DOI:
10.1039/D3TA07710A
(Paper)
J. Mater. Chem. A, 2024,
12, 9651-9660
MOF-modified C3N4 for efficient photo-induced removal of uranium under air without sacrificial agents†
Received
13th December 2023
, Accepted 11th March 2024
First published on 11th March 2024
Abstract
Photocatalytic reduction of uranium from U(VI) to U(IV) is an effective method to remove uranium from wastewater, while it often requires anaerobic conditions and/or the addition of sacrificial agents, which hinders its further application. Herein, a MOF-modified C3N4 composite material was prepared for uranium removal under air atmosphere without the addition of sacrificial agents, achieving a notable uranium removal capacity of 1355 mg g−1. The introduction of the MOF enhanced the band structure and the photoelectric properties of C3N4, making it able to generate and separate electrons and holes efficiently. The electrons were applied to reduce O2 to form H2O2 and the holes could oxidize H2O to O2. The generated H2O2 could react with UO22+ to form (UO2)O2·2H2O to realize the solidification of uranium under both air and N2 atmospheres. This work may give a new direction to the design of photocatalysts for highly efficient uranium removal under air atmosphere without sacrificial agents.
1 Introduction
As the major nuclear fuel, uranium is a kind of heavy metal with chemical toxicity and radioactive properties. It makes economic and environmental sense to recover and remove uranium from the wastewater generated during its production and use.1–3 Besides, enrichment of uranium from wastewater for recycling is also an effective way to relieve the shortage of uranium resources. In recent years, many methods have been proposed, including extraction, membrane separation, adsorption, precipitation and so on, for the remediation of uranium-containing wastewater.4–8 However, these methods still possess some drawbacks, such as the low removal capacity and complex procedures making uranium enrichment a critical task and a huge challenge.8,9
The photocatalytic method has been widely used to remove uranium due to its high efficiency, ease of operation and designable catalysts.10–13 Normally, uranium could be reduced from soluble U(VI) to insoluble U(IV) with proper photocatalysts such as TiO2, g-C3N4, inorganic oxides, COFs,13–24etc. Although the removal properties of these materials for uranium reduction were remarkable, there are still some problems that restrict its development. Some photocatalytic experiments must be performed under anaerobic conditions, since photogenerated electrons (e−) as well as the newly generated U(IV) are easily reduced by oxygen in water. Moreover, due to the poor separation efficiency of photogenerated electron–hole (e−–h+) pairs on photocatalysts, most systems required the addition of sacrificial agents (methanol, ethanol, etc.) to capture h+ so that the remaining e− can effectively reduce U(VI).18 As a result, maintaining an anaerobic environment and adding sacrificial agents will increase costs and may bring secondary pollution, restricting the further application of the photocatalytic method. Therefore, it is urgently necessary to develop new photocatalysts for the removal of uranium under air atmosphere without the addition of sacrificial agents.11,12,25
In our previous work, we found that uranyl ions can react with photogenerated H2O2 to form metastudtite under air atmosphere, which can achieve efficient solidification and separation of uranyl ions.26 Therefore, the synthesis of photocatalysts that can efficiently produce H2O2 is expected to be an effective way for the removal of uranium. Among various catalysts, carbon nitride (C3N4) is widely used in the field of photocatalysis due to its medium band gap (∼2.7 eV), easy preparation, non-toxicity and low cost.24,25 However, the weak visible light absorption and the easy recombination rate of electrons and holes limit the application of C3N4.27–29 To overcome these deficiencies, a number of methods for modifying C3N4 for the generation of H2O2 have been used, including element doping, structural engineering, and heterojunction structures.30–33 Metal organic frameworks (MOFs) have excellent chemical stability, special electronic structures and suitable band structures to assist C3N4 in the generation of H2O2 driven by visible light.34,35 Shi et al.36 proposed a thermally induced strategy to inhibit the recombination of MOF photogenerated carriers, thereby improving the optoelectronic performance of MOFs. It confirmed the unlimited potential of MOFs in the field of photocatalysis. In addition, the modification with MOFs can not only improve the utilization rate of visible light but also provide additional reaction sites for U(VI). As a result, it is of great significance to study the activity of MOF-modified C3N4 in photocatalytic removal of U(VI).
In this paper, we select Ni3 (HHTP)2, which is a 2D conductive hexagonal MOF with strong light response and excellent electron transfer ability,37 for the modification of C3N4 with a facile method. The morphology, structure and optical absorption properties of the C3N4–MOF composite material were systematically characterized. It was found that the composites could oxidize H2O to generate H2O2 efficiently without the addition of any sacrificial agents under visible light. Then the photocatalytic performance for uranium removal was tested under different conditions and the mechanism was also discussed deeply. This work verifies that the photocatalyst for the generation of H2O2 could be applied in removal of uranium, broadening the way to select photocatalysts and deepening the understanding of photocatalytic uranium removal. It also provides an effective photocatalyst for uranium removal without sacrificial agents under air atmosphere, which presents great advantages for further application.
