into the reactor and absorption equilibrium was achieved in dark, the photocatalytic CO2 reduction reaction was started under simulated sunlight. For the liquid–solid reaction mode, the reaction suspension containing 5 mg of photocatalyst and 30 mL of water was transformed in the reactor, which was bubbled with CO2 gas for 20 min. After the sealed reactor had been kept in dark for 30 min, the photocatalytic CO2 reduction reaction was started under simulated sunlight from a Xe lamp. The photocatalytic reactions were conducted for 4 h to obtain the average generation rates of products. The gas products (CO, CH4, C2H4) were quantitatively analyzed using a gas chromatograph (GC 9790 II, FuLi) equipped with both a thermal conductivity detector (TCD) and a flame ionization detector (FID). The liquid products were identified using 1H nuclear magnetic resonance (NMR) spectrometry in D2O, which was performed on a Bruker AVANCE AV 400 MHz spectrometer. For the recycling experiments, the reaction system was re-blown or re-bubbled with CO2 gas after each reaction cycle and then irradiated for the next reaction cycle.
2.6.
In situ DRIFTS measurement
The in situ diffused reflection infrared Fourier transform spectroscopy (DRIFTS) analysis was performed on a Bruker TENSOR II FTIR spectrometer equipped with a mercury–cadmium–tellurium (MCT) detector cooled with liquid nitrogen. For the gas–solid reaction mode, 30 mg of the photocatalyst was placed in the test chamber, which was then subjected to a vacuum. After the background spectrum was collected from 400 to 4000 cm−1, humid CO2 gas was introduced into the test chamber to achieve adsorption–desorption equilibrium in the dark for 30 min. Subsequently, the time-dependent DRIFTS spectra were recorded upon irradiation from the Xe lamp. For the liquid–solid reaction mode, the operations were the same as those described above, except that a drop of deionized water was dripped beside the photocatalyst on the sample holder.
3. Results and discussion
3.1. Structures and properties of photocatalysts
The Cux/3DOM-TiO2 photocatalysts were synthesized via a template-assisted pyrolysis method as illustrated in Fig. 1a. The monodisperse PS sub-microspheres were assembled into colloidal crystal templates which were impregnated with the precursor solution containing Cu2+ ions and titanium butoxide. The Cux/3DOM-TiO2 photocatalysts with different Cu contents were obtained via calcination of the colloidal crystal templates loaded with precursors in air at 550 °C for 3 h. The thermogravimetric (TG) analysis (Fig. S1, ESI†) indicated that the PS-sphere template decomposes at 400 °C in air and the precursors are completely converted into the final material below 500 °C. Therefore, it is reasonable to conduct the calcination at 550 °C for 3 h, during which the individual Cu single atoms are in situ anchored in the 3DOM-TiO2 matrix. The XRD patterns of the typical Cu0.01/3DOM-TiO2 photocatalyst and the control samples (3DOM-TiO2, Cu0.01/TiO2 and TiO2) are shown in Fig. 1b. The photocatalysts obtained from the template-assisted synthesis (Cu0.01/3DOM-TiO2 and 3DOM-TiO2) consist of anatase TiO2 (JCPDS no. 01-562) accompanied by a small quantity of rutile TiO2 (JCPDS no. 01-1292), while the Cu0.01/TiO2 and TiO2 samples synthesized by the direct pyrolysis of precursors without the PS template present only the diffraction peaks of anatase TiO2. The mass fractions of rutile TiO2 (WRutile) in Cu0.01/3DOM-TiO2 and 3DOM-TiO2 are estimated to be 13.86 wt% and 13.80 wt%, respectively, according to the previously reported formula:34,35WRutile = ARutile/(0.884 AAnatase + ARutile), where AAnatase and ARutile are the integrated intensities of the anatase (101) and rutile (110) XRD peaks. The previous study demonstrated that the calcination process assisted by a PS colloidal crystal template could produce mixed phases of anatase and rutile.27 As a light-harvesting material, TiO2 containing both anatase and rutile phases is desirable because the anatase/rutile junction facilitates spatial charge separation, allowing more photogenerated charge carriers to migrate to the surface and participate in redox reactions.36 The diffraction peaks of Cu species are not observed in the XRD patterns of both the Cu0.01/3DOM-TiO2 and the Cu0.01/TiO2 samples due to the high dispersity and/or low content of Cu. Compared with TiO2 and 3DOM-TiO2, the typical diffraction peaks of Cu0.01/3DOM-TiO2 and Cu0.01/TiO2 present obvious shifts to lower 2θ positions (Fig. S2, ESI†), which are caused by the doping of Cu cations with larger ionic radii (Cu2+, 0.73 Å; Cu+, 0.77 Å; and Ti4+, 0.61 Å) into TiO2.37,38 In addition, the anatase (101) diffraction peak of Cu0.01/3DOM-TiO2 and 3DOM-TiO2 presents further shifts towards lower 2θ positions compared with those of Cu0.01/TiO2 and TiO2, possibly because the special 3DOM structures result in little lattice distortion.39,40 The SEM images (Fig. 1c; Fig. S3a, ESI†) affirm that the Cu0.01/3DOM-TiO2 photocatalyst exhibits a honeycomb morphology with 3D ordered macroporous structures, where the spherical voids have an average diameter of 160 nm and the walls have an average thickness of about 10 nm. These spherical voids are well arranged and interconnected. The 3DOM-TiO2 sample (Fig. S3b, ESI†) presents a similar honeycombed morphology to that of the Cu0.01/3DOM-TiO2 photocatalyst. In contrast, Cu0.01/TiO2 (Fig. S3c, ESI†) and TiO2 (Fig. S3d, ESI†) synthesized by the direct pyrolysis of precursors without a PS template are irregular bulks without porosity. The TEM image shown in Fig. 1d indicates that the 3DOM frameworks are constructed from numerous TiO2 small particles. The HRTEM image (Fig. 1e) reveals the crystalline phase boundary of anatase and rutile, where the characteristic spacings of 0.352 and 0.229 nm are assignable to the (101) lattice of anatase TiO2 and the (200) lattice of rutile TiO2, respectively. The crystalline lattices of CuO or Cu2O are not found in the HRTEM image of Cu0.01/3DOM-TiO2, suggesting that Cu element exists as highly dispersed species in this material.
