Issue |
Natl Sci Open
Volume 4, Number 1, 2025
Special Topic: Nuclear Environment Advances
|
|
---|---|---|
Article Number | 20240003 | |
Number of page(s) | 15 | |
Section | Earth and Environmental Sciences | |
DOI | https://doi.org/10.1360/nso/20240003 | |
Published online | 13 September 2024 |
RESEARCH ARTICLE
Selective capture of iodate anions by a cerium-based metal-organic framework
State Key Laboratory of Radiation Medicine and Protection, School of Radiation Medicine and Protection, Collaborative Innovation Center of Radiological Medicine of Jiangsu Higher Education Institutions, Soochow University, Suzhou 215123, China
* Corresponding author (email: shuaowang@suda.edu.cn)
Received:
24
January
2024
Revised:
19
March
2024
Accepted:
13
May
2024
Effective remediation of radioactive IO3− is highly desirable for fuel reprocessing, medical waste disposal, and nuclear accidents. However, the nature of high solubility, strong mobility, and extremely hard to bind with minerals for IO3− makes this task an enormous challenge. Herein, a metal-organic framework material [Ce(IV)-MOF-808] with available Ce(IV) sites was used to efficiently remove IO3−. Ce(IV)-MOF-808 exhibits superior selectivity and one of the highest adsorption capacities (623 mg·g−1) for IO3− removal. Moreover, Ce(IV)-MOF-808 shows great adsorption performance for IO3− at low concentrations, and the distribution coefficient (Kd) value was calculated to be 2.60 × 106 mL·g−1. The exceptional IO3− uptake performance is attributed to the high affinity between Ce(IV) cluster and the oxyanion based on the comprehensive analysis of zeta potential and X-ray photoelectron spectroscopy (XPS) results, in which IO3− can easily replace the weakly coordinated ligand and form a strong coordination structure of Ce-O-I-O2. More importantly, Ce(IV)-MOF-808 exhibits excellent uptake performance for IO3− from both the simulated Beishan groundwater system and Hanford groundwater system in the dynamic column separation test, indicating the highly promising practical application of Ce(IV)-MOF-808 in IO3− remediation from actual radioactive wastes.
Key words: iodate / metal-organic framework / selectivity / distribution coefficient / adsorption
© The Author(s) 2024. Published by Science Press and EDP Sciences.
This is an Open Access article distributed under the terms of the Creative Commons Attribution License (https://creativecommons.org/licenses/by/4.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
INTRODUCTION
Radioiodine is a great risk driver for human health due to its high toxicity and specific bioaccumulation in the thyroid gland [1–3]. 129I is one of the most detrimental radioisotopes with a high fission yield and a long half-life (1.57 × 107 years) [4] and as a result, leads to a large inventory in nuclear wastes. However, the highly soluble and mobile nature of radioiodine in aqueous solution endows it with an extremely high risk of release during nuclear waste disposal, waste storage, or nuclear accidents. In fact, a large amount of radioiodine has been leaked into the environment in the Fukushima disaster [5] and several regions where nuclear wastes are processed or stored. For example, groundwater from the Department of Energy’s Savannah River Site (SRS) is highly contaminated with 129I [6]. As one kind of medical radionuclide, 131I (t1/2 = 8 d) has been well-used in cancer diagnosis and treatment, resulting in a large quantity of medical radioiodine wastewater. With the development of nuclear medicine, there will be an increasing risk of radioiodine contamination from the growing volume of medical wastewater [7–10]. Radioiodine primarily exists in the form of iodide (I−) and iodate (IO3−) in an aqueous solution. IO3− is the dominant species in highly aerobic conditions, which accounts for approximately 70.6% of the total iodine in the groundwater at the Hanford site [11]. Therefore, the safe and effective removal of radioactive iodate is crucial for radioactive waste management, medical waste disposal, and contaminated water remediation.
Recently, numerous works have been dedicated to the investigation of diverse iodine sorbents for the efficient removal of radioactive iodine under aqueous and/or gas environments, such as minerals [12,13], zeolites [14,15], activated carbon [16–19], and covalent organic frameworks (COFs) [20–23]. However, compared with I−, I2, and CH3I, research on IO3− removal is seriously inadequate. Metal oxides and their composite materials [24–26], anion exchange resins [27], and layered double hydroxides (LDHs) [3,28] are the most commonly used sorbents for the separation of IO3−. Metal oxides and metal composites, such as Cu2O@CH, a Cu2O-loaded three-dimensional bulk cationic hydrogel composite exhibit excellent radiation resistance in radioiodine removal, but show poor capability and low distribution coefficient for IO3− (Kd = 5.33 × 102 mL·g−1) [1]. Traditional anion exchange resins, such as Purolite A530E, show an inferior adsorption capacity for IO3−, and exhibit low uptake selectivity and poor radiation resistance in IO3− removal [27]. LDHs are a class of well-used water remediation sorbents with the merits of low cost, high adsorption capacity and good radiation resistance. However, it shows poor uptake selectivity and low Kd values for IO3− removal [29]. Therefore, it is of great significance to develop promising sorbents with the virtues of high uptake capacity, excellent selectivity, and high Kd values aiming at radioactive IO3−.
According to Pearson’s theory of hard-soft-acid base (HSAB), high-valent metal ions such as Ce4+, which are characterized as hard Lewis acids, show high affinity to hard Lewis bases, such as IO3− [30]. In addition, the extremely low solubility of Ce(IO3)4 in aqueous solution (Ksp = 5 × 10−17) [31] reveals the strong combination between Ce(IV) and IO3− in aqueous solution. Therefore, Ce(IV)-based adsorbents will be promising sorbents for remediating IO3− contamination. However, most Ce(IV)-based sorbents, such as CeO2, show extremely poor uptake capacity for IO3− remediation due to the lack of available coordination sites in the structure. Therefore, aiming to make the Ce(IV) site available, metal ions and ligands with open metal sites or weak coordination bonds are highly desirable. Metal-organic frameworks (MOFs) constructed by organic ligands bridging inorganic nodes (metal ions or clusters) are one of the most promising materials for fabricating Ce(IV)-based adsorbents with accessible Ce(IV) clusters due to the variability of ligands and coordination structures. However, the systematic design and development of MOFs with open metal sites or weak coordination bonds for the efficient removal of radioactive IO3− is still in its infancy.
