Open Access Article
Zhimin
Dong‡
a,
Dongling
Zeng‡
a,
Zifan
Li
a,
Junjie
Chen
a,
Youqun
Wang
a,
Xiaohong
Cao
a,
Guoping
Yang
*a,
Zhibin
Zhang
*a,
Yunhai
Liu
a and
Feng
Yang
*b
aState Key Laboratory of Nuclear Resources and Environment, East China University of Technology, Nanchang 330013, China. E-mail: erick@ecut.edu.cn; zhbzhang@ecut.edu.cn
bDepartment of Chemistry, Southern University of Science and Technology, Shenzhen 518055, China. E-mail: yangf3@sustech.edu.cn
First published on 22nd October 2024
Recycling uranium (U) via adsorption and controlled conversion is crucial for the sustainable development of nuclear energy, in which photocatalytic reduction of U(VI) from aqueous solutions is considered one of the most effective strategies. The primary challenge in the photocatalytic elimination of U(VI) resides in the demand for photocatalysts with exceptional properties for effective U(VI) adsorption and charge separation. Herein, we developed the hybrids of polyoxometalate@Cu-metal–organic frameworks (POM@Cu-MOFs) through a self-assembly strategy and demonstrated the efficient removal of U(VI) via synergistic adsorption and photocatalysis. The abundant oxygen-rich groups in POM served as the adsorption sites, endowing POM@Cu-MOFs with a remarkable removal capacity (1987.4 mg g−1 under light irradiation) to remove 99.4% of UO22+. The attraction of electrons from Cu atoms within Cu-MOFs effectively accelerated the carrier dynamics due to their pronounced electronegativity. A mechanism associated with the synergetic effects of adsorption and photocatalytic reduction of U(VI) was proposed. This work provides a feasible approach for efficiently eliminating U(VI) from aqueous solutions in environmental pollution cleanup using the POM@Cu-MOF photocatalyst.
The surface reaction plays a crucial role in the photocatalytic process. This involves the target substance adhering to the material's surface, and subsequently interacting with electrons or holes generated from photoexcitation, leading to the oxidation or reduction process. To improve both catalytic activity and selectivity, various surface engineering strategies have been employed, including the construction of an active surface with high adsorption and activation capacity, the control of surface properties to ensure that more carriers can reach the surface-active site to participate in the reaction, and the improvement of the electron or hole reduction or oxidation capacity of the catalyst on its surface.10–12
Polyoxometalate (POM), a metal oxygen cluster with an oxygen-enriched surface, exhibits rapid and reversible multielectron-transfer reactions while maintaining a stable structure.13 These remarkable attributes endow POMs with excellent photocatalytic reduction and adsorption capacities, facilitating the segregation and immobilization of radioactive waste.14 However, the performance and stability of pure POM may be limited by agglomeration, disordered arrangements, and a restricted surface area.15 Therefore, selecting an appropriate solid matrix for dispersing POM would be a promising alternative strategy. Recently, MOFs have emerged as excellent platforms for various applications due to their unique porous structures and chemical stability, thus showing great promise for applications in lithium-ion batteries, gas separation, luminescent sensors, supercapacitors, photocatalysis, and ion capture.16–21 For instance, Wang et al. reported POM-based MOFs with enhanced U(VI) removal efficiency by three distinct sorption mechanisms (complexation, chemical reduction, and photocatalytic reduction).22 The study on the super-sodalite cage constructed from POM reported by Xu et al. also showed effective U(VI) adsorption capacity.23 Therefore, incorporating POM and MOFs with diverse sorption mechanisms would be a feasible approach for constructing effective sorbents that are still rarely explored.
Here, we developed a feasible self-assembly strategy to construct POM@Cu-MOFs, in which redox POM units are uniformly confined in the pores of MOFs that are constructed through the in situ growth of Cu2(CO2)4 paddle wheel-based structures. Serving as both the photocatalyst and adsorbent for efficient U(VI) removal, POM@Cu-MOFs exhibit synergetic advantages by combining the photocatalysis activity of POM with the strong pre-enrichment capability of MOFs, thereby enhancing the uranium removal efficiency.