|
| Fig. 1 (a) Schematic diagram of the synthesis of Cu0.01/3DOM-TiO2 photocatalyst. (b) XRD patterns of different samples; (c) SEM, (d) TEM and (e) HRTEM images of Cu0.01/3DOM-TiO2. | |
The specific surface areas and pore size distributions of Cu0.01/3DOM-TiO2, 3DOM-TiO2 and Cu0.01/TiO2 were determined using N2 adsorption–desorption measurements (Fig. 2; Fig. S4, ESI†). Both Cu0.01/3DOM-TiO2 and 3DOM-TiO2 samples display the typical type IV isotherm with a hysteresis loop, and the dominant adsorption of both samples can be identified as macropores corresponding to the step at P/P0 = 0.8–0.995 (Fig. 2a).41 In contrast, Cu0.01/TiO2 presents an extremely weak N2 adsorption capacity (Fig. S4, ESI†) due to its large bulk morphology without porosity. Cu0.01/3DOM-TiO2 presents a specific surface area of 52.80 m2 g−1 and a pore volume of 0.28 cm3 g−1, which are higher than those of 3DOM-TiO2 (Table S2, ESI†). The Cu0.01/TiO2 sample shows a very low specific surface area (3.74 m2 g−1) and a negligible pore volume (0.0014 cm3 g−1). The large surface area and pore volume of Cu0.01/3DOM-TiO2 are conducive to the sufficient exposure of active sites and the easy diffusion of CO2 and H2O molecules during the photocatalytic CO2 reduction reaction.42
|
| Fig. 2 (a) N2 adsorption–desorption isotherms and (b) pore size distribution plots of Cu0.01/3DOM-TiO2 and 3DOM-TiO2. | |
The Raman spectroscopy analysis shed light on the variations in local structures of TiO2 in different photocatalysts. As shown in Fig. 3a, the Raman spectra of the three samples (Cu0.01/3DOM-TiO2, 3DOM-TiO2 and Cu0.01/TiO2) exhibit five representative peaks of anatase TiO2, which can be ascribed to the Eg(1), Eg(2) and Eg(3) modes (144, 197 and 639 cm−1), the B1g mode (397 cm−1) and the doublet of the A1g/B1g mode (517 cm−1), respectively.41,43 Compared with Cu0.01/TiO2, 3DOM-TiO2 shows a slight shift and broadening in the characteristic A1g and B1g vibration peaks due to the increased structural disorder resulting from the formation of the 3DOM structure. Cu0.01/3DOM-TiO2 presents further broadening of the characteristic peaks because of Cu doping in 3DOM TiO2, but no peaks of copper oxide or copper hydroxide are observed. Moreover, both 3DOM-TiO2 and Cu0.01/3DOM-TiO2 show a weak peak at 448 cm−1 (inset in Fig. 3a), which corresponds to a characteristic vibration mode of rutile TiO2, further confirming the coexistence of the anatase and rutile phases in the two samples.44 The XPS technique was employed to acquire information on the surface elemental compositions and chemical states of the photocatalysts. For the three samples (Cu0.01/3DOM-TiO2, 3DOM-TiO2 and Cu0.01/TiO2), each of the Ti 2p spectra (Fig. 3b) shows two characteristic peaks at binding energies of 458.6 and 464.3 eV, which are assignable to Ti 2p3/2 and 2p1/2 of Ti4+, respectively.45,46 Compared with Cu0.01/TiO2, the Ti 2p3/2 and 2p1/2 peaks of both 3DOM-TiO2 and Cu0.01/3DOM-TiO2 present a 0.1 eV positive shift, which is probably caused by the lattice distortions of 3DOM-TiO2.47 Two peaks are deconvoluted from each O 1s spectrum (Fig. 3c), where the main peak around 529.8 eV is assigned to the lattice oxygen of TiO2 and the minor peak around 531.0 eV is attributed to the surface chemisorbed oxygen on oxygen-vacancy (VO) sites.48–50 Cu0.01/3DOM-TiO2 and 3DOM-TiO2 have higher oxygen-vacancy contents (34.63% and 30.74%) than Cu0.01/TiO2 (24.87%), because the high specific surface area of 3DOM-TiO2 exposes more surface Ti sites with unsaturated coordination.51 In Fig. 3d, two characteristic peaks centered at binding energies of 932.3 and 952.0 eV appear in the Cu 2p spectra of both Cu0.01/3DOM-TiO2 and Cu0.01/TiO2, which are approximate to the Cu 2p3/2 and 2p1/2 of Cu+ ions,52,53 indicating that the Cu single atoms incorporated in TiO2 have an electronic state close to that of the Cu+ cation. In addition, another pair of minor peaks centered at 934.0 and 954.3 eV deconvoluted from the Cu 2p spectrum of Cu0.01/TiO2 can be assigned to a Cu2+ cation with a higher oxidation state, probably due to the formation of CuO clusters.54,55 Compared with 3DOM-TiO2, the lack of peak shifts in the XPS spectra of Cu0.01/3DOM-TiO2 may be ascribable to the formation of the atomically dispersed Cuδ+ in the lattice of 3DOM-TiO2 during the high-temperature calcination process as well as the low Cu content. Furthermore, the electronic structure and local coordination environment of Cu atoms in the Cu0.01/3DOM-TiO2 photocatalyst were studied using X-ray absorption fine structure (XAFS) analysis, in comparison with Cu foil and CuO. Fig. 3e shows the normalized Cu K-edge X-ray absorption near-edge structure (XANES) spectra, where the adsorption edge of Cu0.01/3DOM-TiO2 is located between those of Cu foil and CuO, indicating that the electronic state of Cu in Cu0.01/3DOM-TiO2 lies between those of Cu0 and Cu2+.52 The adsorption peak of Cu0.01/3DOM-TiO2 is centered at 8997.1 eV and shifts a little to a lower energy compared with CuO, meaning a higher electron density of Cu single atoms in this photocatalyst than for Cu2+ in CuO.56 The coordination configuration of Cu atoms in Cu0.01/3DOM-TiO2 is further revealed by the extended X-ray adsorption fine structure (EXAFS) spectrum (Fig. 3f). The Cu K-edge EXAFS spectrum of Cu0.01/3DOM-TiO2 presents two characteristic distances. The major peak centered at 1.4 Å corresponds to the direct binding of Cu atoms with lattice oxygen (Cu–O), while the minor peak located at 2.4 Å is assignable to Cu–Ti coordination in the TiO2 environment.57 No peak of Cu–Cu metallic binding is observed, indicating that the Cu atoms are individually doped into the lattice of TiO2 and atomically distributed in the 3DOM-TiO2 support.
|
| Fig. 3 (a) Raman spectra of different photocatalysts. XPS spectra of (b) Ti 2p, (c) O 1s, and (d) Cu 2p of different photocatalysts. (e) Cu K-edge XANES and (f) EXAFS spectra of Cu0.01/3DOM-TiO2 with the reference samples of Cu foil and CuO. | |
High light-harvesting and charge separation efficiencies are fundamental requirements for efficient photocatalysis. The photoabsorption properties of bulk TiO2, 3DOM-TiO2, Cu0.01/TiO2 and Cu0.01/3DOM-TiO2 were measured using UV–vis diffuse reflectance spectroscopy (UV–vis DRS), as shown in Fig. 4a. The bulk TiO2 shows absorption only in the ultraviolet region (λ < 400 nm) corresponding to the intrinsic transition from valence band (VB) to conduction band (CB). 3DOM-TiO2 exhibits a great enhancement in its ultraviolet-harvesting ability owing to the unique 3DOM structure as well as additional absorption from 400 nm to longer wavelengths due to the formation of more oxygen vacancies.58 With the introduction of Cu atoms, the photoabsorption above 400 nm becomes more conspicuous, suggesting the presence of more VO defect states in TiO2.19 Moreover, the Cu-incorporated TiO2 samples (Cu0.01/TiO2 and Cu0.01/3DOM-TiO2) exhibit a near-infrared absorption band near 800 nm, corresponding to d–d transitions of the dopant Cu cations.59 For Cu0.01/TiO2, the photoabsorption decreases in the ultraviolet region while increasing greatly in the visible region, which is attributable to the formation of CuO on TiO2. Photographs of the different photocatalysts are provided in Fig. 4a and Fig. S12 (ESI†), respectively. Typically, the Cu0.01/3DOM-TiO2 photocatalyst is light yellow while the Cu0.01/TiO2 sample is dark green. The color of the Cux/3DOM-TiO2 photocatalysts changes from light yellow to brown as the Cu content increases gradually (Fig. S12, ESI†). Photoluminescence (PL) spectroscopy is employed to evaluate the separation efficiency of the photogenerated electron–hole pairs under excitation of 300 nm. The Cu0.01/3DOM-TiO2 photocatalyst exhibits the weakest PL emission (Fig. 4b), suggesting the most efficient separation of photogenerated charge carriers.60 The photocurrent responses and electrochemical impedance spectra (EIS) of photocatalysts were measured to further verify their photoexcited charge separation and transfer efficiencies. As shown in Fig. 4c, Cu0.01/3DOM-TiO2 shows the highest photocurrent response with a light on/off mode, reflecting the most efficient separation and transfer of photoexcited charge carriers. The EIS Nyquist plot of Cu0.01/3DOM-TiO2 presents the smallest arc radius (Fig. 4d), corresponding to the fastest charge transfer. These results demonstrate that the incorporation of Cu single atoms into 3DOM-TiO2 affords outstanding superiority in enhancing light absorption and photogenerated charge separation/transfer efficiency. The highly dispersed Cu single-atom incorporation in TiO2 enables efficient electron transfer via the Cu+–Cu2+ shift.22 The unique structure and properties of Cu0.01/3DOM-TiO2 make it a fascinating photocatalyst for CO2 conversion.