Herein, MOFs with Ce(IV) hexanuclear clusters [Ce(IV)-MOF-808] [32] were selected for the removal of radioactive IO3− due to the capability to make the Ce(IV) site available through the unsaturated coordination of Ce(IV) and the weak coordination groups upon Ce(IV) (OH− and H2O). Ce(IV)-MOF-808 not only exhibits ultrahigh capture capacity (623 mg·g−1) and superior distribution coefficient (Kd = 2.60 × 106 mL·g−1) for IO3− but also shows high selectivity for IO3− under the condition of superabundant competing ions. More importantly, it exhibits an excellent capability for IO3− remediation in the simulated Beishan and Hanford groundwater system, indicating the powerful potential application of Ce(IV)-MOF-808 for the efficient removal of IO3− from nuclear waste liquid solutions or contaminated natural water systems.
RESULTS AND DISCUSSION
Characterization of Ce(IV)-MOF-808
Ce(IV)-MOF-808 [Ce6(μ3-O)4(μ3-OH)4(BTC)2(OH)6(H2O)6] was obtained under a solvothermal reaction from a mixture of the corresponding organic ligand (H3BTC) dissolved in DMF and an aqueous cerium ammonium nitrate solution (Figure 1a, 1b) [32]. In this structure, each [Ce6(μ3-O)4(μ3-OH)4]12+ cluster is consisted with six hydroxyl groups and six H2O molecules through weak coordination. Ce(IV)-MOF-808 exhibits a 3,6-connected 3D framework with spn and tetrahedral cages of approximately 4 Å and large adamantane pores of 18 Å in diameter (Figure 1c, 1d). The crystalline structure of Ce(IV)-MOF-808 was characterized by Fourier transform infrared (FT-IR) and Powder X-ray diffraction (PXRD). As shown in Figure 1e, the presence of a wide band range of 2500–3700 cm−1 in Ce(IV)-MOF-808 indicates the presence of the –OH group in the Ce(IV)-MOF-808 structure [33]. The v(C=O) of the COOH group disappears at 1717 cm−1, while v(COO-) appears. At the same time, the vas(C=O) of v(COO-) in Ce(IV)-MOF-808 is split into three peaks (1647 cm−1, 1610 cm−1 and 1551 cm−1) and vs(C=O) into two peaks (1434 cm−1 and 1366 cm−1) [34]. The new peak at 556 cm−1 is caused by the tensile vibration of Ce–O in Ce(IV)-MOF-808 [35]. These results indicate the formation of a coordination bond between H3BTC and Ce(IV). Furthermore, the PXRD pattern of synthesized Ce(IV)-MOF-808 (Figure 1f) is consistent with that of the simulated result, confirming the successful synthesis of Ce(IV)-MOF-808.
![]() |
Figure 1 Crystal structure of Ce(IV)-MOF-808. Ball-and-stick model (a) and polyhedral model (b) of the Ce6 cluster. The flat sheet (c) and 3D framework (d) of Ce(IV)-MOF-808. Atom colors: Ce = purple; C = gray; O = red; O (OH/H2O) = green; H atoms are omitted for clarity. (e) FT-IR spectra of Ce(IV)-MOF-808 (blue) and H3BTC (black). (f) PXRD pattern of Ce(IV)-MOF-808: simulated (black) and synthesized (red). |
pH effect analysis
Considering the high concentration of H+/OH− in radioactive waste, the influence of different pH conditions on the adsorption of IO3− by Ce(IV)-MOF-808 was first investigated. As shown in Figure 2a, the uptake percentage of IO3− by Ce(IV)-MOF-808 shows negligible changes and over 95% of IO3− can be removed in the range of pH 4–10. Besides, when the pH value changed to 2, the removal of IO3− obviously increased and nearly 99% of IO3− was removed. To further explore the relation between pH and adsorption behavior, the zeta potential of Ce(IV)-MOF-808 was measured. As shown in Figure S2, Ce(IV)-MOF-808 exhibits positive charge under acidic conditions (pH < 7.6) and shows a strong attraction toward IO3− through electrostatic interactions. Besides, the high uptake capability of Ce(IV)-MOF-808 for IO3− removal is likely attributed to the vast –OH sites on the surface of the material that can be exchanged with IO3− (Ce-OH + IO3− → Ce-O-I-O2 + OH−), leading to a high capture rate of IO3−. Therefore, the subsequent batch experiments (except the related experiments of simulated waste liquid) explored the adsorption effect of Ce(IV)-MOF-808 on IO3− solutions at unregulated pH (pH~7) and pH 2, respectively.
![]() |
Figure 2 (a) Effect of pH value on the sorption properties of IO3− by Ce(IV)-MOF-808 (Conditions: contact time = 12 h, [I]initial = 50 mg·L−1, msorbent/Vsolution = 1 g·L−1). Adsorption kinetics curve under the condition of pH 7 (b) and pH 2 (c) (Conditions: [I]initial = 50 mg·L−1, msorbent/Vsolution = 1 g·L−1). Adsorption isotherm curve under the condition of pH 7 (d) and pH 2 (e) (Conditions: contact time = 12 h, msorbent/Vsolution = 1 g·L−1). (f) Comparison of equilibrium time and the maximum capacity for reported adsorbents and Ce(IV)-MOF-808 in this work. The detailed experimental conditions, including the solid/liquid ratio and pH values, are listed in Table S6. |
Sorption kinetics analysis
The sorption kinetics analysis of Ce(IV)-MOF-808 for IO3− was studied in neutral and acidic solutions (pH 2). Ce(IV)-MOF-808 was mixed with IO3− solution (pH 7 or pH 2) at a solid-liquid ratio of 1 g·L−1 and stirred at room temperature. IO3− was converted to I3− at 2% KI and 0.1 M H3PO4 (Figures S3, S4) [27], and the concentration of IO3− in aqueous solution was monitored by the characteristic absorption peak of I3− in the UV-Vis spectrum at 350 nm (Figure S5). The characteristic peak of I3− at 350 nm in the UV absorption spectrum disappeared at 60 min in both neutral and acidic solutions, indicating that Ce(IV)-MOF-808 can reach adsorption equilibrium (pH 7, > 93%; pH 2, > 98%) within 60 min (Figure 2b, 2c). This rapid adsorption kinetics may be attributed to the accessible and replaceable hydroxyl sites in Ce(IV)-MOF-808 [36]. The kinetics of Ce(IV)-MOF-808 for IO3− is faster than most of the previously reported materials, e.g., dried duckweed powder (5760 min) [37], DNTD (1440 min) [38], and pomelo peel (5760 min) [39]. In addition, pseudo-first-order and pseudo-second-order kinetic models were used to fit the adsorption processes in both environments. As shown in Table S3, the kinetics data is in line with the pseudo-second-order kinetic model, indicating that the uptake of IO3− on Ce(IV)-MOF-808 is accord with the chemisorption process (Figure S6). Additionally, the distribution coefficient (Kd) of Ce(IV)-MOF-808 toward IO3− reaches as high as 2.60 × 106 mL·g−1 (Table S4), indicating the powerful applications of Ce(IV)-MOF-808 in practical IO3− remediation.