2 Experimental section
2.1 Chemicals and reagents
Urea (CO(NH2)2) and melamine (C3N3(NH2)3) were obtained from Beijing Chemical Works. Ni(OAc)2·4H2O (nickel(II) acetate tetrahydrate) and HHTP (2,3,6,7,10,11-hexahydroxytriphenylene) were purchased from Aladdin Reagent Co., Ltd (Shanghai, China). Uranyl nitrate hydrate (UO2(NO3)2·6H2O) was bought from a commercial supplier. All reagents were used without further purification.
2.2 Synthetic method
2.2.1 Synthesis of C3N4.
Urea (1.5 g), melamine (1.5 g), NaOH (0.2 g) and KCl (7.5 g) were fully mixed, and then the mixture was put into a tube furnace and heated up to 550 °C for 4 h with a heating rate of 2.5 °C min−1. After the reaction, the sample was washed with deionized water and dried in an oven for further application.
2.2.2 Synthesis of the MOF.
Ni(OAc)2·4H2O (80 mg, 0.320 mmol) and HHTP (56 mg, 0.168 mmol) were dispersed in 32 mL of ultrapure water. After uniform ultrasonic dispersion, the solution mixture was placed in an autoclave and kept at 85 °C for 12 h. After the reaction, the product was obtained by centrifugation and washed with deionized water.
2.2.3 Synthesis of C3N4–MOF.
Ni(OAc)2·4H2O (80 mg, 0.320 mmol), HHTP (56 mg, 0.168 mmol) and C3N4 (0.24 g) were put in 32 mL of ultrapure water. After being fully ultrasound dispersed, the solution mixture was placed in an autoclave and kept at 85 °C for 12 h. After the reaction, the product was obtained by centrifugation and washed with deionized water three times.
2.3 Photocatalytic experiments
In the photochemical experiment, 50 mL of U(VI) solution with a concentration of 0.4 mM was mixed with 30 mg C3N4–MOF (C3N4 and MOF) in a 100 mL reactor. The pH of the mixture was adjusted in the range of 3.0 ± 0.1 to 6.0 ± 0.1 by adding negligible volumes of HNO3 and NaOH solutions (∼0.5 M). The suspension was magnetically stirred in the dark for 1 h, then irradiated under visible light with a 300 W xenon lamp equipped with a UV-cut filter (λ > 420 nm), and the reactor was kept at a constant temperature of 25 °C with a circulation pump. At intervals of 20 min, 1 mL of the suspension was filtered through a 0.22 μm membrane filter after being pipetted through a syringe. The concentration of uranyl ions left in the solution was determined on a spectrophotometer.
3 Results and discussion
3.1 Characterization of the photocatalysts
The morphology and structure of C3N4, MOF and the C3N4–MOF composite material were tested by SEM (Fig. 1a–c). Fig. 1a shows that C3N4 had a typical block structure with rough and wrinkled surfaces, and the MOF in Fig. 1b presented a rod-like morphology with different lengths and a width of about 50 nm. Fig. 1c shows that in C3N4–MOF, the MOF crossed and tightly connected in the structure of C3N4, indicating that the MOF has been composited with C3N4 successfully.
 |
| | Fig. 1 The SEM images of (a) C3N4, (b) MOF and (c) C3N4–MOF. (d) XRD spectra of C3N4, MOF and C3N4–MOF. | |
Then XRD was performed to evaluate the crystallinity of C3N4, MOF and C3N4–MOF (Fig. 1d). It was showed that the MOF has several diffraction peaks at 9.3°, 13.9°, 16.2°, 18.1°, 19.3° and 26.6°, corresponding to the (020), (121), (022), (022), (032) and (004) lattice planes, respectively.37,38 The crystal structure of C3N4–MOF shown in Fig. 1d contains the characteristic peaks of both substances, which further confirms the successful combination of the C3N4 and MOF. The nitrogen adsorption–desorption isotherms of C3N4, MOF and C3N4–MOF are shown in Fig. S1† and the BET surface area of each sample was 66.2 m2 g−1, 302.6 m2 g−1 and 81.1 m2 g−1, respectively.
Furthermore, XPS was performed to analyze the surface chemical state and the elemental composition of different materials. The survey spectrum of materials in Fig. S2a–c† displayed the existence of C, N, O and Ni elements. The atomic ratios of different elements in C3N4, MOF and C3N4–MOF are listed in Table S1.†Fig. 2 present the element spectra of C3N4–MOF. The C 1s spectra (Fig. 2a) displayed three peaks at 288.5 eV, 285.2 eV and 284.4 eV, which can be attributed to the N–C
N, C–N and C–C bonds,39,40 respectively. Two peaks centered at 873.9 eV and 856.1 eV in Ni 2p spectra can be attributed to Ni 2p1/2 and Ni 2p3/2.38,41 The O 1s signal was fitted with two peaks at 532.9 eV and 530.73 eV and they were O–H and Ni–O bonds, respectively.38,41 Furthermore, the peaks at 401.1 eV, 399.9 eV and 398.9 eV of N 1s spectra can be indexed to the free amino group (C–N–H), tertiary nitrogen N–(C)3 and the sp2-hybridized aromatic N bonded to carbon atoms (C
N–C),38,42 respectively. In summary, the morphologies and the chemical structures based on SEM, XRD and XPS characterization indicated the successful synthesis of C3N4–MOF composite materials.