|
| Fig. 4 (a) Kubelka–Munk transformed UV–vis absorption spectra, (b) photoluminescence spectra, (c) transient photocurrent responses, (d) electrochemical impedance spectra of different photocatalysts. | |
3.2. Photocatalytic CO2 reduction performance
Two typical heterogeneous photocatalytic reaction systems of gas–solid and liquid–solid modes were employed to evaluate the performance of the Cu0.01/3DOM-TiO2 photocatalyst toward CO2 reduction with water under simulated sunlight. The effect of the Cu content in Cu0.01/3DOM-TiO2 on the photocatalysis was probed, as shown in Fig. S5,† and a volcano relationship was obtained for the photocatalytic CO2 reduction performance relative to the Cu content. The Cu0.01/3DOM-TiO2 photocatalyst with a Cu/Ti molar ratio of 1% exhibited optimal performance in both reaction systems. Interestingly, the main product of the photocatalytic CO2 reduction reaction in the gas–solid system was CH4, while that in the liquid–solid system was C2H4 (Fig. S6, ESI†). Fig. 5a shows the time-dependent evolution of CH4 and CO over the Cu0.01/3DOM-TiO2 photocatalyst in the gas–solid catalytic system, without H2 evolution. As a main product, CH4 was generated predominately from CO2 reduction in the gas–solid catalytic reaction, with a remarkable production rate of 43.15 μmol gcat−1 h−1 and a high selectivity of 83.3% (Fig. 5b). Notably, in the liquid–solid catalytic reaction, the Cu0.01/3DOM-TiO2 photocatalyst exhibits an excellent performance for CO2 reduction to C2H4 (Fig. 5d), accompanied by the generation of CO and CH4 by-products. The production rate of C2H4 reaches 6.99 μmol gcat−1 h−1, with a selectivity of 58.4%, which is of great significance due to the crucial role of C2H4 in the chemical industry. No liquid product is detected in the reaction solution after 4 h of irradiation (Fig. S7, ESI†). In comparison, the Cu0.01/3DOM-TiO2 photocatalyst presents a higher activity and product selectivity for CO2 reduction in the gas–solid catalytic reaction while producing more valuable C2H4 in the liquid–solid catalytic reaction. These results indicate that the photocatalytic CO2 reduction reactions over Cu0.01/3DOM-TiO2 in the two catalytic systems actually follow different reaction pathways. The promotion effects of Cu doping and the 3DOM structure on CO2 conversion are testified by the control experiments of photocatalytic reactions with 3DOM-TiO2, Cu0.01/TiO2 and TiO2, respectively. As shown in Fig. 5b and e, all the control catalysts without Cu doping or 3DOM structure present very poor photocatalysis toward CO2 reduction in both the gas–solid and the liquid–solid systems, and none of them has the ability to convert CO2 into CH4 or C2H4, affirming that both Cu single-atom incorporation and the 3DOM structure are crucial for enhancing the photocatalytic CO2 reduction performance. Subsequently, the durability of the Cu0.01/3DOM-TiO2 photocatalyst was examined by four cycles of the photocatalytic CO2 reduction reaction with 4 h of irradiation for each cycle in the gas–solid and liquid–solid reaction systems (Fig. 5c and f), respectively. After each cycle, the photocatalyst was exposed to air for recovery. Both the excellent activity and the high selectivity were well maintained after four cycles of the photocatalytic reaction, indicating the good durability of the Cu0.01/3DOM-TiO2 photocatalyst. The XPS analyses (Fig. S8, ESI†) indicate that the surface composition of the Cu0.01/3DOM-TiO2 catalyst remains unchanged after the long-term photocatalytic CO2 reduction reaction in both the gas–solid and liquid–solid systems, affirming the good stability of the photocatalyst. In addition, several control experiments of the photocatalytic CO2 reduction reaction were carried out to provide evidence of the CO2 conversion in both catalytic reaction systems. As shown in Fig. S9 (ESI†), the products are barely generated in the reaction system without CO2, light or catalyst, indicating that all the products originate from the photocatalytic CO2 reduction reaction. When CO2 is replaced by N2 in the reaction systems, small amounts of CO and CH4 are produced at negligible rates over the Cu0.01/3DOM-TiO2 photocatalyst (Fig. S9b and S9d, ESI†), and the production stops after 2 h of irradiation. The slight production of CO and CH4 can be ascribed to the reduction of surface-adsorbed CO2 on the photocatalyst by air.
|
| Fig. 5 Photocatalytic CO2 reduction performance. (a–c) Gas–solid catalytic reaction: (a) product evolutions over Cu0.01/3DOM-TiO2, (b) generation rates of products over different photocatalysts, (c) cyclic test of Cu0.01/3DOM-TiO2. (d–f) Liquid–solid catalytic reaction: (d) product evolutions over Cu0.01/3DOM-TiO2, (e) generation rates of products over different photocatalysts, (f) cycling test of Cu0.01/3DOM-TiO2. | |
3.3. Photocatalytic reaction mechanism
The photoinduced charge transfer from the catalyst to the reactant molecules is critical for photocatalytic reactions. To probe the interfacial charge kinetics, the photocurrent responses of photocatalysts were measured in different atmospheres. The transient photocurrent recorded under an Ar atmosphere is assignable to the electron transfer from catalyst to electrode. Notably, the photocurrent measured in the CO2-saturated electrolyte is lower than that measured in the Ar-saturated electrolyte for each photocatalyst (Fig. 6a and b), which is caused by the competitive interfacial electron transfer from the photocatalyst to the surface-adsorbed CO2 molecules.61 In comparison, the photocurrents of the Cu-incorporated photocatalysts (Cu0.01/3DOM-TiO2 and Cu0.01/TiO2) decline more than those of the photocatalysts without Cu (3DOM-TiO2 and TiO2), indicating that the incorporation of Cu is greatly beneficial to the electron transfer from the photocatalyst to the CO2 molecules. In particular, the Cu0.01/3DOM-TiO2 photocatalyst shows the greatest decrease in photocurrent measured in the CO2-saturated electrolyte, corresponding to the most efficient electron delivery from the catalyst to the CO2 molecules. It is understandable that the high charge separation and transfer efficiencies as well as the favorable CO2 adsorption ability of the Cu0.01/3DOM-TiO2 photocatalyst jointly contributes to enhancing the electron provision to CO2 molecules. The significantly enhanced electron transfer efficiency can accelerate the CO2 activation and the subsequent proton-coupled electron transfer process for hydrocarbon generation.62 Furthermore, in situ ESR detection reveals the generation of radical intermediates during photocatalytic CO2 reduction with water over the Cu0.01/3DOM-TiO2 catalyst (Fig. 6c). After 10 min of irradiation, the characteristic signal of the carbon-centered radical is obviously detected, affirming the activation of chemisorbed CO2 on the photocatalyst.63 Meanwhile, the characteristic signal of the hydroxyl radical (˙OH) is also observed in the photocatalytic reaction system, which should stem from the oxidation of surface-adsorbed H2O molecules by consuming photogenerated holes (h+).64 Accordingly, hydrogen peroxide (H2O2) as the product of the oxidation half reaction in both the photocatalytic CO2 reduction systems is determined by iodometry (Fig. 6d).65 The H2O2 production rates in gas–solid and liquid–solid reaction systems of the Cu0.01/3DOM-TiO2 photocatalyst are 16.85 and 15.12 μmol gcat−1 h−1, respectively, which are much higher than those of other photocatalysts (Fig. S10, ESI†). The results demonstrate that H2O2 has been produced from the oxidation of surface-adsorbed H2O by the photogenerated holes and also confirm the superior activity of the Cu0.01/3DOM-TiO2 photocatalyst toward CO2 reduction with water.