Sorption isotherm analysis
The adsorption capacity of Ce(IV)-MOF-808 toward IO3− was further evaluated by adding Ce(IV)-MOF-808 to a series of IO3− solutions with initial concentrations ranging from 0 to 750 mg·L−1 (pH 7 or pH 2), and the adsorption isotherm curve was fitted by the Freundlich and Langmuir models (Figure 2d, 2e). The adsorption curve plotted by the equilibrium concentration and adsorption capacities conform to the Langmuir model. The saturated adsorption capacities for pH 7 and pH 2 were calculated to be 361 mg·g−1 and 623 mg·g−1, respectively (Table S5). More importantly, the adsorption capacity at pH 2 is slightly higher than the reported highest capacity (ZrSbO2: 612.5 mg·g−1) [40], setting a new record in adsorption capacity towards IO3−. The high adsorption capacity under acidic conditions is attributed to the positively charged surface of Ce(IV)-MOF-808 under acidic conditions which introduces more adsorption sites into Ce(IV)-MOF-808 based on electrostatic interactions. In fact, this capacity is significantly higher than most of previously reported adsorbents, including oxide composites (Cu2O@CH, 313.4 mg·g−1) [1], natural materials (dried duckweed powder, 5.1 mg·g−1) [38], resins (Purolite A530E, 53.3 mg·g−1) [27], naturally occurring porous minerals (HDPy-bent, 50.2 mg·g−1) [41], and LDHs (NiAl LDH, 395.5 mg·g−1) [29] (Figure 2f and Table S6). This result indicates that Ce(IV)-MOF-808 may be a promising candidate for practical IO3− remediation from nuclear waste with the virtue of generating less amount of secondary radioactive waste.
Selectivity
Since competing anions, such as SO42−, NO3−, and Cl−, often excessively coexist with IO3− in the radioactive waste solutions. The uptake selectivity of Ce(IV)-MOF-808 for IO3− was studied under the condition of different concentrations of competing anions. As shown in Figure 3a, the removal rate of IO3− shows negligible changes as the concentrations of NO3− or Cl− increased under the condition of pH 7. When the concentration of NO3− and Cl− increases to 100 times than that of IO3−, the removal rates of IO3− by Ce(IV)-MOF-808 remain above 92.1% and 91.8%, respectively. Meanwhile, Ce(IV)-MOF-808 shows the same trend on IO3− removal under the condition of pH 2 that 94% of IO3− is removed under 100 times excess of both NO3− and Cl−, indicating the excellent uptake selectivity of Ce(IV)-MOF-808 for IO3− removal (Figure 3b). Moreover, anions (SO42−) with higher charge density usually have greater competition than lower charge density anions IO3− according to the anti-Hofmeister bias [42]. However, Ce(IV)-MOF-808 can still remove 46.6% and 47.1% of IO3− when the molar ratio of SO42− to IO3− is as high as 100: 1 under the condition of pH 7 and pH 2, respectively. This excellent uptake selectivity of Ce(IV)-MOF-808 for IO3− makes it a promising candidate to selectively remove IO3− from wastes with high ionic strengths.
![]() |
Figure 3 Effect of competing anions on the removal percentage of IO3− by Ce(IV)-MOF-808 under the condition of pH 7 (a) and pH 2 (b). Conditions: contact time = 12 h, [I]initial = 20 mg·L−1, msorbent/Vsolution = 1 g·L−1. |
Radiation stability
Considering the high radioactivity of nuclear wastes, the radiation resistance of Ce(IV)-MOF-808 was evaluated by comparing the FT-IR spectrum, PXRD pattern and adsorption capacity of Ce(IV)-MOF-808 before and after irradiation. As shown in Figure S7, the FT-IR spectrum and the PXRD pattern of Ce(IV)-MOF-808 remain almost unchanged after being irradiated for different doses, showing that the structure of Ce(IV)-MOF-808 remains stable under irradiation. In addition, Ce(IV)-MOF-808 shows a high uptake capacity for IO3− after undergoing an extremely high dose of β-ray irradiation (200 kGy) as the original sample (Figure 4a, 4b). These results indicate the outstanding radiation resistance of Ce(IV)-MOF-808.
![]() |
Figure 4 The removal percentage of IO3− under the condition of pH 7 (a) and pH 2 (b) after β irradiation compared with the original Ce(IV)-MOF-808 sample (Conditions: contact time = 12 h, [I]initial = 200 mg·L−1, msorbent/Vsolution = 1 g·L−1). The removal percentage of IO3− by Ce(IV)-MOF-808 from the simulated Hanford groundwater (c) and Beishan groundwater (d) (Conditions: contact time = 12 h, [I]initial = 1 mg·L−1). Dynamic adsorption curve of Ce(IV)-MOF-808-packed column for IO3− capture from the simulated Hanford groundwater (e) and Beishan groundwater (f) (Conditions: [I]initial = 1 mg·L−1, msorbent = 200 mg, flow rate = 0.15 mL·min−1). |
IO3− removal from simulated nuclear wastes
To further explore the practical application of Ce(IV)-MOF-808 in the treatment of radioactive waste, we analyzed the capture capacity of Ce(IV)-MOF-808 towards IO3− in the simulated groundwater at the Hanford site in the United States and the Beishan high-level waste disposal site in China. As shown in Table S1, high amounts of competing ions coexist in the simulated wastes. For example, the amounts of Cl− and SO42− far exceed than that of IO3− by more than 1200 times in the simulated BGW (Table S2), which presents a huge challenge for the selective removal of IO3−. Exhilaratingly, over 99% IO3− in both simulated systems was removed by Ce(IV)-MOF-808 with a solid/liquid ratio of 5 g·L−1 (Figure 4c, 4d and Table S7). These results show the bright practical application of Ce(IV)-MOF-808 in IO3− remediation from actual radioactive waste.
