Issue
Natl Sci Open
Volume 4, Number 1, 2025
Special Topic: Nuclear Environment Advances
Article Number 20240029
Number of page(s) 15
Section Earth and Environmental Sciences
DOI https://doi.org/10.1360/nso/20240029
Published online 19 September 2024

© The Author(s) 2024. Published by Science Press and EDP Sciences.

Licence Creative CommonsThis 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

Nuclear energy stands out as an indispensable green energy source due to its reliability and consistency in mitigating a range of environmental issues stemming from substantial greenhouse gas emissions [1,2]. As energy shortages intensify and nuclear energy consumption rises, uranium as a basic raw material for the nuclear industry has garnered considerable attention [35]. However, uranium mining, nuclear industry processes, and even nuclear accidents generate substantial volumes of radioactive wastewater containing uranium [6,7]. U(VI) predominantly exists in the form of soluble UO22+ with significant mobility in aquatic environments. The high mobility of UO22+ allows it to spread rapidly, exerting a profound impact on the ecological environment [7,8]. The severe chemical toxicity and radiological hazard of uranium can cause damage on various human organs upon contact or exposure, posing a significant threat to human health when it enters the body through the mouth, nose, or skin [7,8]. Consequently, the efficient removal and capture of uranium from complex radioactive wastewater are of vital importance not only to address environmental pollution but also to ensure the safety and sustainable development of nuclear energy.

Various treatment methods of wastewater containing uranium have been developed such as solvent extraction [9], chemical precipitation [9], membrane filtration [10], chemical reduction, and adsorption [6,1113]. Among them, adsorption is considered as a reliable technique due to the simple operation, low cost, and high efficiency. Many uranium adsorption materials have been reported, including minerals [14,15], metal oxides [16,17], metal chalcogenides [7,1820], resin [21,22], metal organic frameworks (MOFs) [2124], porous organic polymers (POPs) [25,26], covalent organic frameworks (COFs) [2729], and so on. Transition metal dichalcogenides stand out due to their sulfur or selenium-rich surfaces and their highly tunable properties. These materials can effectively track and remove metal ions through the interaction between sulfur or selenium and metal ions [11,18,30]. Molybdenum disulfide (MoS2), a prominent example of transition metal dichalcogenides, has garnered significant interest due to its remarkable attributes such as a narrow bandgap energy, non-toxicity, biocompatibility, low cost, and easy synthesis [31,32]. More importantly, it is a promising material for removing uranium by the interaction of Lewis soft base S2− with the relatively soft UO22+ [11,18,30]. However, the previous study indicated that the raw MoS2 has a low affinity for UO22+ [18]. To address this limitation, researchers have prepared various MoS2 composite materials to enhance the U(VI) adsorption performance of MoS2 by either increasing the number of uranium adsorption sites or improving the material’s hydrophilicity [3335], such as layered double hydroxide modified MoS2 [33], phosphonate-functionalized MoS2 [34], and sodium dodecyl sulfate modified MoS2 [35]. It is the key to select the appropriate carrier to prepare high-performance composite materials based on molybdenum sulfide.

Collagen fibers (CF), a structural protein known for its mechanical strength and hydrophilicity, have found extensive applications in separation processes due to their fibrous structure, which promotes efficient mass transfer [36,37]. CF is endowed with numerous active functional groups, such as –OH, –CONH2, –COOH, and –NH2, that can interact with metal ions, making them an excellent candidate for removing pollutants [7,36]. Consequently, CF emerges as a promising adsorbent for the removal of U(VI) from aqueous solutions. Currently, a variety of CF-based adsorption materials have been developed for uranium capture [7,3840]. The primary strategies involve: (i) activating and modifying the CF structure to increase the active sites of materials for effective U(VI) adsorption, such as alkaline treatment or grafting organic functional molecules onto CF [39,40]; (ii) loading metal ions or nanoparticles. For example, CF loaded with Ti4+ [38] or nanoparticles [41], have shown the ability of effective uranium removal from wastewater.

Our strategy is to composite CF with metal chalcogenides. Metal chalcogenides contribute specific adsorption sites for U(VI), while CF enriches the composite with additional functional groups, such as amino and carboxyl groups, which are beneficial to U(VI) capture. Moreover, the hydrophilic nature and fibrous structure of CF are advantageous for promoting mass transfer and thus facilitating U(VI) adsorption. However, only one case of nano-ZnS/alkali-activated CF composite has been reported by us [7]. It indicates that CF serves as an excellent carrier for metal sulfides, enabling the in-situ growth of zinc sulfide on CF to prepare composite material. Additionally, the carboxyl and amino groups of CF can assist in binding uranium, further enhancing the adsorption capacity of the composite material. Compared with bulk zinc sulfide, this composite material exhibits higher uranium removal performance, which proves the correctness of our compositing strategy. Therefore, combining metal chalcogenides with CF can give composite materials the advantages of both metal chalcogenides and CF. However, CF with the characteristic of lower thermal stability is not suitable for in situ growth of other metal chalcogenides such as molybdenum sulfide and molybdenum selenide, as their synthesis temperature is higher. Therefore, there is a pressing need to develop a green, simple, and universally applicable compositing strategy or method to prepare composite materials with transition metal dichalcogenides and CF for higher uranium adsorption performance.