|
| Fig. 6 (a, b) Transient photocurrents of different catalysts measured in the Ar or CO2 saturated electrolyte, (c) in situ ESR detection of the photocatalytic CO2 reduction system with Cu0.01/3DOM-TiO2, (d) UV–vis absorption spectra of the iodometry solution of photocatalytic CO2 reduction over different photocatalysts in gas–solid and liquid–solid systems. | |
In situ DRIFTS spectra collected from the photocatalytic reaction systems are helpful to identify the surface-bound intermediates of CO2 transformation and provide crucial clues for understanding the reaction pathway. In situ DRIFTS measurements were carried out for both the gas–solid and the liquid–solid catalytic reaction systems under Xe lamp irradiation. In both cases, four broad absorption bands appear in the 3750–3550 cm−1 region (Fig. S11, ESI†). The bands centered at 3727 and 3707 cm−1 are attributable to the stretching vibrations of surface-adsorbed –OH groups on Ti4+ and Ti3+, respectively.66 The bands peaking at 3626 and 3600 cm−1 correspond to the stretching vibrations of O–H in HCO3− and H2O, respectively.67 In comparison, the intensities of the four absorption bands in the liquid–solid catalytic system are much higher than those of the gas–solid catalytic system due to the water-rich environment of the former. The infrared adsorption signals of the CO2-reduction intermediates are displayed in Fig. 7. The peak at 1542 cm−1 in the gas–solid catalytic system is attributable to the surface-bound monodentate carbonate (m-CO32−), which is located at 1536 cm−1 for the liquid–solid catalytic system.68,69 Some infrared adsorption peaks associated with the CO2 transformation intermediates are obviously detected in both photocatalytic systems. The broad peak at 1448 cm−1 (Fig. 7a) or 1444 cm−1 (Fig. 7b) is ascribed to *COOH, a typical one-electron-reduction intermediate.70 The emergence of the peaks at 1346 and 1559 cm−1 (1350 and 1548 cm−1) is indicative of the formation of *HCOOH via the protonation process of *COOH.18 The adsorption bands in the region of 1614–1774 cm−1 are assignable to *CHO, which is an intermediate generated from *HCOOH.18,71,72 Notably, the infrared signals of the gas–solid catalytic system (Fig. 7a) are much stronger than those of the liquid–solid system (Fig. 7b), due to the higher conversion rate of the former. The low solubility of CO2 in water restricts the reaction rate of the photocatalytic CO2 reduction in the liquid–solid system. In addition, several distinctive signals are observed in the in situ DRIFTS of both catalytic systems, which provide crucial evidence to reveal the differences in the photocatalytic CO2 reduction pathways. For the gas–solid catalytic system (Fig. 7a), the sharp peak at 1455 cm−1 appearing after prolonged irradiation is attributed to the *CH3 group originating from *CHO protonation,38,73 which is generally considered as a key intermediate for CH4 production. For the liquid–solid catalytic system (Fig. 7b), the characteristic peak of *CH3 is absent, but a new peak at 2076 cm−1 is increasingly apparent with the irradiation, which can be assigned to the chemisorbed *CO.70,74 The previous report proved that the intermolecular interaction between surface-adsorbed CO and interfacial water is critical for the formation of ethylene during the electrochemical reduction of CO.75 The interfacial water is helpful to stabilize the CO dimer, a key intermediate in ethylene formation, by hydrogen bonding to the terminal oxygens of the adsorbed CO molecules. The porous structure of the Cu0.01/3DOM-TiO2 photocatalyst provides a water-rich local environment during CO2 reduction in the liquid–solid system, which is beneficial to the stabilization and dimerization of the *CO intermediate.76
|
| Fig. 7
In situ DRIFTS of the photocatalytic CO2 reduction with Cu0.01/3DOM-TiO2 catalyst in two different reaction systems: (a) gas–solid catalytic system, (b) liquid–solid catalytic system. | |
According to the above analyses, the photocatalytic mechanism and possible CO2 reduction pathways over the Cu0.01/3DOM-TiO2 photocatalyst in two different systems are illustrated in Fig. 8. 3DOM-TiO2 is excited to generate electron–hole pairs under simulated sunlight, and the photogenerated electrons are efficiently captured by the positively charged Cuδ+ (0 < δ ≤ 2) single-atom sites. The enhanced electron density at Cu sites facilitates the conversion of CO2 into hydrocarbons (CH4 or C2H4) via the multi-proton coupled multi-electron reduction process. Specifically, the adsorption and activation of CO2 molecules are achieved at the Cuδ+ active sites of photocatalyst, where the adsorbed CO2 undergoes a continuous hydrogenation process to form *COOH, followed by the generation of *HCOOH and *CHO intermediates. In the gas–solid photocatalytic system (Fig. 8a), CH4 is produced as the main product via the hydrogenation process of *CHO. On the other hand, it is widely accepted that the generation of the C2 product undergoes three different reaction paths: (1) a *COCO intermediate from *CO dimerization, (2) a *COCHO intermediate from the dimerization between *CO and *CHO, and (3) a *CHOCHO intermediate from *CHO dimerization.71,77,78 In the liquid–solid system of photocatalytic CO2 reduction (Fig. 8b), the coexistence of *CO and *CHO intermediates is a crucial clue to speculate that the formation of C2H4 follows the pathway of the *COCHO intermediate from dimerization between *CO and *CHO. Therefore, the possible reaction pathways of photocatalytic CO2 reduction in the present gas–solid and liquid–solid systems are described in eqn (1)–(5) and (6)–(11), respectively, where * denotes the catalytically active sites on the photocatalyst and ↑ represents the release of the gas product.
| * + CO2 + e− + H+ → *COOH | (1) |
| *COOH + e− + H+ → *HCOOH | (2) |
| *HCOOH + e− + H+ → *CHO + H2O | (3) |
| *CHO + 4e− + 4H+ → *CH3 + H2O | (4) |
| * + CO2 + e− + H+ → *COOH | (6) |
| *COOH + e− + H+ → *CO + H2O | (7) |
| *COOH + e− + H+ → *HCOOH | (8) |
| *HCOOH + e− + H+ → *CHO + H2O | (9) |
| *CO–CHO + 7e− + 7H+ → C2H4↑ + 2H2O | (11) |
|
| Fig. 8 The possible mechanisms of photocatalytic CO2 reduction over Cu0.01/3DOM-TiO2 catalyst in two different reaction systems: (a) gas–solid catalytic system, (b) liquid–solid catalytic system. | |
The oxidation half-reactions in both the catalytic systems are the formation of H2O2via the consumption of photogenerated holes by water (eqn (12) and (13)).