Dynamic column separation
To further validate the potential of Ce(IV)-MOF-808 in practical scenarios, we evaluated the uptake performance of Ce(IV)-MOF-808 via a dynamic column breakthrough experiment. Ce(IV)-MOF-808 (200 mg) was packed in a plastic column with an inner diameter of 12.5 mm and further connected to the peristaltic pump for the dynamic column experiment (Figure S8). The flow rate of the simulated HGW or BGW solution was set to 0.15 mL·min−1 through the control system. As shown in Figure 4e, almost complete adsorption of IO3− by Ce(IV)-MOF-808 is observed within the initial 40 mL in the simulated HGW. Then, the concentrations of IO3− in the effluent increased continuously, hinting that the adsorption sites were gradually occupied by IO3− and the remaining adsorption sites gradually decreased during this period. The dynamic adsorption process reached equilibrium after being carried out with ~72 mL of IO3− stock solution, and the effluent concentration (Ce) was very close to that of the stock solution (C0). Similarly, for the simulated BGW (Figure 4f), Ce(IV)-MOF-808 could almost completely adsorb IO3− within the initial 100 mL and reached equilibrium after being carried out with 120 mL of IO3− stock solution. The above results show that Ce(IV)-MOF-808 exhibits excellent uptake selectivity for IO3− and is highly desirable for actual IO3− separation from radioactive wastes.
Investigation of the adsorption mechanism
To further explore the adsorption mechanism, the structure of Ce(IV)-MOF-808 before and after IO3− adsorption was characterized by energy-dispersive X-ray spectroscopy (EDS) mapping, FT-IR spectroscopy, and X-ray photoelectron spectroscopy (XPS). In the EDS mapping of Ce(IV)-MOF-808_IO3−, the uniform distribution of element I indicates its homogeneous adsorption (Figures S9 and S10). The FT-IR spectra of Ce(IV)-MOF-808 before and after IO3− adsorption were collected in the range of 4000–500 cm−1 (Figure S11). The characteristic peak of –OH on Ce(IV)-MOF-808 decreased significantly after adsorption, indicating that the excellent IO3− uptake capability possibly originates from the replacement of –OH. The changes of surface elements and interaction mechanism before and after Ce(IV)-MOF-808 adsorption IO3− were analyzed by XPS (Figure 5). As shown in Figure 5a, new peaks identified at 624 and 54 eV in full XPS spectra appear after adsorption of IO3− and correspond to the characteristics of I 3d orbitals and I 4d orbitals, respectively [43]. This result indicates that IO3− is successfully captured onto Ce(IV)-MOF-808. Meanwhile, I 3d3/2 (pH = 7, 635.80 eV; pH = 2, 635.73 eV) and I 3d5/2 (pH = 7, 624.38 eV; pH = 2, 624.24 eV) in the I 3d spectra after adsorption (Figure 5b) also confirm the above conclusion.
![]() |
Figure 5 XPS spectra of Ce(IV)-MOF-808 before and after IO3− adsorption. (a) Full survey XPS spectra. (b) I 3d of Ce(IV)-MOF-808 before and after adsorption. (c) O 1s of Ce(IV)-MOF-808 before and after adsorption. |
The I 3d and O 1s spectra were widened and studied in detail to obtain a better understanding of the adsorption mechanism. The binding energies of I 3d3/2 and I 3d5/2 in Ce(IV)-MOF-808 after adsorption are shifted toward higher binding energies compared to KIO3 (Figure 5b), suggesting that the electron cloud density around I decreases after adsorption, likely because of the reduction of electrons given to I by O in IO3− after adsorption caused by the formation of strong coordination between O in IO3− and Ce in Ce(IV)-MOF-808.
The O 1s spectrum of the original Ce(IV)-MOF-808 can be divided into three peaks at 532.72, 531.25, and 529.46 eV, which were associated with Ce-OH, Ce-O-C, and Ce-O-Ce, respectively (Figure 5c). The peak of Ce-OH on Ce(IV)-MOF-808 decreases obviously after adsorption compared with that of the original Ce(IV)-MOF-808, accompanied by the new peaks of Ce-O-I [530.44 eV (pH = 7) and 530.51 eV (pH = 2)] caused by the coordination between O in IO3− and Ce in Ce(IV)-MOF-808. This result illustrates that the weak coordination functional group, –OH, is replaced by IO3− in the adsorption process and forms a strong coordination between IO3− and Ce in Ce(IV)-MOF-808. Therefore, combined with the XPS result of I 3d, the adsorption mechanism can function as following equation:
Ce-OH + IO3− → Ce-O-I-O2 + OH− (1)
Besides, the peak area of Ce-O-I adsorbed in the acidic environment (22.68%) is larger than that of Ce-O-I in the neutral environment (19.55%). This is due to high hydroxyl protonation on the surface of Ce(IV)-MOF-808 under acidic conditions, making the formation of Ce-O-I easier. This result is consistent with the bath adsorption results that the adsorption capacity of Ce(IV)-MOF-808 for IO3− in the acidic environment is higher than that in the neutral environment.
CONCLUSIONS
In summary, Ce(IV)-MOF-808 exhibits excellent adsorption properties toward IO3− removal including fast adsorption kinetics, ultrahigh adsorption capacity (623 mg·g−1), excellent selectivity, and high Kd value. Moreover, superior IO3− remediation is achieved both in static adsorption and dynamic column breakthrough from the simulated Hanford groundwater and Beishan groundwater, coupled with decent radiation resistance, endowing Ce(IV)-MOF-808 with promising application prospects in practical scenarios. The XPS results demonstrate that the excellent adsorption performance of Ce(IV)-MOF-808 is attributed to the high affinity between IO3− and Ce that the weakly coordinated hydroxyl groups in Ce(IV)-MOF-808 can be easily replaced by IO3−, leading to the strongly coordinated Ce-O-I between IO3− and Ce. This work not only reports a promising candidate for IO3− remediation from complex radioactive water systems, but also highlights new opportunities for the rational design of promising adsorbent materials for efficient removal of IO3−.