 |
| | Fig. 2 XPS spectra of C3N4–MOF: (a) C 1s, (b) Ni 2p, (c) O 1s and (d) N 1s. | |
3.2 Optical, photoelectrochemical, and electrochemical properties
To evaluate the activity of different materials, the optical, photoelectrochemical and electrochemical properties were investigated. First, the optical properties were demonstrated with the UV-vis absorption spectrum in Fig. 3a. It can be seen that the MOF exhibited excellent light absorption capacity in the visible light range. C3N4 had stronger absorption capacity in the UV-light range than that in the range of 450–800 nm, indicating poor visible light harvesting capabilities.42,43 However C3N4–MOF can capture sunlight in a wider range than C3N4 and the absorption of light greater than 450 nm was mainly contributed by the MOF. The bandgap energy of C3N4 and C3N4–MOF was calculated to be 2.67 eV and 2.28 eV from the UV-vis absorption spectrum and a narrower band gap was obtained after compositing with the MOF (Fig. S3a–c†). Therefore, compared with pure C3N4, C3N4–MOF can obtain more sunlight energy to generate more electron–hole pairs and also to accelerate the separation of electrons and holes.
 |
| | Fig. 3 (a) UV-vis absorption spectra, (b) transient photocurrent responses (TPR), (c) the electrochemical impedance spectroscopy (EIS) and (d) energy band structure of different materials. | |
In order to investigate the interfacial electron transfer rate of the materials, the transient photocurrent responses and electrochemical impedance spectra were recorded. Fig. 3b shows the transient photocurrent response (TPR) curves of materials with on–off cycles of intermittent irradiation. When light illuminated the materials, the photocurrent value enhanced rapidly, indicating that much more efficient separation of charge carriers and interfacial charge transfer were obtained with the materials. Obviously, the C3N4–MOF composite material possessed a stronger photogenerated current than pure C3N4 and MOF, indicating that more surface photogenerated charge existed on it. The interfacial charge transfer of C3N4, MOF, and C3N4–MOF was studied by electrochemical impedance spectroscopy (EIS).44,45 The arc radius represents the internal resistance of charge transfer. The EIS results (Fig. 3c) indicated that C3N4–MOF had much smaller semicircles compared with C3N4 and MOF. It illustrated that C3N4–MOF exhibited a lower resistance, and more efficient charge separation, which was consistent with photocurrent results.
To further investigate the energy band structure of the materials, Mott–Schottky experiments were performed as shown in Fig. S3d–f.† The conduction band (CB) potential of different materials was −0.59 V (C3N4), −0.56 V (MOF) and −0.74 V (C3N4–MOF), which were all more negative than the reduction potential of O2/H2O2 (+0.68 V, vs. RHE). From the band structure in Fig. 3d, the CB energy of C3N4 was more negative than that of the MOF, so the photogenerated electrons induced by C3N4 could be transferred to the CB of the MOF. The reason was that the MOF made a connection between the block structures of C3N4, thereby boosting charge transfer between them. The PL spectra of C3N4 and C3N4–MOF with an excitation wavelength of 380 nm are exhibited in Fig. S4.† It can be seen that C3N4 exhibits strong fluorescence intensity at 425 nm while the fluorescence intensity decreased significantly when combined with the MOF, which further confirmed the rapid transfer of photoelectrons from C3N4 to the MOF. However, the valence band (VB) of MOF was only 1.19 V as calculated, which was not enough to oxidize water. Therefore, the water oxidation process mainly occurs on C3N4. The results above revealed that C3N4–MOF had a suitable band structure to initiate both the oxygen reduction reaction and water oxidation reaction under light irradiation.
3.3 Photocatalytic removal of uranium without sacrificial agents
To evaluate the performance of different materials, the photocatalytic removal properties of U(VI) were investigated with and without sacrificial agents in Fig. 4a and b, respectively. At the adsorption stage, the uranium solution mixed with materials was stirred in the dark for 1 h to reach adsorption equilibrium. And then the visible light was employed for the removal of uranium at the photocatalytic stage. As shown in Fig. 4a and b, about 60% of U(VI) was adsorbed by the MOF and C3N4, whereas the adsorption ratio was only 25% for C3N4–MOF at the adsorption stage. Combining the morphology observation, it was because pure MOF and C3N4 were well distributed that would provide more sites for U(VI) adsorption. When C3N4 was combined with the MOF, C3N4–MOF formed aggregates, reducing the exposure of adsorption sites and leading to a decreased adsorption of U(VI).46 With methanol acting as a sacrificial agent in Fig. 4a, the removal efficiency of U(VI) with C3N4 and C3N4–MOF reached up to 95% after 20 minutes of light exposure. However, the concentration of U(VI) remained basically unchanged with the MOF in the photocatalytic stage, which demonstrated that the MOF possessed little photocatalytic properties for uranium. In the system without sacrificial agents in Fig. 4b, there was a visible difference in the photocatalytic stage and about 90% of U(VI) was removed with C3N4–MOF, but no significant decrease of U(VI) was observed with C3N4 and the MOF. The results above revealed that C3N4–MOF worked still well in the photocatalytic removal of uranium without sacrificial agents.