The photocatalytic CO2 reduction reaction in liquid–solid system is restricted by the low solubility of CO2 in water (33 mM at 1 atm and 25 °C) and the poor mass transfer of CO2 in the liquid phase.79 However, the diffusion coefficient of CO2 in the gas phase (∼0.1 cm2 s−1) is about 10000 times that in the liquid phase (∼1 × 10−5 cm2 s−1), which allows the rapid delivery of CO2 molecules to the surface of photocatalysts with uncovered active sites. The adsorption and activation of CO2 and H2O can be implemented more efficiently in the gas–solid system, and thus the photocatalyst shows higher activity.
4. Conclusion
In summary, the Cu single atom-incorporated 3DOM-TiO2 photocatalyst was synthesized using a low-cost and mass-production template-assisted pyrolysis strategy and further demonstrated for solar-powered CO2 conversion into hydrocarbons. The 3DOM structure with large surface area maximizes the exposed sites for anchoring Cu single atoms in a TiO2 matrix, which enhances the light absorption as well as affording more accessible active sites for CO2 reduction. The photocatalytic CO2 reduction with H2O was conducted in the gas–solid and liquid–solid systems, respectively, to explore the effects of reaction modes on CO2 conversion efficiency and product selectivity. The Cu0.01/3DOM-TiO2 photocatalyst exhibits higher activity and selectivity for CH4 production from CO2 reduction in the gas–solid system while showing a favorable performance for converting CO2 to C2H4 in the liquid–solid system. Both the excellent activity and the selectivity are well maintained after a long irradiation time, indicating the good durability of the Cu0.01/3DOM-TiO2 photocatalyst. Compared with the 3DOM-TiO2-based photocatalysts reported previously (Table S8, ESI†), the Cu0.01/3DOM-TiO2 catalyst exhibits superior photocatalytic CO2 reduction performance without using any noble metal cocatalysts. The photocatalytic mechanisms in both reaction systems were investigated by in situ DRIFTS analyses. It is proposed that CH4 is generated as the main product via the *CHO intermediate in the gas–solid system while C2H4 is produced via the dimerization of *CO and *CHO in the liquid–solid system. This work develops a scalable strategy for preparing Cu-SA-incorporated photocatalysts but also offers a new idea for regulating the product selectivity of CO2 photoreduction by altering the reaction system.
Author contributions
Cong Chen: methodology, investigation, data curation, visualization, writing – original draft. Ting Wang: methodology, investigation, validation. Ke Yan: investigation, validation. Shoujie Liu: software, data curation. Yu Zhao: software, funding acquisition. Benxia Li: supervision, conceptualization, writing – review & editing, funding acquisition.
Conflicts of interest
The authors declare no conflict of interest.
Acknowledgements
This work was supported by the National Natural Science Foundation of China (21471004), the Open Research Fund Program of Key Laboratory of Surface & Interface Science of Polymer Materials of Zhejiang Province (SISPM-2021-02) and the Science Foundation of Zhejiang Sci-Tech University (17062002-Y).
Notes and references
- X. B. Li, Z. K. Xin, S. G. Xia, X. Y. Gao, C. H. Tung and L. Z. Wu, Semiconductor nanocrystals for small molecule activation via artificial photosynthesis, Chem. Soc. Rev., 2020, 49, 9028–9056 RSC.
- C. Gao and Y. Xiong, Solar-driven artificial carbon cycle, Chin. J. Chem., 2021, 40, 153–159 CrossRef.
- E. Gong, S. Ali, C. B. Hiragond, H. S. Kim, N. S. Powar, D. Kim, H. Kim and S.-I. In, Solar fuels: Research and development strategies to accelerate photocatalytic CO2 conversion into hydrocarbon fuels, Energy Environ. Sci., 2022, 15, 880–937 RSC.
- T. Kong, J. Low and Y. Xiong, Catalyst: how material chemistry enables solar-driven CO2 conversion, Chem, 2020, 6, 1035–1038 CAS.
- Z.-H. Xue, D. Luan, H. Zhang and X. W. Lou, Single-atom catalysts for photocatalytic energy conversion, Joule, 2022, 6, 92–133 CrossRef CAS.
- S. Shin, R. Haaring, J. So, Y. Choi and H. Lee, Highly durable heterogeneous atomic catalysts, Acc. Chem. Res., 2022, 55, 1372–1382 CrossRef CAS PubMed.
- L. Zeng and C. Xue, Single metal atom decorated photocatalysts: Progress and challenges, Nano Res., 2020, 14, 934–944 CrossRef.
- J. Wei, F. L. Meng, T. Li, T. Zhang, S. Xi, W. L. Ong, X. Q. Wang, X. Zhang, M. Bosman and G. W. Ho, Spontaneous atomic sites formation in wurtzite CoO nanorods for robust CO2 photoreduction, Adv. Funct. Mater., 2021, 32, 2109693 CrossRef.
- X. Xiong, C. Mao, Z. Yang, Q. Zhang, G. I. N. Waterhouse, L. Gu and T. Zhang, Photocatalytic CO2 reduction to CO over Ni single atoms supported on defect-rich zirconia, Adv. Energy Mater., 2020, 10, 2002928 CrossRef CAS.
- Z. Chen, S. Wu, J. Ma, S. Mine, T. Toyao, M. Matsuoka, L. Wang and J. Zhang, Non-oxidative coupling of methane: N-type doping of niobium single atoms in TiO2–SiO2 induces electron localization, Angew. Chem., Int. Ed., 2021, 60, 11901–11909 CrossRef CAS PubMed.
- L. Luo, L. Fu, H. Liu, Y. Xu, J. Xing, C.-R. Chang, D.-Y. Yang and J. Tang, Synergy of Pd atoms and oxygen vacancies on In2O3 for methane conversion under visible light, Nat. Commun., 2022, 13, 2930 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, Rare-Earth single erbium atoms for enhanced photocatalytic CO2 reduction, Angew. Chem., Int. Ed., 2020, 59, 10651–10657 CrossRef CAS PubMed.
- S. Ali, A. Razzaq, H. Kim and S.-I. In, Activity, selectivity, and stability of Earth-abundant CuO/Cu2O/Cu0-based photocatalysts toward CO2 reduction, Chem. Eng. J., 2022, 429, 131579 CrossRef CAS.
- Z. Sun, Y. Hu, D. Zhou, M. Sun, S. Wang and W. Chen, Factors influencing the performance of copper-bearing catalysts in the CO2 reduction system, ACS Energy Lett., 2021, 6, 3992–4022 CrossRef CAS.