EXPERIMENTAL
Materials and reagents
Cerium ammonium nitrate [(NH4)2Ce(NO3)6, purity > 99.0%] was supplied by Shanghai Macklin Biochemical Technology Co., Ltd. 1,3,5-Benzenetricarboxylic acid (H3BTC, purity > 99.0%) and potassium iodate (KIO3, purity > 99.0%) were purchased from Shanghai Titan Scientific Co., Ltd. N, N-Dimethylformamide (DMF) was supplied by Jiangsu Qiangsheng Functional Chemical Co., Ltd. Formic acid (HCOOH) was purchased from Sinopharm Chemical Reagent Co., Ltd. Acetone (CH3COCH3) was supplied by Shanghai Lingfeng Chemical Reagent Co., Ltd. NaCl, NaNO3, Na2SO4, HCl, NaOH, H3PO4, KI and other reagents were purchased and used from suppliers in analytically pure form.
Synthesis of Ce(IV)-MOF-808
Ce(IV)-MOF-808 was synthesized based on previous reports [32]. Ce(IV)-MOF-808 was prepared by a solvothermal method, as shown in Scheme S1. H3BTC (22.4 mg, 106 μmol), (NH4)2Ce(NO3)6 (175.3 mg, 318 μmol), H2O (0.6 mL), DMF (1.2 mL) and HCOOH (257 μL) were successively added to a 10 mL scintillation bottle, and heated at 100°C on an oil bath magnetic stirring heater for 15 min. The products were separated by centrifugation, washed with DMF (2 mL × 2) and acetone (2 mL × 4), and dried at 70°C for 2 h to obtain a light yellow solid (Figure S1).
Characterizations
The β irradiation of materials was performed using the electron irradiation accelerator device of Shanghai Institute of Applied Physics (1.5 MeV; 1.4 mA; 80 kW). Scanning electron microscopy (SEM) images and energy dispersive X-ray spectroscopy (EDS) were obtained by an EVO 18 scanning electron microscope from Carl Zeiss, Germany. Fourier transform infrared (FT-IR) spectra in the range of 4000–400 cm−1 were recorded by a Thermo Nicolet iS50 conventional Fourier infrared spectrometer (USA). Power X-ray diffraction (PXRD) datas were collected at 2θ from 3° to 40° with a step of 0.02° by D8 Advance X-ray powder diffraction instrument in Bruker, Germany (Cu Kα rays (λ = 1.54056 Å; 40 kV; 40 mA)). The zeta potential was measured with Malvern Zetasizer Nano ZS90 (UK). X-ray photoelectron spectroscopy (XPS) spectra were measured using Thermo Scientific ESCALAB Xi+ (USA). A monochromatic Al Kα source (Mono Al Kα) with an energy of 1486.6 eV was used as the excitation source. The data were fitted and analyzed by Avantage XPS software, and the binding energies were calibrated through C 1s hydrocarbon peak at 284.8 eV. The concentrations of IO3− in the solution were determined by monitoring the characteristic absorption peak of I3− derived from IO3− at 350 nm by a UV-2600 Vis-spectrometer in Shimadzu, Japan.
Batch experiments
All adsorption experiments were performed at 25°C by using the batch adsorption method. Typically, Ce(IV)-MOF-808 (5 mg) was placed into aqueous solution containing certain contents of IO3− (5 mL). The pH was adjusted with diverse concentrations of NaOH and/or HCl solutions. After stirring the mixture on a magnetic mixer at 300 r·min−1 for desired adsorption time, the solid-liquid phase was separated with a 0.22 μm nylon membrane filter (SANJIA Biochemical Supplies). Taking into account the phenomenon that IO3− has no UV characteristic absorption peak in UV-Vis spectrum, the concentration of IO3− was measured by the characteristic peak of I3− at 350 nm based on the reaction that IO3− is converted into I3− after mixing 2 mL of sample with 1 mL of 0.1 M H3PO4 solution and 1 mL of 2% KI solution. In an acidic solution containing excess I−, IO3− reacts with I− and converts to elemental iodine, which then interacts with I− to form I3− [44]. The specific reactions are as follows:
IO3−+5I−+6H+→3I2+3H2O (2)
I2+I−→I3− (3)
IO3−+8I−+6H+→3I3−+3H2O (4)
The removal percentage (η) and sorption amount (qe) of Ce(IV)-MOF-808 for IO3− were calculated by the following formulas:
where C0 (mg·L−1) is the initial concentration of IO3− in solution; Ce (mg·L−1) is the residual concentration of IO3− after adsorption; V (mL) and m (mg) are the volume of IO3− solution and the mass of Ce(IV)-MOF-808, respectively.
Effect of pH study
The pH values were adjusted by adding negligible volumes of HCl and/or NaOH solutions. Ce(IV)-MOF-808 (5 mg) was added to IO3− aqueous solution (5 mL, 50 mg·L−1) at different pH values (2 to 10). The suspension was separated with a 0.22 μm nylon membrane filter after being stirred for 6 h, and the concentration of IO3− in aqueous solution was measured by a UV-Vis spectrometer.
Sorption kinetics study
The kinetics experiments were performed at pH 7 or pH 2 with a solid/liquid ratio of 1 g·L−1. Typically, Ce(IV)-MOF-808 (100 mg) was added to IO3− solution (100 mL, 50 mg·L−1). After being stirred for a certain time (0, 1, 2, 5, 10, 20, 30, 40, 50, 60, 90, 120, 150, 180, 210, 240, 270, and 330 min), the liquid phase was separated with a 0.22 μm nylon membrane filter and the concentration of IO3− was analyzed by a UV-Vis spectrometer.
Two kinetic models, pseudo-first-order (PFO) and pseudo-second-order (PSO) models, were used to fit the kinetics data. The linearized forms of the two models are as follows [45]:
Pseudo-first-order model:
Pseudo-second-order model:
where qe (mg·g−1) and qt (mg·g−1) are the sorption amount at equilibrium time and time t (min), respectively; k1 (min−1) and k2 (g·mg−1·min−1) are the kinetic rate constants of PFO and PSO, respectively. The linearized plot of PFO and PSO was obtained when we plotted ln(qe - qt) and t/qt against t, respectively.