C and the symmetrical stretching of the C
O bond.26 All the above characteristic bands of Cu-MOFs are also observed in the FT-IR spectrum of POM@Cu-MOFs, indicating that the framework of Cu-MOFs is preserved. Besides, the emerging bands at 600–1100 cm−1 in the FT-IR spectrum of POM@Cu-MOFs correspond to P–O, Mo
O, and Mo–O–Mo bonds of POM, demonstrating the confined assembly of POM into the Cu-MOF (Fig. 1c and S2†).27
The N2 adsorption–desorption isotherms and pore size distribution were obtained to demonstrate the encapsulation of POM into Cu-MOFs (Fig. 1d and e). The Cu-MOFs exhibits the typical type I isotherm,28,29 indicating the microporous adsorption behaviour. However, the POM@Cu-MOFs exhibit a certain adsorption capacity in the low-pressure region and hysteresis loop in the medium–high pressure region, revealing the appearance of increased microporosity in the POM@Cu-MOFs attributed to the encapsulation of POM.30,31 The substantially decreased specific surface area of POM@Cu-MOFs (872.5 m2 g−1) compared with Cu-MOFs (1365.7 m2 g−1), calculated by using the Brunauer–Emmett–Teller (BET) method, also suggests the encapsulation of POMs into the Cu-MOF framework. The average pore size of Cu-MOFs calculated by the Barrett–Joyner–Halenda (BJH) method was found to be approximately 3.5 nm, which decreased to 3.1 nm after the formation of POM@Cu-MOFs.
The XPS spectra of POM@Cu-MOFs further demonstrate the surface compositions and valence states (Fig. S3†). The signals of C 1s for Cu-MOFs and POM@Cu-MOFs both exhibit three peaks at 288.2, 285.8, and 284.3 eV, which could be assigned to O–C
O, C–O, and C–C (Fig. 2a), respectively.32 The O 1s spectra of Cu-MOFs and POM@Cu-MOFs both show four peaks related to Cu–O, C–O, O–C
O, and O–H (Fig. 2b).33 As shown in Fig. 2c, the Cu element in Cu-MOFs with strong spectral peaks at 954.8 eV and 935.0 eV correspond to Cu 2p1/2 and Cu 2p3/2, respectively.34,35 For POM@Cu-MOFs, the peaks at 133.8 and 134.7 eV, and 235.6 and 232.5 eV are attributed to P 2p3/2 and 2p1/2, and Mo 3d5/2 and 3d3/2, respectively (Fig. 2d and e).36 Compared with the spectra of the Cu-MOFs, the POM@Cu-MOFs showed increased binding energies of Cu 2p. It is worth noting that Cu 2p in POM@Cu-MOFs shows an upshift toward high binding energy compared with Cu-MOFs, which indicates the electron transfer from the Cu-MOF to POM (Fig. 2f).
When the initial concentration of UO22+ is 100 and 200 mg L−1 and the solid–liquid ratio is 0.1 g L−1, the saturated adsorption capacity under dark conditions is measured to be 709.6 and 1637.4 mg g−1, respectively (Fig. 3a and S5c†). In contrast, POM@Cu-MOFs still show high removal capacity even at high initial concentrations under light conditions, achieving 985.7 and 1987.4 mg g−1, respectively. POM@Cu-MOFs offer notable advantages over Cu-MOFs owing to the synergistic effect of light-induced adsorption and photocatalysis. Compared with the photocatalysts reported elsewhere, POM@Cu-MOFs display superior removal capacity (Table S1†),14,22,37–40 highlighting their promising prospects as an effective photocatalyst for uranium-bearing wastewater treatment. We fitted the data by using the pseudo-first-order and the pseudo-second-order models,41,42 respectively. As shown in Fig. 3b and S6,† the R2 factor for the quasi-first-order dynamics (R2 = 0.96) is superior to that of the quasi-second-order dynamics (R2 = 0.95), indicating that the pseudo-first-order model provides a more precise depiction of the reaction kinetics. These results affirm that the rate-limiting step for the removal of U(VI) is attributed to the photocatalytic reduction reaction. Therefore, the enhanced adsorption performance of the POM@Cu-MOFs is more conducive to enhancing photocatalytic reduction performance for U(VI). Then the reaction rate constant (k) was calculated by using the equation −ln(Ce/C0) = kt, where C0 and Ce represent the initial and final concentrations of U(VI), respectively. Under light irradiation, the k values for U(VI) photoreduction by Cu-MOFs and POM@Cu-MOFs are found to be 0.006 and 0.061 min−1, respectively. Under light irradiation, the highest k value for POM@Cu-MOFs is 2.9 times larger than that under dark conditions (Fig. 3b).