- T. Yang, X. Mao, Y. Zhang, X. Wu, L. Wang, M. Chu, C.-W. Pao, S. Yang, Y. Xu and X. Huang, Coordination tailoring of Cu single sites on C3N4 realizes selective
CO2 hydrogenation at low temperature, Nat. Commun., 2021, 12, 6022 CrossRef CAS PubMed.
- J. Wang, T. Heil, B. Zhu, C.-W. Tung, J. Yu, H. M. Chen, M. Antonietti and S. Cao, A single Cu-center containing enzyme-mimic enabling full photosynthesis under CO2 reduction, ACS Nano, 2020, 14, 8584–8593 CrossRef CAS PubMed.
- H. X. Zhang, Q. L. Hong, J. Li, F. Wang, X. Huang, S. Chen, W. Tu, D. Yu, R. Xu, T. Zhou and J. Zhang, Isolated square-planar copper center in boron imidazolate nanocages for photocatalytic reduction of CO2 to CO, Angew. Chem., Int. Ed., 2019, 58, 11752–11756 CrossRef CAS PubMed.
- Y. Yu, X. Dong, P. Chen, Q. Geng, H. Wang, J. Li, Y. Zhou and F. Dong, Synergistic effect of Cu single atoms and Au–Cu alloy nanoparticles on TiO2 for efficient CO2 photoreduction, ACS Nano, 2021, 15, 14453–14464 CrossRef CAS.
- Y. Zhao, Y. Zhao, R. Shi, B. Wang, G. I. N. Waterhouse, L. Z. Wu, C. H. Tung and T. Zhang, Tuning oxygen vacancies in ultrathin TiO2 nanosheets to boost photocatalytic nitrogen fixation up to 700 nm, Adv. Mater., 2019, 31, 1806482 CrossRef.
- Q. Zhao, C. Zhang, R. Hu, Z. Du, J. Gu, Y. Cui, X. Chen, W. Xu, Z. Cheng, S. Li, B. Li, Y. Liu, W. Chen, C. Liu, J. Shang, L. Song and S. Yang, Selective etching quaternary MAX phase toward single atom copper immobilized MXene (Ti3C2Clx) for efficient CO2 electroreduction to methanol, ACS Nano, 2021, 15, 4927–4936 CrossRef CAS PubMed.
- H. Bao, Y. Qiu, X. Peng, J.-A. Wang, Y. Mi, S. Zhao, X. Liu, Y. Liu, R. Cao, L. Zhuo, J. Ren, J. Sun, J. Luo and X. Sun, Isolated copper single sites for high-performance electroreduction of carbon monoxide to multicarbon products, Nat. Commun., 2021, 12, 238 CrossRef CAS PubMed.
- Y. Zhang, J. Zhao, H. Wang, B. Xiao, W. Zhang, X. Zhao, T. Lv, M. Thangamuthu, J. Zhang, Y. Guo, J. Ma, L. Lin, J. Tang, R. Huang and Q. Liu, Single-atom Cu anchored catalysts for photocatalytic renewable H2 production with a quantum efficiency of 56%, Nat. Commun., 2022, 13, 58 CrossRef CAS.
- Y. Zhang, B. Xia, J. Ran, K. Davey and S. Z. Qiao, Atomic-level reactive sites for semiconductor-based photocatalytic CO2 reduction, Adv. Energy Mater., 2020, 10, 1903879 CrossRef CAS.
- F. Wen and W. Liu, Three-dimensional ordered macroporous materials for photocatalysis: Design and applications, J. Mater. Chem. A, 2021, 9, 18129–18147 RSC.
- A.-Y. Lo and F. Taghipour, Ordered mesoporous photocatalysts for CO2 photoreduction, J. Mater. Chem. A, 2021, 9, 26430–26453 RSC.
- X. Z. Zheng, J. Han, Y. Fu, Y. Deng, Y. Y. Liu, Y. Yang, T. Wang and L. W. Zhang, Highly efficient CO2 reduction on ordered porous Cu electrode derived from Cu2O inverse opals, Nano Energy, 2018, 48, 93–100 CrossRef CAS.
- J. Jiao, Y. Wei, Y. Zhao, Z. Zhao, A. Duan, J. Liu, Y. Pang, J. Li, G. Jiang and Y. Wang, AuPd/3DOM-TiO2 catalysts for photocatalytic reduction of CO2: High efficient separation of photogenerated charge carriers, Appl. Catal., B, 2017, 209, 228–239 CrossRef CAS.
- C. Wang, X. Liu, W. He, Y. Zhao, Y. Wei, J. Xiong, J. Liu, J. Li, W. Song, X. Zhang and Z. Zhao, All-solid-state Z-scheme photocatalysts of g-C3N4/Pt/macroporous-(TiO2@carbon) for selective boosting visible-light-driven conversion of CO2 to CH4, J. Catal., 2020, 389, 440–449 CrossRef CAS.
- X. Tao, R. Long, D. Wu, Y. Hu, G. Qiu, Z. Qi, B. Li, R. Jiang and Y. Xiong, Anchoring positively charged Pd single atoms in ordered porous ceria to boost catalytic activity and stability in suzuki coupling reactions, Small, 2020, 16, 2001782 CrossRef CAS PubMed.
- H. S. Jeon, J. Timoshenko, C. Rettenmaier, A. Herzog, A. Yoon, S. W. Chee, S. Oener, U. Hejral, F. T. Haase and B. R. Cuenya, Selectivity control of Cu nanocrystals in a gas-fed flow cell through CO2 pulsed electroreduction, J. Am. Chem. Soc., 2021, 143, 7578–7587 CrossRef CAS.
- J. Fu, K. Jiang, X. Qiu, J. Yu and M. Liu, Product selectivity of photocatalytic CO2 reduction reactions, Mater. Today, 2020, 32, 222–243 CrossRef CAS.
- Y. C. Tan, K. B. Lee, H. Song and J. Oh, Modulating local CO2 concentration as a general strategy for enhancing C–C coupling in CO2 electroreduction, Joule, 2020, 4, 1104–1120 CrossRef CAS.
- X. Wang, P. Ou, J. Wicks, Y. Xie, Y. Wang, J. Li, J. Tam, D. Ren, J. Y. Howe, Z. Wang, A. Ozden, Y. Z. Finfrock, Y. Xu, Y. Li, A. S. Rasouli, K. Bertens, A. H. Ip, M. Graetzel, D. Sinton and E. H. Sargent, Gold-in-copper at low *CO coverage enables efficient electromethanation of CO2, Nat. Commun., 2021, 12, 3387 CrossRef CAS PubMed.
- X. Wang, S. Shen, Z. Feng and C. Li, Time-resolved photoluminescence of anatase/rutile TiO2 phase junction revealing charge separation dynamics, Chin. J. Catal., 2016, 37, 2059–2068 CrossRef CAS.
- Q. Xu, Y. Ma, J. Zhang, X. Wang, Z. Feng and C. Li, Enhancing hydrogen production activity and suppressing CO formation from photocatalytic biomass reforming on Pt/TiO2 by optimizing anatase–rutile phase structure, J. Catal., 2011, 278, 329–335 CrossRef CAS.
- C. Peng, T. Zhou, P. Wei, H. Ai, B. Zhou, H. Pan, W. Xu, J. Jia, K. Zhang, H. Wang and H. Yu, Regulation of the rutile anatase TiO2 phase junction in situ grown on -OH terminated Ti3C2Tx (MXene) towards remarkably enhanced photocatalytic hydrogen evolution, Chem. Eng. J., 2022, 439, 135685 CrossRef CAS.