The distribution coefficient (Kd) of Ce(IV)-MOF-808 toward IO3− was calculated by the following equation:
Sorption isotherm study
The sorption isotherm experiments of Ce(IV)-MOF-808 were carried out by varying the initial IO3− concentrations from 0 to 750 mg·L−1. In each sample, Ce(IV)-MOF-808 (5 mg) was added into IO3− aqueous solution (5 mL) under the condition of pH 7 or pH 2 with a solid/liquid ratio of 1 g·L−1. After being stirred for 6 h, the suspension liquid was separated and the concentration of IO3− was measured by a UV-Vis spectrometer.
Two isotherm models, the Freundlich and Langmuir isotherm models were used to fit the sorption process. The Freundlich equation is based on the empirical equation for heterogeneous surface adsorption. The stronger binding sites are occupied first, and the binding strengths are gradually decreased as the occupation of the active sites. The Freundlich isotherm model is expressed in the following equations:
where KF ((mg·g−1) × (L1/n·mg−1/n)) and n are the Freundlich constants corresponding to the sorption capacity and strength, respectively. The linearized plot was obtained by plotting lnqe against lnCe, and KF and n could be calculated from the slope and intercept.
The Langmuir isotherm model assumes that the sorption process is based on monolayer and that all sites are equal. The linear equation of the Langmuir isotherm model is expressed as follows [46]:
where qm (mg·g−1) is the maximum sorption amount of IO3− corresponding to complete monolayer coverage; KL (L·mg−1) is a constant related to the sorption capacity and energy of sorption, which characterizes the affinity of the adsorbate with the adsorbent. The linearized plot was obtained when we plotted Ce/qe against Ce, and qm and KL could be calculated from the slope and intercept.
Anion selectivity study
The effects of competing anions, including NO3−, SO42−, and Cl−, were determined by adding different concentrations of NaNO3, Na2SO4, or NaCl solutions (0.157 mM, 0.785 mM, 1.57 mM, and 15.7 mM) into a 0.157 mM IO3− solution. Ce(IV)-MOF-808 (5 mg) was added to the above solution (5 mL) under the condition of pH 7 or pH 2. The suspension was separated after being stirred for 6 h, and the concentration of IO3−, after sorption in aqueous solution was measured by a UV-Vis spectrometer.
Radiation resistance measurements
The β irradiation experiment was conducted using an electron beam by an electron accelerator. Ce(IV)-MOF-808 powder was exposed to β-irradiation for four doses (50, 100, 150, and 200 kGy). The anti-irradiation property of Ce(IV)-MOF-808 was characterized by FT-IR and PXRD spectroscopy, while the irradiated samples were further subjected to IO3− adsorption experiments (pH 7 or pH 2) to determine the irradiation stability of Ce(IV)-MOF-808.
Removal of IO3− in simulated nuclear wastes
The simulated Hanford groundwater (HGW) and Beishan groundwater (BGW) were prepared according to Table S1 and Table S2. Ce(IV)-MOF-808 (5 mg/25 mg) was added to the above simulated solutions (5 mL). The suspension was separated after being stirred for 6 h, and the concentration of IO3− after sorption in aqueous solution was measured by a UV-Vis spectrometer.
Dynamic column adsorption experiments
Typically, Ce(IV)-MOF-808 (200 mg) was packed in a plastic column (12.5 mm I.D.). The simulated BGW or HGW solution flowed through the column at a flow rate of 0.15 mL·min−1. The concentrations of IO3− in the effluent were evaluated by a UV-Vis spectrometer.
Funding
This work was supported by the Intergovernmental International Cooperation of the National Key R&D Program of China (2022YFE0105300), the National Key R&D Program of China (2021YFB3200400), the National Natural Science Foundation of China (22306136, 22425061, 22176139, U2267222, and U1967217), the China National Postdoctoral Program for Innovative Talents (BX2021206), the China Postdoctoral Science Foundation (2021M702390), the Natural Science Foundation of Jiangsu (BK20230510), the New Cornerstone Science Foundation through the XPLORER PRIZE, and the Priority Academic Program Development of Jiangsu Higher Education Institutions (PAPD).
Author contributions
S.W. and Z.C. conceived and supervised the project. Y.Z. synthesized and characterized the material. Y.Z., J.L., Q.G. and L.L. performed the adsorption experiments and analyzed the data. L.H. assisted in the drawing of TOC. Y.Z. and F.Z. collected and analyzed the XPS data. M.Z. aided in the irradiation experiments. Y.Z., J.L., L.C., and S.W. wrote the paper. All authors discussed the results and commented on the manuscript.
Conflict of interest
The authors declare no conflict of interest.
Supplementary information
Supplementary file provided by the authors. Access here
The supporting information is available online at https://doi.org/10.1360/nso/20240003. The supporting materials are published as submitted, without typesetting or editing. The responsibility for scientific accuracy and content remains entirely with the authors.