 |
| | Fig. 4 Photocatalytic removal rate of U(VI) with different materials in the presence of methanol (a) and in the absence of methanol (b). (c) The production rate of H2O2 of materials in the presence or absence of methanol. Under different atmospheric conditions, the generation rate of H2O2 catalyzed by C3N4–MOF (d) photocatalytic removal rate of U(VI) catalyzed by C3N4–MOF (e) and the pseudo-first-order rate constant (k) of U(VI) removal by C3N4–MOF (f). | |
We have previously verified that H2O2 played an important role in the photo-induced removal of uranium and it could react with uranyl ions to form metastudtite to realize the solidification of uranium.26 Herein the potential of different materials in photocatalytic production of H2O2 was demonstrated in Fig. 4c in the presence or absence of sacrificial agents. The concentration of H2O2 was determined by TiSO4 absorbance.47 Due to the rapid electron–hole recombination, C3N4 showed the H2O2 production of 0.096 mmol h−1 g−1 in the absence of methanol. And MOF only showed the H2O2 evolution of 0.022 mmol h−1 g−1. Especially, the production rate of H2O2 catalyzed by C3N4–MOF attained 1.02 mmol h−1 g−1, which was about 10 times that of C3N4. With the addition of methanol as a sacrificial agent, the H2O2 generation rate of all materials was increased, which was consistent with the trend without methanol. Theoretically, H2O2 can be generated from H2O and O2 by semiconductor photocatalysis. The photogenerated valence band holes (VB h+) oxidized H2O to generate O2 (eqn (1)), while conduction band electrons (CB e−) promoted the double-electron reduction of O2 to generate H2O2 (eqn (3)).48,49 The photocatalytic H2O2 generation via oxygen reduction can be improved by the use of methanol because the oxidation of methanol is thermodynamically more favorable than the direct oxidation of water.50 In the presence of methanol, the holes were employed to oxidize H2O (eqn (1)) and methanol (eqn (2)), and the liberated electrons can be used for oxygen reduction (eqn (3)) to generate H2O2.51,52
| | | 2H2O + 2h+ ⇌ O2 + 4H+ + 2e− (+1.23 V vs. RHE) | (1) |
| | | CH3OH + ·OH + 2h+ → CO2 + 5H+ + 3e− | (2) |
| | | O2 + 2H+ + 2e− ⇌ H2O2 (+0.68 V vs. RHE) | (3) |
As mentioned above, H2O2 was obtained by the reduction of O2 and a series of experiments were carried out to identify the role of oxygen in the photocatalytic process. Fig. 4d illustrates the H2O2 production rate of C3N4–MOF under different atmospheric conditions and it was visible that the production rate followed the sequence of O2> air >N2, indicating that O2 was an important participant component in this catalytic process. However, in N2-saturated systems, the production of H2O2 was reduced compared to that of air but still detected, which could be attributed to the oxidation of water. Then the photocatalytic removal activity of uranium is shown in Fig. 4e with C3N4–MOF under different atmospheric conditions. And the removal rate was coincident with the H2O2 production rate.
Then the pseudo-first-order rate constant (k) of UO22+ reduction of C3N4–MOF under different atmospheric conditions is shown in Fig. 4f; the photocatalytic reaction rate constant (k) was denoted by applying a pseudo-first-order reaction:27
The photocatalytic reaction rate constant (k) was fitted and k for the O2 saturated system was 0.038 min−1, which was 5 times higher than that for N2 saturated test system (0.007 min−1). The results confirmed the importance of oxygen in the generation of H2O2 and the photocatalytic removal of uranium and it also verified that C3N4–MOF can reduce O2 to H2O2 by a two-electron reaction without adding any sacrificial agents.
3.4 Comprehensive research on the photocatalytic performance
To evaluate the uranium removal activity of C3N4–MOF, different amounts of C3N4–MOF were first used to remove U(VI) with a concentration of 0.4 mM and pH = 5. As shown in Fig. 5a, it was found that 20 mg of C3N4–MOF removed 85% of U(VI) solution effectively, and when the dosage of the catalyst increased, the efficiency increased significantly.53 It was attributed to the increased likelihood of U(VI) coming in contact with the active site of C3N4–MOF. Under different acid-base conditions, the efficiency of C3N4–MOF removal of U(VI) was investigated, as shown in Fig. 5b. It was clear that after 60 min of irradiation at pH = 5, 95% of U(VI) was eliminated, and the removal performance was quicker in weak acid systems. However, the removal efficiency of U(VI) under strong acid conditions was extremely low, which was due to the adsorption competition between U(VI) and H+ on the surface of the catalyst, as well as the unstable photocatalytic products of uranium under acid conditions, which might have redissolved into the solution.54 Besides, at pH 3–6, the U(VI) ions mainly existed as UO22+, UO2OH+, (UO2)2(OH)22+ and (UO2)3(OH)42+ in the aqueous solution. From the zeta potential results (Fig. S5†), the surface of C3N4–MOF was positively charged at pH = 3, and the electrostatic repulsion between the catalyst surface and U(VI) resulted in the weak adsorption of U(VI). As the pH increased, the surface charge of the catalyst gradually became negative, which could trap electrons on the surface of the catalyst due to the electrostatic attraction, thus greatly improving the removal efficiency of U(VI).50 To assess the application of the C3N4–MOF in uranium removal, the elimination capacity was studied and calculated to be about 1355 mg g−1 without sacrificial agents. The stability and reusability of photocatalysts are important factors affecting their photocatalytic performance. Fig. 5c shows the availability of C3N4–MOF over five photocatalytic cycles. After each photocatalytic reaction cycle, the photocatalyst was separated and collected from the suspension by centrifugation and reacted with 1.0 M HNO3 solutions successively to remove the U(VI) deposits on the surface of C3N4–MOF. The photocatalytic activity of C3N4–MOF did not decrease significantly and the removal rate remained above 80% after five repeated cycles, indicating the good stability and reusability of C3N4–MOF. In order to test the practical application of C3N4–MOF, the effects of Cs, La, Zr, Ce, Cu, Zn, and Ba on the extraction performance of U(VI) were further studied (Fig. 5d). The percentage of U(VI) extracted from solutions containing multiple interfering ions showed a slight decrease compared to the absence of interfering ions, but the selectivity for uranium extraction was still excellent. The above experimental data indicated that the catalyst was feasible for the actual removal of uranium.