- B.-R. Chen, V.-H. Nguyen, J. C. S. Wu, R. Martin and K. Kočí, Production of renewable fuels by the photohydrogenation of CO2: effect of the Cu species loaded onto TiO2 photocatalysts, Phys. Chem. Chem. Phys., 2016, 18, 4942–4951 RSC.
- N. Li, B. Wang, Y. Si, F. Xue, J. Zhou, Y. Lu and M. Liu, Toward high-value hydrocarbon generation by photocatalytic reduction of CO2 in water vapor, ACS Catal., 2019, 9, 5590–5602 CrossRef CAS.
- D. Qi, L. Lu, Z. Xi, L. Wang and J. Zhang, Enhanced photocatalytic performance of TiO2 based on synergistic effect of Ti3+ self-doping and slow light effect, Appl. Catal., B, 2014, 160–161, 621–628 CrossRef CAS.
- M. Zhang, W. Liao, Y. Wei, C. Wang, Y. Fu, Y. Gao, L. Zhu, W. Zhu and H. Li, Aerobic oxidative desulfurization by nanoporous tungsten oxide with oxygen defects, ACS Appl. Nano Mater., 2021, 4, 1085–1093 CrossRef CAS.
- S. Wu, Z. Chen, W. Yue, S. Mine, T. Toyao, M. Matsuoka, X. Xi, L. Wang and J. Zhang, Single-atom high-valent Fe(IV) for promoted photocatalytic nitrogen hydrogenation on porous TiO2-SiO2, ACS Catal., 2021, 11, 4362–4371 CrossRef CAS.
- Y. Zhang, C. Cao, X.-T. Wu and Q.-L. Zhu, Three-dimensional porous copper-decorated bismuth-based nanofoam for boosting the electrochemical reduction of CO2 to formate, Inorg. Chem. Front., 2021, 8, 2461–2467 RSC.
- J. A. Benavides, C. P. Trudeau, L. F. Gerlein and S. G. Cloutier, Laser selective photoactivation of amorphous TiO2 films to anatase and/or rutile crystalline phases, ACS Appl. Energy Mater., 2018, 1, 3607–3613 CrossRef CAS.
- Z. Luo, A. S. Poyraz, C.-H. Kuo, R. Miao, Y. Meng, S.-Y. Chen, T. Jiang, C. Wenos and S. L. Suib, Crystalline mixed phase (anatase/rutile) mesoporous titanium dioxides for visible light photocatalytic activity, Chem. Mater., 2014, 27, 6–17 CrossRef.
- Y. Ma, X. Yi, S. Wang, T. Li, B. Tan, C. Chen, T. Majima, E. R. Waclawik, H. Zhu and J. Wang, Selective photocatalytic CO2 reduction in aerobic environment by microporous Pd-porphyrin-based polymers coated hollow TiO2, Nat. Commun., 2022, 13, 1400 CrossRef CAS PubMed.
- S. Jiang, K. Zhao, M. Al-Mamun, Y. L. Zhong, P. Liu, H. Yin, L. Jiang, S. Lowe, J. Qi, R. Yu, D. Wang and H. Zhao, Design of three-dimensional hierarchical TiO2/SrTiO3 heterostructures towards selective CO2 photoreduction, Inorg. Chem. Front., 2019, 6, 1667–1674 RSC.
- T. Wu, X. Zhu, Z. Xing, S. Mou, C. Li, Y. Qiao, Q. Liu, Y. Luo, X. Shi, Y. Zhang and X. Sun, Greatly improving electrochemical N2 reduction over TiO2 nanoparticles by Iron doping, Angew. Chem., Int. Ed., 2019, 58, 18449–18453 CrossRef CAS PubMed.
- J. Y. Do, R. K. Chava, K. K. Mandari, N.-K. Park, H.-J. Ryu, M. W. Seo, D. Lee, T. S. Senthil and M. Kang, Selective methane production from visible-light-driven photocatalytic carbon dioxide reduction using the surface plasmon resonance effect of superfine silver nanoparticles anchored on lithium titanium dioxide nanocubes (Ag@LixTiO2), Appl. Catal., B, 2018, 237, 895–910 CrossRef CAS.
- M. Xiao, L. Zhang, B. Luo, M. Lyu, Z. Wang, H. Huang, S. Wang, A. Du and L. Wang, Molten-salt-mediated synthesis of an atomic nickel Co-catalyst on TiO2 for improved photocatalytic H2 evolution, Angew. Chem., Int. Ed., 2020, 59, 7230–7234 CrossRef CAS PubMed.
- Z. Wang, X. Mao, P. Chen, M. Xiao, S. A. Monny, S. Wang, M. Konarova, A. Du and L. Wang, Understanding the roles of oxygen vacancies in hematite-based photoelectrochemical processes, Angew. Chem., Int. Ed., 2019, 58, 1030–1034 CrossRef CAS PubMed.
- C. Liu, Y. Qin, W. Guo, Y. Shi, Z. Wang, Y. Yu and L. Wu, Visible-light-driven photocatalysis over nano-TiO2 with different morphologies: From morphology through active site to photocatalytic performance, Appl. Surf. Sci., 2022, 580, 152262 CrossRef CAS.
- G. Wang, C.-T. He, R. Huang, J. Mao, D. Wang and Y. Li, Photoinduction of Cu single atoms decorated on UiO-66-NH2 for enhanced photocatalytic reduction of CO2 to liquid fuels, J. Am. Chem. Soc., 2020, 142, 19339–19345 CrossRef CAS PubMed.
- J. Yang, X. Zhu, Z. Mo, J. Yi, J. Yan, J. Deng, Y. Xu, Y. She, J. Qian, H. Xu and H. Li, A multidimensional In2S3–CuInS2 heterostructure for photocatalytic carbon dioxide reduction, Inorg. Chem. Front., 2018, 5, 3163–3169 RSC.
- J. Y. Do, N. Son, R. K. Chava, K. K. Mandari, S. Pandey, V. Kumaravel, T. S. Senthil, S. W. Joo and M. Kang, Plasmon-induced hot electron amplification and effective charge separation by Au nanoparticles sandwiched between copper titanium phosphate nanosheets and improved carbon dioxide conversion to methane, ACS Sustainable Chem. Eng., 2020, 8, 18646–18660 CrossRef CAS.
- A. A. Liu, L. C. Liu, Y. Cao, J. M. Wang, R. Si, F. Gao and L. Dong, Controlling dynamic structural transformation of atomically dispersed CuOx species and influence on their catalytic performances, ACS Catal., 2019, 9, 9840–9851 CrossRef CAS.
- Y. Wang, Z. Chen, P. Han, Y. Du, Z. Gu, X. Xu and G. Zheng, Single-atomic Cu with multiple oxygen vacancies on ceria for electrocatalytic CO2 reduction to CH4, ACS Catal., 2018, 8, 7113–7119 CrossRef CAS.
- B.-H. Lee, S. Park, M. Kim, A. K. Sinha, S. C. Lee, E. Jung, W. J. Chang, K.-S. Lee, J. H. Kim, S.-P. Cho, H. Kim, K. T. Nam and T. Hyeon, Reversible and cooperative photoactivation of single-atom Cu/TiO2 photocatalysts, Nat. Mater., 2019, 18, 620–626 CrossRef CAS PubMed.