References
- Yan C, Li J, Tan C, et al. Cuprous oxide-based cationic hydrogel by the integration of enrichment and immobilization of radioiodine (I– , IO3– ) in aqueous solution. ACS Appl Mater Interfaces 2023; 15: 28135-28148. [Article] [Google Scholar]
- Chen P, He X, Pang M, et al. Iodine capture using Zr-based metal–organic frameworks (Zr-MOFs): Adsorption performance and mechanism. ACS Appl Mater Interfaces 2020; 12: 20429-20439. [Article] [Google Scholar]
- Kang J, Levitskaia TG, Park S, et al. Nanostructured MgFe and CoCr layered double hydroxides for removal and sequestration of iodine anions. Chem Eng J 2020; 380: 122408. [Article] [NASA ADS] [CrossRef] [Google Scholar]
- Asmussen RM, Westesen A, Cordova EA, et al. Iodine removal from carbonate-containing alkaline liquids using strong base resins, hybrid resins, and silver precipitation. Ind Eng Chem Res 2023; 62: 3271-3281. [Article] [Google Scholar]
- Mclaughlin PD, Jones B, Maher MM. An update on radioactive release and exposures after the Fukushima Dai-ichi nuclear disaster. BJR 2012; 85: 1222-1225. [Article] [Google Scholar]
- Beals DM, Hayes DW. Technetium-99, iodine-129 and tritium in the waters of the Savannah River Site. Sci Total Environ 1995; 173-174: 101-115. [Article] [NASA ADS] [CrossRef] [PubMed] [Google Scholar]
- Rose PS, Smith JP, Cochran JK, et al. Behavior of medically-derived 131I in the tidal Potomac River. Sci Total Environ 2013; 452-453: 87-97. [Article] [NASA ADS] [CrossRef] [PubMed] [Google Scholar]
- Jiménez F, López R, Pardo R, et al. The determination and monitoring of 131I activity in sewage treatment plants based on A2/O processes. Radiat Meas 2011; 46: 104-108 [CrossRef] [MathSciNet] [Google Scholar]
- Barquero R, Agulla MM, Ruiz A. Liquid discharges from the use of radionuclides in medicine (diagnosis). J Environ Radioact 2008; 99: 1535-1538. [Article] [NASA ADS] [CrossRef] [PubMed] [Google Scholar]
- Fischer HW, Ulbrich S, Pittauerová D, et al. Medical radioisotopes in the environment—Following the pathway from patient to river sediment. J Environ Radioact 2009; 100: 1079-1085. [Article] [Google Scholar]
- Moore RC, Pearce CI, Morad JW, et al. Iodine immobilization by materials through sorption and redox-driven processes: A literature review. Sci Total Environ 2020; 716: 132820. [Article] [NASA ADS] [CrossRef] [PubMed] [Google Scholar]
- Choung S, Kim M, Yang JS, et al. Effects of radiation and temperature on iodide sorption by surfactant-modified bentonite. Environ Sci Technol 2014; 48: 9684-9691. [Article] [Google Scholar]
- Kaplan DI, Serne RJ, Parker KE, et al. Iodide sorption to subsurface sediments and illitic minerals. Environ Sci Technol 2000; 34: 399-405. [Article] [Google Scholar]
- Inglezakis VJ, Satayeva A, Yagofarova A, et al. Surface interactions and mechanisms study on the removal of iodide from water by use of natural zeolite-based silver nanocomposites. Nanomaterials 2020; 10: 1156. [Article] [CrossRef] [PubMed] [Google Scholar]
- Pham TCT, Docao S, Hwang IC, et al. Capture of iodine and organic iodides using silica zeolites and the semiconductor behaviour of iodine in a silica zeolite. Energy Environ Sci 2016; 9: 1050-1062. [Article] [Google Scholar]
- Ikari M, Matsui Y, Suzuki Y, et al. Removal of iodide from water by chlorination and subsequent adsorption on powdered activated carbon. Water Res 2015; 68: 227-237. [Article] [CrossRef] [PubMed] [Google Scholar]
- Li D, Kaplan DI, Sams A, et al. Removal capacity and chemical speciation of groundwater iodide (I−) and iodate (IO3−) sequestered by organoclays and granular activated carbon. J Environ Radioact 2018; 192: 505-512. [Article] [Google Scholar]
- Li D, Kaplan DI, Price KA, et al. Iodine immobilization by silver-impregnated granular activated carbon in cementitious systems. J Environ Radioact 2019; 208-209: 106017. [Article] [Google Scholar]
- Hoskins JS, Karanfil T, Serkiz SM. Removal and sequestration of iodide using silver-impregnated activated carbon. Environ Sci Technol 2002; 36: 784-789. [Article] [Google Scholar]
- Lin Y, Jiang X, Kim ST, et al. An elastic hydrogen-bonded cross-linked organic framework for effective iodine capture in water. J Am Chem Soc 2017; 139: 7172-7175. [Article] [NASA ADS] [CrossRef] [PubMed] [Google Scholar]
- Liu X, Zhang A, Ma R, et al. Experimental and theoretical insights into copper phthalocyanine-based covalent organic frameworks for highly efficient radioactive iodine capture. Chin Chem Lett 2022; 33: 3549-3555. [Article] [CrossRef] [Google Scholar]
- He L, Chen L, Dong X, et al. A nitrogen-rich covalent organic framework for simultaneous dynamic capture of iodine and methyl iodide. Chem 2021; 7: 699-714. [Article] [NASA ADS] [CrossRef] [Google Scholar]
- Zhang Y, He L, Pan T, et al. Superior iodine uptake capacity enabled by an open metal-sulfide framework composed of three types of active sites. CCS Chem 2023; 5: 1540-1548. [Article] [CrossRef] [Google Scholar]
- Liu S, Kang S, Wang H, et al. Nanosheets-built flowerlike micro/nanostructured Bi2O2.33 and its highly efficient iodine removal performances. Chem Eng J 2016; 289: 219-230. [Article] [NASA ADS] [CrossRef] [Google Scholar]
- Zhao Q, Chen G, Wang Z, et al. Efficient removal and immobilization of radioactive iodide and iodate from aqueous solutions by bismuth-based composite beads. Chem Eng J 2021; 426: 131629. [Article] [NASA ADS] [CrossRef] [Google Scholar]
- Wang N, Zhang G, Xiong R, et al. Synchronous moderate oxidation and adsorption on the surface of γ-MnO2 for efficient iodide removal from water. Environ Sci Technol 2022; 56: 9417-9427. [Article] [NASA ADS] [CrossRef] [PubMed] [Google Scholar]
- Zhao Y, Li J, Chen L, et al. Efficient removal of iodide/iodate from aqueous solutions by Purolite A530E resin. J Radioanal Nucl Chem 2023; 332: 1193-1202. [Article] [Google Scholar]
- Zhang D, Liu XY, Zhao HT, et al. Application of hydrotalcite in soil immobilization of iodate (IO3− ). RSC Adv 2018; 8: 21084-21091. [Article] [NASA ADS] [CrossRef] [PubMed] [Google Scholar]