 |
| | Fig. 5 Comprehensive research of C3N4–MOF for photocatalytic reduction of U(VI): (a) effect of dosage of the catalyst, (b) effect of initial solution pH, (c) effect of cycle performance, and (d) effect of ion selection performance. | |
3.5 Photocatalytic mechanism
To gain an insight into the interaction mechanism and transformation process of uranium, C3N4–MOF after photoreaction under air and N2 atmosphere was characterized. Under an air atmosphere, the products presented agglomerated short rod-like crystals as shown by the SEM image in Fig. 6a. In addition, the EDS presented the uniform distributions of O and U. In Fig. 6b, the XRD patterns confirmed that U(VI) was converted to (UO2)O2·2H2O. The XPS spectra of C3N4–MOF before and after photocatalytic experiments (Fig. S2c and d†) were recorded to verify the presence of uranium on the surface of the catalyst. It was obvious that a new characteristic peak appeared in the range of 380 eV to 395 eV, which was attributed to U 4f. In Fig. 6c, two peaks at 392.57 eV and 381.72 eV were investigated which corresponded to U 4f5/2 and U 4f7/2 of U(VI),51,55 respectively. Besides, the photoproducts under a N2 atmosphere were also characterized with XRD in Fig. S6.† As mentioned in our previous work and other references,26,56,57 U(VI) preferred to form UO2 under a N2 atmosphere after photoreaction with catalysts. However in this work, it was interesting that U(VI) was transformed into (UO2)O2·2H2O other than UO2 (ref. 58) under a N2 atmosphere, which may be attributed to the photocatalytic generation of H2O2 with C3N4–MOF. In N2-saturated systems, the valence band potential of C3N4–MOF was 1.54 V, which was more positive than the oxidation potential of H2O/O2 (+1.23 V, vs. RHE). Therefore, it can oxidize H2O to generate O2, and then the conduction band electrons can promote the double-electron reduction of O2 to generate H2O2. Large amounts of H2O2 were generated along with the irradiation time, leading to the generation of (UO2)O2·2H2O by the direct reaction of UO22+ and H2O2. This result also demonstrated that U(VI) could be solidified to (UO2)O2·2H2O under both air and N2 atmospheres due to the high generation of H2O2 with C3N4–MOF. Furthermore, the FTIR spectrum of C3N4–MOF before and after the reaction was characterized (Fig. 6d). There were no significant changes in functional groups and chemical configurations on the catalyst, indicating that the catalyst was stable before and after the reaction.
 |
| | Fig. 6 (a) SEM images of C3N4–MOF after the photoreaction and elemental distribution of O and U. (b) XRD patterns of C3N4–MOF after the photocatalytic reaction under air. (c) U 4f XPS spectrum of C3N4–MOF after the reaction, (d) FTIR spectrum of C3N4–MOF before and after the photoreaction. | |
Based on the above discussion, we proposed a possible mechanism for the photocatalysis of U(VI) by C3N4–MOF. The introduction of MOF narrowed the band gap of C3N4 and broadened the visible light response area, thereby improving the utilization efficiency of solar energy. Under visible-light irradiation, C3N4–MOF was excited by visible light, and the photogenerated charge was rapidly separated at the surface. The transfer of photoelectrons from C3N4 to the MOF was accelerated by chemical bonding, thereby effectively inhibiting the recombination of photogenerated electron–hole pairs. Due to the influx of extraneous electrons, there would be more free electrons on the MOF that were available for further oxygen reduction reactions. At the same time, the electron–hole separation efficiency of C3N4 was improved and more holes were employed to oxidize water. H2O2 was also produced by C3N4–MOF through the conversion of the dissolving O2 to O2˙− radicals under visible light irradiation (eqn (5)). Large amounts of H2O2 were generated with the further increase of irradiation time, leading to the generation of (UO2)O2·2H2O by the direct reaction of UO22+ and H2O2 (eqn (6)).55
| | | UO22+ + 2H2O + H2O2 → (UO2)O2·2H2O (s) + 2H+ | (6) |
4 Conclusions
In conclusion, a novel C3N4–MOF composite material was synthesized by a hydrothermal reaction. Its morphology, structural and photoelectric properties were characterized and the results indicated the successful synthesis of C3N4–MOF. It has excellent optical adsorption capacity and a narrower band gap compared with C3N4, which was attributed to the introduction of the MOF. The electrons and holes on C3N4–MOF were generated and separated efficiently under visible light, where O2 could be reduced to H2O2 with electrons and holes could oxidize H2O to O2. The production rate of H2O2 was 10 times higher than that of C3N4. In the uranium removal experiments, C3N4–MOF presented satisfactory photocatalytic properties without adding any sacrificial agents under air. In weakly acidic uranium solution, more catalyst and a higher concentration of uranium were beneficial for a faster removal rate. A high uranium elimination capacity of 1355 mg g−1 was achieved with C3N4–MOF and it also displayed good stability in the recycle tests and excellent selectivity in the presence of other metal ions. The mechanism was studied deeply with SEM, XRD, XPS, and FT-IR methods and the results revealed that the product of uranium was (UO2)O2·2H2O under both air and N2 atmospheres, which was formed by the reaction of uranyl ions and H2O2. This work provided a novel photocatalyst for uranium removal without any sacrificial agents under an air atmosphere, promoting the further application of the photocatalytic method. It also proved that the photocatalysts that were used to produce H2O2 were also effective for uranium removal, broadening the selection of catalysts and also providing a new photocatalytic way to remove uranium.