The pseudo-first-order kinetics derived from the linear correlation between −ln(Ce/C0) and reaction time (t) (Fig. S7†) indicates that adsorption of UO22+ on the catalyst surface is the rate-determining step in terms of the Langmuir–Hinshelwood model.28 At different initial concentrations of UO22+, POM@Cu-MOFs remove 99.4% of UO22+, which is higher than that removed by Cu-MOFs. The photocatalytic reduction rate constant of POM@Cu-MOFs is 11–20 times larger than that of Cu-MOFs at a UO22+ concentration of 20–200 mg L−1, indicating the higher photocatalytic efficiency of POM@Cu-MOFs.
Additionally, the existence of uranium-containing wastewater containing numerous coexisting ions may potentially affect photocatalytic reduction. With particular emphasis on Ca2+ as a major competing ion, we conducted experiments using various concentrations of CaCl2 in the range of 0.001–1 mol L−1 to evaluate its influence on the photocatalysis efficiency. The removal efficiency of uranium by photocatalytic reduction exceeds 80% across all concentration conditions (Fig. S8†). Therefore, the presence of CaCl2 does not significantly impact the photocatalytic reduction efficiency.
Fig. 3c illustrates the impact of pH on the adsorption and photocatalytic reactions of U(VI) solution. The removal efficiency decreases from 86.3% to 31.8% on increasing the pH value from 3.0 to 5.5 without light irradiation, which increases first and then decreases under light irradiation. Therefore, the optimal pH value for photocatalytic reduction of U(VI) is 4.0. The surface potential of the material becomes more negative with increasing pH value (Fig. S9†). This surface charge, being negative in nature, is more conducive to the adsorption of positively charged UO22+ ions due to the electrostatic interactions.43 Therefore, the surface of the material carries negative charges, and the electrostatic attraction between the negatively charged surface and positively charged UO22+ ions plays a significant role in the photocatalytic removal of U(VI). The evaluation of a photocatalyst’s in treatment efficiency must consider not just its U(VI) removal kinetics in a complex environment but also its recyclability and stability during catalyst development.44 Essentially undiminished photocatalytic reduction efficiency of U(VI) (90%) is achieved during 5 reaction cycles using one-batch POM@Cu-MOFs recovered after each run of the evaluation (Fig. 3d), indicating that POM@Cu-MOFs have excellent structural stability and good cyclic reusability.
| (αhν)1/n = A(hν − Eg) | (1) |
The bandgaps of Cu-MOFs, POM, and POM@Cu-MOFs are calculated to be 3.02, 2.25, and 2.31 eV, respectively (Fig. S10a†). From the Mott–Schottky curves, the positive slopes of the plots suggest that Cu-MOFs are a n-type semiconductor (Fig. 4b).46 From the intercept on the abscissa, the obtained flat band potential (Efb) is −1.27 eV (vs. Ag/AgCl), with the conduction band nearly overlapping the flat band potential, and the position of Efb relative to the normal hydrogen electrode (NHE) is 0.2 higher than the flat band potential;47 thus, the conduction potential (ECB) of Cu-MOFs is calculated to be −1.01 V vs. NHE. Moreover, the flat band potential of POM is around −0.28 V vs. NHE.15 These data, combined with the band gap energies, allow us to calculate the valence band (EVB) edges of Cu-MOFs and POM at 2.01 eV and 1.97 eV (vs. NHE), respectively. The band structure of the POM@Cu-MOFs, along with the ECB value of Cu-MOFs and POM is more negative than the U(VI) reduction potential (UO22+/U4+, +0.267 V) and (UO22+/UO2, +0.411 V). Therefore, the conduction band of POM@Cu-MOFs can provide sufficient driving force for photocatalytic reduction from U(VI) to U(IV).