- Kang J, Cintron-Colon F, Kim H, et al. Removal of iodine (I− and IO3−) from aqueous solutions using CoAl and NiAl layered double hydroxides. Chem Eng J 2022; 430: 132788. [Article] [CrossRef] [Google Scholar]
- Ho TL. Hard soft acids bases (HSAB) principle and organic-chemistry. Chem Rev 1975; 75: 1-20 [CrossRef] [Google Scholar]
- Lide DR. CRC Handbook of Chemistry and Physics. Boca Raton: CRC Press, 2004 [Google Scholar]
- Lammert M, Glißmann C, Reinsch H, et al. Synthesis and characterization of new Ce(IV)-MOFs exhibiting various framework topologies. Cryst Growth Des 2017; 17: 1125-1131. [Article] [NASA ADS] [CrossRef] [Google Scholar]
- Pervez MN, Chen C, Li Z, et al. Tuning the structure of cerium-based metal-organic frameworks for efficient removal of arsenic species: The role of organic ligands. Chemosphere 2022; 303: 134934. [Article] [Google Scholar]
- Chu L, Guo J, Huang Z, et al. Excellent catalytic performance over acid-treated MOF-808(Ce) for oxidative desulfurization of dibenzothiophene. Fuel 2023; 332: 126012. [Article] [Google Scholar]
- He J, Pei C, Yang Y, et al. The structural design and valence state control of cerium-based metal-organic frameworks for their highly efficient phosphate removal. J Cleaner Production 2021; 321: 128778. [Article] [Google Scholar]
- Jacobsen J, Ienco A, D′Amato R, et al. The chemistry of Ce-based metal–organic frameworks. Dalton Trans 2020; 49: 16551-16586. [Article] [CrossRef] [PubMed] [Google Scholar]
- Zhang K, Chen T. Sorption and removal of iodate from aqueous solution using dried duckweed (Landoltia punctata) powder. J Radioanal Nucl Chem 2018; 316: 543-551. [Article] [NASA ADS] [CrossRef] [Google Scholar]
- Liu P, Chen T, Zheng J. Removal of iodate from aqueous solution using diatomite/nano titanium dioxide composite as adsorbent. J Radioanal Nucl Chem 2020; 324: 1179-1188. [Article] [Google Scholar]
- Da T, Chen T. Optimization of experimental factors on iodate adsorption: A case study of pomelo peel. J Radioanal Nucl Chem 2020; 326: 511-523. [Article] [Google Scholar]
- Suorsa V, Otaki M, Virkanen J, et al. Pure and Sb-doped ZrO2 for removal of IO3− from radioactive waste solutions. Int J Environ Sci Technol 2022; 19: 5155-5166. [Article] [NASA ADS] [CrossRef] [Google Scholar]
- Yang J, Tai W, Wu F, et al. Enhanced removal of radioactive iodine anions from wastewater using modified bentonite: Experimental and theoretical study. Chemosphere 2022; 292: 133401. [Article] [Google Scholar]
- Zhang Y, Cremer PS. Chemistry of Hofmeister anions and osmolytes. Annu Rev Phys Chem 2010; 61: 63-83. [Article] [Google Scholar]
- Li J, Ma X, Zhao C, et al. A novel Ce(IO3)4 catalyst: Facile preparation and high activity in degradation of organic dyes without light irradiation at room temperature. J Phys Chem Solids 2017; 100: 33-39. [Article] [NASA ADS] [CrossRef] [Google Scholar]
- Schmitz G. Kinetics and mechanism of the iodate-iodide reaction and other related reactions. Phys Chem Chem Phys 1999; 1: 1909-1914 [Google Scholar]
- Li J, Dai X, Zhu L, et al. 99TcO4− remediation by a cationic polymeric network. Nat Commun 2018; 9: 3007. [Article] [CrossRef] [PubMed] [Google Scholar]
- Langmuir I. The adsorption of gases on plane surfaces of glass, mica and platinum. J Am Chem Soc 1917; 40: 1361-1403 [Google Scholar]
All Figures
![]() |
Figure 1 Crystal structure of Ce(IV)-MOF-808. Ball-and-stick model (a) and polyhedral model (b) of the Ce6 cluster. The flat sheet (c) and 3D framework (d) of Ce(IV)-MOF-808. Atom colors: Ce = purple; C = gray; O = red; O (OH/H2O) = green; H atoms are omitted for clarity. (e) FT-IR spectra of Ce(IV)-MOF-808 (blue) and H3BTC (black). (f) PXRD pattern of Ce(IV)-MOF-808: simulated (black) and synthesized (red). |
In the text |
![]() |
Figure 2 (a) Effect of pH value on the sorption properties of IO3− by Ce(IV)-MOF-808 (Conditions: contact time = 12 h, [I]initial = 50 mg·L−1, msorbent/Vsolution = 1 g·L−1). Adsorption kinetics curve under the condition of pH 7 (b) and pH 2 (c) (Conditions: [I]initial = 50 mg·L−1, msorbent/Vsolution = 1 g·L−1). Adsorption isotherm curve under the condition of pH 7 (d) and pH 2 (e) (Conditions: contact time = 12 h, msorbent/Vsolution = 1 g·L−1). (f) Comparison of equilibrium time and the maximum capacity for reported adsorbents and Ce(IV)-MOF-808 in this work. The detailed experimental conditions, including the solid/liquid ratio and pH values, are listed in Table S6. |
In the text |
![]() |
Figure 3 Effect of competing anions on the removal percentage of IO3− by Ce(IV)-MOF-808 under the condition of pH 7 (a) and pH 2 (b). Conditions: contact time = 12 h, [I]initial = 20 mg·L−1, msorbent/Vsolution = 1 g·L−1. |
In the text |
![]() |
Figure 4 The removal percentage of IO3− under the condition of pH 7 (a) and pH 2 (b) after β irradiation compared with the original Ce(IV)-MOF-808 sample (Conditions: contact time = 12 h, [I]initial = 200 mg·L−1, msorbent/Vsolution = 1 g·L−1). The removal percentage of IO3− by Ce(IV)-MOF-808 from the simulated Hanford groundwater (c) and Beishan groundwater (d) (Conditions: contact time = 12 h, [I]initial = 1 mg·L−1). Dynamic adsorption curve of Ce(IV)-MOF-808-packed column for IO3− capture from the simulated Hanford groundwater (e) and Beishan groundwater (f) (Conditions: [I]initial = 1 mg·L−1, msorbent = 200 mg, flow rate = 0.15 mL·min−1). |
In the text |
![]() |
Figure 5 XPS spectra of Ce(IV)-MOF-808 before and after IO3− adsorption. (a) Full survey XPS spectra. (b) I 3d of Ce(IV)-MOF-808 before and after adsorption. (c) O 1s of Ce(IV)-MOF-808 before and after adsorption. |
In the text |
Current usage metrics show cumulative count of Article Views (full-text article views including HTML views, PDF and ePub downloads, according to the available data) and Abstracts Views on Vision4Press platform.
Data correspond to usage on the plateform after 2015. The current usage metrics is available 48-96 hours after online publication and is updated daily on week days.
Initial download of the metrics may take a while.