Author contributions
Lingyu Zhang: conceptualization, methodology, data curation, writing original draft. Yuhao Yang: methodology, investigation, data curation. Nan Zhao: validation, methodology. Shuang Liu: writing review & editing. Zhe Wang: validation, methodology, funding acquisition. Xiangke Wang: resources, writing review & editing. Yuexiang Lu: resources, supervision.
Conflicts of interest
The authors declare no competing financial interest.
Acknowledgements
This work was financially supported by the National Natural Science Foundation of China (Grant No. 22376059 and 21976104), Young Elite Scientists Sponsorship Program (2021QNRC001) of the China Association for Science and Technology, and Brain Pool program funded by the Ministry of Science and ICT through the National Research Foundation of Korea (grant number: NRF-2022H1D3A2A02051548).
References
- Z. Xiao-teng, J. Dong-mei, X. Yi-qun, C. Jun-chang, H. Shuai and X. Liang-shu, J. Chem., 2019, 2019, 1–10 CrossRef.
- S. Abdi, M. Nasiri, A. Mesbahi and M. H. Khani, J. Hazard. Mater., 2017, 332, 132–139 CrossRef CAS PubMed.
- F. Yu, Z. Zhu, S. Wang, Y. Peng, Z. Xu, Y. Tao, J. Xiong, Q. Fan and F. Luo, Chem. Eng. J., 2021, 412, 127558 CrossRef CAS.
- X. Zhang, X. Liu, Y. Peng, X. Wu, Y. Tan, Q. Zeng, Z. Song and M. Li, Sep. Purif. Technol., 2022, 287, 120550 CrossRef CAS.
- Y. Xu, J. Yu, J. Zhu, Q. Liu, H. Zhang, J. Liu, R. Chen, Y. Li and J. Wang, Appl. Catal., B, 2022, 316, 121677 CrossRef CAS.
- F. Chen, B. Fan, C. Wang, J. Qian, B. Wang, X. Tang, Z. Qin, Y. Chen, L. Bin, W. Liu, A. Wang, Y. Ye and Y. Wang, J. Hazard. Mater., 2022, 439, 129622 CrossRef CAS PubMed.
- Y. Yuan, S. Feng, L. Feng, Q. Yu, T. Liu and N. Wang, Angew. Chem., Int. Ed., 2020, 59, 4262–4268 CrossRef CAS PubMed.
- D.-m. Fu, M. Li, Y.-l. Hua, F.-y. Gao, X.-y. Wu, X.-w. Zhang, Q. Fang, L. Bi and T. Cai, Sep. Purif. Technol., 2022, 291, 120957 CrossRef CAS.
- S. Prusty, P. Somu, J. K. Sahoo, D. Panda, S. K. Sahoo, S. K. Sahoo, Y. R. Lee, T. Jarin, L. S. Sundar and K. S. Rao, Chemosphere, 2022, 308, 136278 CrossRef CAS PubMed.
- J. Zhao, C. Lyu, R. Zhang, Y. Han, Y. Wu and X. Wu, J. Hazard. Mater., 2023, 442, 130018 CrossRef CAS PubMed.
- Y. Ye, J. Jin, F. Chen, D. D. Dionysiou, Y. Feng, B. Liang, H.-Y. Cheng, Z. Qin, X. Tang, H. Li, D. Yntema, C. Li, Y. Chen and Y. Wang, Chem. Eng. J., 2022, 450, 138317 CrossRef CAS.
- T. Chen, K. Yu, C. Dong, X. Yuan, X. Gong, J. Lian, X. Cao, M. Li, L. Zhou, B. Hu, R. He, W. Zhu and X. Wang, Coord. Chem. Rev., 2022, 467, 214615 CrossRef CAS.
- S. Zhang, L. Chen, Z. Qu, F. Zhai, X. Yin, D. Zhang, Y. Shen, H. Li, W. Liu, S. Mei, G. Ji, C. Zhang, X. Dai, Z. Chai and S. Wang, Chem, 2023, 9, 3172–3184 CAS.