Additionally, electrochemical impedance spectroscopy (EIS) was performed to further probe the properties of charge transport of POM@Cu-MOF heterostructures. POM@Cu-MOFs exhibited a smaller radius in its semicircle compared to Cu-MOFs, indicating a lower charge transfer resistance and higher interfacial charge separation efficiency (Fig. S10b†). Photoluminescence spectroscopy is widely used to evaluate the charge recombination probability, in which the low spectral intensity implies a lower recombination rate of photogenerated electrons and holes, while higher spectral intensity indicates higher recombination efficiency of photogenerated electrons and holes.48 The PL spectra show that the emission intensity of POM@Cu-MOFs is reduced compared to that of Cu-MOFs (Fig. S10c†). Therefore, the incorporation of POM and MOFs significantly inhibits the recombination rate of electron–hole pairs, highlighting the advantages of POM in improving carrier transfer for enhanced visible photocatalytic activity.
The primary active species involved in the photocatalytic reduction mechanism were identified through a series of trapping experiments. The results of the quenching experiments were obtained by adding 0.01 mol L−1 radical scavenger (Fig. S11†). Interestingly, the presence of methanol enhanced the photocatalytic performance when the electronic scavenger (AgNO3), hole scavenger (methanol), ˙O2− scavenger (p-benzoquinone, BQ), and hydroxyl radical scavenger (t-butylalcohol, TBA) were added to the reaction solution, respectively.49,50 Addition of AgNO3 could inhibit photocatalytic reduction, indicating that electrons are indeed active species involved in the reaction process. In contrast, the presence of methanol can greatly enhance the photocatalytic performance, enabling complete degradation of U(VI) within 40 min. Furthermore, both TBA and BQ exhibit distinct effects on the removal rate of U(VI) under visible light irradiation, achieving 97% and 89% removal rates for U(VI), respectively. During the photocatalytic process, h+, ˙OH, and ˙O2− species are generally employed as oxidizing agents. Thus, scavenging these oxidative radicals and holes can efficiently prevent re-oxidation of U(VI) and reduce recombination between photogenerated holes and electrons to improve the recovery ability, indicating that e− are the main active species. To further verify the details of the photoelectrons in the as-prepared POM@Cu-MOFs, electron paramagnetic resonance (EPR) was performed. As shown in Fig. 4c, the Cu-MOFs exhibit a weak EPR signal with a g factor of 2.003, suggesting a relatively low photoelectron density.51 Conversely, POM@Cu-MOFs exhibit a much stronger EPR signal due to the increased concentration of photogenerated electrons in the system. This is attributed to the formation of a heterojunction between POM and Cu-MOFs, where photogenerated electrons can transfer freely between POM and Cu-MOFs. We performed the EPR experiment by using 5,5-dimethyl-1-pyrroline N-oxide (DMPO) as a free radical trapping agent to further confirm the presence of ˙O2− and ˙OH radicals. Under dark conditions, no EPR signal was observed in an aqueous suspension containing POM@Cu-MOFs. By contrast, the characteristic EPR signals of DMPO–˙O2− and DMPO–˙OH species were observed after light irradiation, and the intensity of these signals increased with the duration of illumination (Fig. S12†). This observation suggests that the photocatalytic removal of U(VI) by POM@Cu-MOFs is facilitated by the generation of more photogenerated electrons and the dissociation of more O2 and H2O on the active sites of the catalyst under illumination.
In addition, temperature-dependent PL spectroscopy was employed to reveal exciton dissociation kinetics in Cu-MOFs and POM@Cu-MOFs (Fig. 4d and e). As the temperature decreases, the fluorescence intensity gradually increases along with enhanced charge carrier recombination and emission excitons. Then, the dependence of the material's fluorescence intensity on temperature was calculated by using the Arrhenius equation,52 resulting in values of 53.18 and 27.25 meV for Cu-MOFs and POM@Cu-MOFs, respectively (Fig. 4f). Accordingly, Eb of POM@Cu-MOFs is smaller than that of Cu-MOFs, indicating that exciton dissociation in POM@Cu-MOFs is more likely to occur in Cu-MOFs, which can more effectively dissociate into free carriers and improve charge transfer.