- Z. Xu, C. Guo, X. Liu, L. Li, L. Wang, H. Xu, D. Zhang, C. Li, Q. Li and W. Wang, Ag nanoparticles anchored organic/inorganic Z-Scheme 3DOMM-TiO2-based heterojunction for efficient photocatalytic and photoelectrochemical water splitting, Chin. J. Catal., 2022, 43, 1360–1370 CrossRef CAS.
- X. Lv, D. Y. S. Yan, F. L.-Y. Lam, Y. H. Ng, S. Yin and A. K. An, Solvothermal synthesis of copper-doped BiOBr microflowers with enhanced adsorption and visible-light driven photocatalytic degradation of norfloxacin, Chem. Eng. J., 2020, 401, 126012 CrossRef CAS.
- R. K. Chava, J. Y. Do and M. Kang, Enhanced photoexcited carrier separation in CdS–SnS2 heteronanostructures: A new 1D–0D visible-light photocatalytic system for the hydrogen evolution reaction, J. Mater. Chem. A, 2019, 7, 13614–13628 RSC.
- N. Zhang, A. Jalil, D. Wu, S. Chen, Y. Liu, C. Gao, W. Ye, Z. Qi, H. Ju, C. Wang, X. Wu, L. Song, J. Zhu and Y. Xiong, Refining defect states in W18O49 by Mo doping: A strategy for tuning N2 activation towards solar-driven nitrogen fixation, J. Am. Chem. Soc., 2018, 140, 9434–9443 CrossRef CAS PubMed.
- X. Chang, T. Wang and J. Gong, CO2 photo-reduction: Insights into CO2 activation and reaction on surfaces of photocatalysts, Energy Environ. Sci., 2016, 9, 2177–2196 RSC.
- S. Kreft, R. Schoch, J. Schneidewind, J. Rabeah, E. V. Kondratenko, V. A. Kondratenko, H. Junge, M. Bauer, S. Wohlrab and M. Beller, Improving selectivity and activity of CO2 reduction photocatalysts with oxygen, Chem, 2019, 5, 1818–1833 CAS.
- Y. Nosaka and A. Nosaka, Understanding hydroxyl radical (˙OH) generation processes in photocatalysis, ACS Energy Lett., 2016, 1, 356–359 CrossRef CAS.
- Z. Tian, C. Han, Y. Zhao, W. Dai, X. Lian, Y. Wang, Y. Zheng, Y. Shi, X. Pan, Z. Huang, H. Li and W. Chen, Efficient photocatalytic hydrogen peroxide generation coupled with selective benzylamine oxidation over defective ZrS3 nanobelts, Nat. Commun., 2021, 12, 2039 CrossRef CAS PubMed.
- R. Long, Y. Li, Y. Liu, S. Chen, X. Zheng, C. Gao, C. He, N. Chen, Z. Qi, L. Song, J. Jiang, J. Zhu and Y. Xiong, Isolation of Cu atoms in Pd lattice: Forming highly selective sites for photocatalytic conversion of CO2 to CH4, J. Am. Chem. Soc., 2017, 139, 4486–4492 CrossRef CAS PubMed.
- Ş. Neaţu, J. A. Maciá-Agulló, P. Concepción and H. Garcia, Gold–copper nanoalloys supported on TiO2 as photocatalysts for CO2 reduction by water, J. Am. Chem. Soc., 2014, 136, 15969–15976 CrossRef PubMed.
- Y. Huang, K. Li, J. Zhou, J. Guan, F. Zhu, K. Wang, M. Liu, W. Chen and N. Li, Nitrogen-stabilized oxygen vacancies in TiO2 for site-selective loading of Pt and CoOx cocatalysts toward enhanced photoreduction of CO2 to CH4, Chem. Eng. J., 2022, 439, 135744 CrossRef CAS.
- W. He, Y. Wei, J. Xiong, Z. Tang, W. Song, J. Liu and Z. Zhao, Insight into reaction pathways of CO2 photoreduction into CH4 over hollow microsphere Bi2MoO6-supported Au catalysts, Chem. Eng. J., 2022, 433, 133540 CrossRef CAS.
- W. Wang, C. Deng, S. Xie, Y. Li, W. Zhang, H. Sheng, C. Chen and J. Zhao, Photocatalytic C–C coupling from carbon dioxide reduction on copper oxide with mixed-valence copper(I)/copper(II), J. Am. Chem. Soc., 2021, 143, 2984–2993 CrossRef CAS PubMed.
- W. Ma, S. Xie, T. Liu, Q. Fan, J. Ye, F. Sun, Z. Jiang, Q. Zhang, J. Cheng and Y. Wang, Electrocatalytic reduction of CO2 to ethylene and ethanol through hydrogen-assisted C–C coupling over fluorine-modified copper, Nat. Catal., 2020, 3, 478–487 CrossRef CAS.
- B. N. Choi, J. Y. Seo, Z. An, P. J. Yoo and C.-H. Chung, An in situ spectroscopic study on the photochemical CO2 reduction on CsPbBr3 perovskite catalysts embedded in a porous copper scaffold, Chem. Eng. J., 2022, 430, 132807 CrossRef CAS.
- R. Zhang, H. Wang, S. Tang, C. Liu, F. Dong, H. Yue and B. Liang, Photocatalytic oxidative dehydrogenation of ethane using CO2 as a soft oxidant over Pd/TiO2 catalysts to C2H4 and syngas, ACS Catal., 2018, 8, 9280–9286 CrossRef CAS.
- M. Wang, M. Shen, X. Jin, J. Tian, M. Li, Y. Zhou, L. Zhang, Y. Li and J. Shi, Oxygen vacancy generation and stabilization in CeO2−x by Cu introduction with improved CO2 photocatalytic reduction activity, ACS Catal., 2019, 9, 4573–4581 CrossRef CAS.
- J. Li, X. Li, C. M. Gunathunge and M. M. Waegele, Hydrogen bonding steers the product selectivity of electrocatalytic CO reduction, Proc. Natl. Acad. Sci. U. S. A., 2019, 116, 9220–9229 CrossRef CAS PubMed.
- P. B. Joshi, N. Karki and A. J. Wilson, Electrocatalytic CO2 reduction in acetonitrile enhanced by the local environment and mass transport of H2O, ACS Energy Lett., 2022, 7, 602–609 CrossRef CAS.
- J. X. Hao, D. Yang, J. J. Wu, B. X. Ni, L. J. Wei, Q. J. Xu, Y. L. Min and H. X. Li, Utilizing new metal phase nanocomposites deep photocatalytic conversion of CO2 to C2H4, Chem. Eng. J., 2021, 423, 130190 CrossRef CAS.
- L. Hou, J. Han, C. Wang, Y. Zhang, Y. Wang, Z. Bai, Y. Gu, Y. Gao and X. Yan, Ag nanoparticle embedded Cu nanoporous hybrid arrays for the selective electrocatalytic reduction of CO2 towards ethylene, Inorg. Chem. Front., 2020, 7, 2097–2106 RSC.
- H. Huang, R. Shi, Z. Li, J. Zhao, C. Su and T. Zhang, Triphase photocatalytic CO2 reduction over silver-decorated titanium oxide at a gas–water boundary, Angew. Chem., 2022, 61, e202200802 CAS.
|
This journal is © the Partner Organisations 2022 |
Click here to see how this site uses Cookies. View our privacy policy here.