- M. Chen, T. Liu, S. Tang, T. Wei, A. Gu, R. Zhang, Y. Liu, H. Wang, Z. Xie, Y. Yuan, Z. Li and N. Wang, Chem. Eng. J., 2022, 446, 137264 CrossRef CAS.
- Z. Dai, S. Zhao, J. Lian, L. Li and D. Ding, Sep. Purif. Technol., 2022, 298, 121590 CrossRef CAS.
- Z. Zhou, Q. Liu, J. Zhu, J. Liu, H. Zhang, J. Yu, R. Chen, Y. Li and J. Wang, J. Colloid Interface Sci., 2022, 628, 840–848 CrossRef CAS PubMed.
- X. Liu, X. Wang, W. Jiang, C.-R. Zhang, L. Zhang, R.-P. Liang and J.-D. Qiu, Chem. Eng. J., 2022, 450, 138062 CrossRef CAS.
- P. He, L. Zhang, L. Wu, S. Xiao, X. Ren, R. He, X. Yang, R. Liu and T. Duan, Appl. Catal., B, 2023, 322, 122087 CrossRef CAS.
- Y. Guo, S. Li, F. Yang, C. Li, Y. Guo, K. Xuan, G. Wang, Y. Liu and J. Li, J. Hazard. Mater., 2022, 440, 129734 CrossRef CAS PubMed.
- Z. Dai, Y. Zhen, Y. Sun, L. Li and D. Ding, Chem. Eng. J., 2021, 415, 129002 CrossRef CAS.
- Y. Song, A. Li, P. Li, L. He, D. Xu, F. Wu, F. Zhai, Y. Wu, K. Hu, S. Wang and M. V. Sheridan, Chem. Mater., 2022, 34, 2771–2778 CrossRef CAS.
- H. Li, F. Zhai, D. Gui, X. Wang, C. Wu, D. Zhang, X. Dai, H. Deng, X. Su, J. Diwu, Z. Lin, Z. Chai and S. Wang, Appl. Catal., B, 2019, 254, 47–54 CrossRef CAS.
- H. Deng, L. Yuan, Z. Li, D. Wang, X. Wang, P. Liang, L. Wang, Y. Liu, Y. Fu, Z. Chang, Z. Chai, J. K. Gibson and W. Shi, Adv. Sustainable Syst., 2022, 6, 2200316 CrossRef CAS.
- Q. Guo, C. Y. Zhou, Z. B. Ma and X. M. Yang, Adv. Mater., 2019, 31, 1901997 CrossRef CAS PubMed.
- P. L. Liang, L. Y. Yuan, K. Du, L. Wang, Z. J. Li, H. Deng, X. C. Wang, S. Z. Luo and W. Q. Shi, Chem. Eng. J., 2021, 420, 129831 CrossRef CAS.
- Z. Wang, B. Li, H. Shang, X. Dong, L. Huang, Q. Qing, C. Xu, J. Chen, H. Liu, X. Wang, X. Xiong and Y. Lu, Green Chem., 2022, 24, 7092 RSC.
- C. Liu, Z. Dong, C. Yu, J. Gong, Y. Wang, Z. Zhang and Y. Liu, Appl. Surf. Sci., 2021, 537, 147754 CrossRef CAS.
- S. Li, D. Pan, Z. Cui, Y. Xu, H. Shang, W. Hua, F. Wu and W. Wu, Sep. Purif. Technol., 2022, 301, 121966 CrossRef CAS.
- H. Deng, X. C. Wang, L. Wang, Z. J. Li, P. L. Liang, J. Z. Ou, K. Liu, L. Y. Yuan, Z. Y. Jiang, L. R. Zheng, Z. F. Chai and W. Q. Shi, Chem. Eng. J., 2020, 401, 125977 CrossRef CAS.
- J. Gong, Z. Xie, B. Wang, Z. Li, Y. Zhu, J. Xue and Z. Le, J. Environ., 2021, 9, 104638 CAS.
- Z. Zhang, C. Liu, Z. Dong, Y. Dai, G. Xiong, Y. Liu, Y. Wang, Y. Wang and Y. Liu, Appl. Surf. Sci., 2020, 520, 146352 CrossRef CAS.
- Z. Le, C. Xiong, J. Gong, X. Wu, T. Pan, Z. Chen and Z. Xie, Environ. Pollut., 2020, 260, 114070 CrossRef CAS PubMed.
- J. Zhu, X. Hao, Q. Liu, J. Liu, R. Chen, J. Yu, R. Li, P. Liu and J. Wang, J. Mol. Liq., 2021, 343, 117238 CrossRef CAS.
- T. He, S. Chen, B. Ni, Y. Gong, Z. Wu, L. Song, L. Gu, W. Hu and X. Wang, Angew. Chem., Int. Ed., 2018, 57, 3493–3498 CrossRef CAS PubMed.
- H. Li and S. Wang, Chem, 2021, 7, 279–280 CAS.