The morphology of POM@Cu-MOFs remains essentially unchanged after the photocatalytic reduction of U(VI) under visible light irradiation (Fig. S13a†). Besides this, amorphous sheets are also observed, which was ascribed to the products of uranium-bearing precipitates, as SEM-EDS mapping demonstrates that U is uniformly distributed in POM@Cu-MOFs after photoreduction (Fig. S13b†). The electrons generated within photoexcited POM@Cu-MOFs reduce the soluble U(VI) adsorbed on the surface of POM@Cu-MOFs to form the insoluble uranium-bearing precipitate, thus achieving the reduction and isolation of U(VI). The XRD peak of the POM@Cu-MOF catalyst has minimal alterations after adsorption and photocatalysis compared to that of the original sample (Fig. 5a). The new IR bands appearing at 918.4 and 1534.9 cm−1 for the catalyst after adsorption and photocatalytic reduction are ascribed to the vibration of UO22+ and uranyl hydroxy species, respectively (Fig. 5b).53 These results confirm the complexing ability of POM@Cu-MOFs for UO22+. XPS was performed to evaluate the stability of the POM@Cu-MOF catalyst after adsorption and irradiation. The valence state of uranium after adsorption and the photocatalytic reduction reaction is inconsistent (Fig. S14a†). U 4f7/2 and 4f5/2 can be divided into two peaks: 380.6 eV and 391.5 eV for U(IV), and 383.1 eV and 394.7 eV for U(VI), indicating the coexistence of U(VI) and U(IV) species after the photocatalytic reduction of U(VI) by POM@Cu-MOFs.54 For the uranium complexation produced after POM@Cu-MOF absorption of U(VI) under dark conditions, only the peak of U(VI) is present at 382.2 and 393.2 eV (Fig. 5c). After the photocatalytic reduction, the XPS peaks of Mo and P in POM shift to a lower binding energy (Fig. 5d and e), indicating spontaneous electron transfer at the interface. Meanwhile, compared with the spectra of POM@Cu-MOFs after absorption of U(VI), the POM@Cu-MOFs show increased binding energies of Cu 2p, C 1s, and lattice oxygen in O 1s (Fig. 5f and S14b, c†). More importantly, the XPS peak intensity of oxygen-containing species adsorbed on the POM@Cu-MOF surface (Oads) in O 1s increases after the reaction because unsaturated U
O groups are formed during the reduction process (Fig. S12d†). These results demonstrate that the electrons transfer from Cu-MOFs to POM and finally to UO22+.
Based on the above results, a possible mechanism for reducing U(VI) is proposed (Fig. 6). Due to the porous architecture, the Cu-MOF exhibits commendable adsorption capacity and POM displays reversible multi-electron transfer reactions, endowing POM@Cu-MOFs with notable adsorption and photocatalytic processes. The reaction process as follows: (1) the oxo-metal cluster and P–O bonds within POM@Cu-MOFs were the active coordination sites for UO22+ adsorption, allowing UO22+ in solution to be effectively bound to the catalyst surface; (2) when exposed to light, the complexation of POM and Cu-MOFs could efficiently inhibit the recombination of photogenerated electrons and holes to accelerate the migration of electrons. Ultimately, the adsorbed UO22+ is reduced to an insoluble uranium-bearing precipitate via e− and ˙O2− generated in the photocatalytic process. (3) Photoexcited electrons from Cu-MOFs can easily transfer to POM, because the CB levels of Cu-MOFs are lower than those of POM, thus improving UO22+ reduction and facilitating activation of H2O into active ˙OH radicals. (4) The holes in the VB were transferred from POM to Cu-MOFs, and H2O molecules were consumed at the same time to generate ˙OH. POM, with excellent electrical conductivity, facilitates the separation of photogenerated electrons, thereby improving photocatalytic performance. Free radicals and the reaction process can be described by using the following equation:
| POM@Cu-MOFs + hν → e− + h+ | (2) |
| H2O + h+ → OH + H+ | (3) |
| UO22+ + e− → UO2+x (s, amorphism) | (4) |
| O2 + e− → ˙O2− | (5) |
| UO22+ + 2O2− → UO2+x (s, amorphism) + O2 | (6) |
![]() | ||
| Fig. 6 A plausible photoreduction mechanism of U(VI) over the POM@Cu-MOF heterojunction under light irradiation. | ||
| qt = (C0 − Ct)V/m | (7) |
| U(VI) removal rate = (C0 − Ce)/C0 × 100% | (8) |
Footnotes |
| † Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4sc05349d |
| ‡ These authors contributed equally to this work. |
| This journal is © The Royal Society of Chemistry 2024 |