- Z. W. Huang, K. Q. Hu, X. B. Li, Z. N. Bin, Q. Y. Wu, Z. H. Zhang, Z. J. Guo, W. S. Wu, Z. F. Chai, L. Mei and W. Q. Shi, J. Am. Chem. Soc., 2023, 145, 18148–18159 CrossRef CAS PubMed.
- X. Shi, R. Hua, Y. Xu, T. Liu and G. Lu, Sustainable Energy Fuels, 2020, 4, 4589–4597 RSC.
- H. Yoon, S. Lee, S. Oh, H. Park, S. Choi and M. Oh, Small, 2019, 15, e1805232 CrossRef PubMed.
- Y. Zhou, Y. Sun, C. Zhu, Y. Liu, X. Dai, J. Zhong, Q. Chen, H. Tian, R. Zhou and Z. Kang, J. Mater., 2018, 6, 8955–8961 CAS.
- Q. Liang, H. Jin, Z. Wang, Y. Xiong, S. Yuan, X. Zeng, D. He and S. Mu, Nano Energy, 2019, 57, 746–752 CrossRef CAS.
- Y. Li, Q. Li, S. Zhao, C. Chen, J. Zhou, K. Tao and L. Han, ChemistrySelect, 2018, 3, 13596–13602 CrossRef CAS.
- X. Li, B. Kang, F. Dong, Z. Zhang, X. Luo, L. Han, J. Huang, Z. Feng, Z. Chen, J. Xu, B. Peng and Z. L. Wang, Nano Energy, 2021, 81, 105671 CrossRef CAS.
- F. Guo, W. Shi, C. Zhu, H. Li and Z. Kang, Appl. Catal., B, 2018, 226, 412–420 CrossRef CAS.
- C. Zhu, C. a. Liu, Y. Fu, J. Gao, H. Huang, Y. Liu and Z. Kang, Appl. Catal., B, 2019, 242, 178–185 CrossRef CAS.
- X. Chen, Q. Li, J. Li, J. Chen and H. Jia, Appl. Catal., B, 2020, 270, 118915 CrossRef CAS.
- P. Li, J. Wang, Y. Wang, L. Dong, W. Wang, R. Geng, Z. Ding, D. Luo, D. Pan, J. Liang and Q. Fan, Chem. Eng. J., 2021, 425, 131552 CrossRef CAS.
- S. Meiyazhagan, S. Yugeswaran, P. V. Ananthapadmanabhan, P. R. Sreedevi and K. Suresh, Plasma Chem. Plasma Process., 2020, 40, 1267–1290 CrossRef CAS.
- Y. Shiraishi, T. Takii, T. Hagi, S. Mori, Y. Kofuji, Y. Kitagawa, S. Tanaka, S. Ichikawa and T. Hirai, Nat. Mater., 2019, 18, 985–993 CrossRef CAS PubMed.
- C. a. Liu, Y. Fu, J. Zhao, H. Wang, H. Huang, Y. Liu, Y. Dou, M. Shao and Z. Kang, Chem. Eng. J., 2019, 358, 134–142 CrossRef CAS.
- G. Ding, W. Wang, T. Jiang, B. Han, H. Fan and G. Yang, ChemCatChem, 2013, 5, 192–200 CrossRef CAS.
- X.-H. Jiang, Q.-J. Xing, X.-B. Luo, F. Li, J.-P. Zou, S.-S. Liu, X. Li and X.-K. Wang, Appl. Catal., B, 2018, 228, 29–38 CrossRef CAS.
- S.-Y. Chen, J.-P. Yu, Z.-F. Chai, W.-Q. Shi and L.-Y. Yuan, Chem. Eng. J., 2023, 460, 141742 CrossRef CAS.
- P. Li, J. Wang, Y. Wang, J. Liang, B. He, D. Pan, Q. Fan and X. Wang, Chem. Eng. J., 2019, 365, 231–241 CrossRef CAS.
- H. Asadzadeh Patehkhor, M. Fattahi and M. Khosravi-Nikou, Sci. Rep., 2021, 11, 24177 CrossRef CAS PubMed.
- H. Wang, P. Mei, X. Huang, J. Xiao and Y. Sun, J. Cleaner Prod., 2022, 330, 129821 CrossRef CAS.
- S. Li, Y. Hu, Z. Shen, Y. Cai, Z. Ji, X. Tan, Z. Liu, G. Zhao, S. Hu and X. Wang, Sci. China: Chem., 2021, 64, 1323–1331 CrossRef CAS.
- Z. Cui, S. Li, X. Wei, J. Wang, Y. Xu, M. Zhao, D. Pan and W. Wu, J. Radioanal. Nucl. Chem., 2022, 331, 4159–4168 CrossRef CAS.
- H. Zhang, W. Liu, A. Li, D. Zhang, X. Li, F. Zhai, L. Chen, L. Chen, Y. Wang and S. Wang, Angew. Chem., Int. Ed., 2019, 58, 16110–16114 CrossRef CAS PubMed.
Footnote |
| † Electronic supplementary information (ESI) available: Figures for nitrogen adsorption–desorption isotherms, XPS survey spectrum, and band gaps calculated from the UV-vis absorption spectrum of different materials. See DOI: https://doi.org/10.1039/d3ta07710a |
|
| This journal is © The Royal Society of Chemistry 2024 |
Click here to see how this site uses Cookies. View our privacy policy here.