In this work, we proposed a universal preparation strategy for preparing MQ2-collagen fiber (M = Mo, W; Q = S, Se) composite materials, that is, green, simple and efficient co-ball milling of transition metal dichalcogenides with CF. MQ2 as an adsorbent is uniformly distributed on the carrier CF after compositing, exposing more active sites. Meanwhile, CF can also provide additional functional groups such as amino carboxyl groups as adsorption sites. The presence of CF makes the material more hydrophilic, and its fibrous structure can promote mass transfer and adsorption of U(VI). Therefore, compared with bulk MQ2, composite materials exhibit better U(VI) adsorption performance. We take the MoS2-CF as the representative to study the removal performance of U(VI) in detail. The results show that MoS2-CF has the saturated U(VI) adsorption capacity of 301 mg g−1, and can selectively remove uranium in the presence of a large amount of alkali and alkali earth metal ions and even in actual water samples. At the same time, the material also shows recycle ability for U(VI) adsorption. The mechanism of uranium adsorption is revealed through various characterizations including Fourier transform infrared spectroscopy (FT-IR), X-ray photoelectron spectra (XPS), and elemental distribution mappings. The efficient uranium adsorption originates from the interaction of amino and carboxyl groups on CF and sulfur on the surface of MoS2 for U(VI). Overall, our compositing strategy is universal for low-cost and eco-friendly composite materials combining transition metal dichalcogenide with collagen fiber by co-ball milling. The prepared MQ2-CF composite materials have excellent application potential for highly efficient capture of U(VI) from wastewater.

RESULTS AND DISCUSSION

Characterizations of MoS2-CF

The composite material (MoS2-CF) of MoS2 and CF is constructed using green methods, in which water is the processing solvent, natural polymers (collagen fibers) are as carriers/adsorption enhancers and MoS2 acts as the adsorption unit. In brief, CF, MoS2, and solvent water are placed in a ball milling tank for ball milling, and then the ball milling solution is freeze-dried to obtain composite materials (Figure 1). The final adsorbent material is fibrous with an average diameter at the nanoscale. The morphology of MoS2 that synthesized based on literature can be seen from SEM images as nanospheres formed by stacking nanosheets (Figure 2A). From SEM, it can be seen that CF presents a smooth surface (Figure 2B), while elemental distribution mapping and EDS show that it contains C, H, and N (Figure S1). After compositing, uniform molybdenum sulfide nanoparticles appear on the surface of CF (Figure 2C). The corresponding EDS and elemental distribution mapping images of MoS2-CF display the presence of Mo and S elements originating from MoS2, confirming the successful loading of MoS2 on CF (Figures 2D and S1).

thumbnail Figure 1

(A) Illustration of self-assembly preparation of MoS2-CF. (B) Photos of MoS2, CF, and MoS2-CF; the enlarged image is the SEM of MoS2-CF.

thumbnail Figure 2

Characterization of MoS2-CF. SEM of MoS2 (A), CF (B) and MoS2-CF (C). (D) Elemental distribution mapping of MoS2-CF. (E) XPS of CF, MoS2, and MoS2-CF. (F) Water contact angle of CF, MoS2, and MoS2-CF. (G) Zeta potential of CF, MoS2, and MoS2-CF after ball milling.

To gain a deeper understanding of the composite, the adsorbent is further characterized. SEM, elemental distribution mapping, and EDS confirmed the successful synthesis of MoS2 (Figure S2). The PXRD diffraction peaks and Raman spectroscopic analysis of obtained MoS2 are consistent with 1T phase of MoS2 reported in the literature [42,43] (Figure S2). At the same time, after calcination, it can transform into 2H phase MoS2 (Figure S2d). The PXRD pattern of MoS2-CF confirms MoS2 successfully loaded on CF (Figure S3a). Importantly, the characteristic peak position of MoS2 remains unchanged after compositing, indicating that the structure of MoS2 is not affected throughout the entire preparation process [44]. FTIR of various materials can provide a deeper understanding of the surface chemical groups present on the materials (Figure S3b). The characteristic peaks of CF are as follows: a broad absorption peak can be seen at 3100–3650 cm−1, which corresponds to the vibrational absorption of O–H and N–H. Vibration absorption of C–H in CF can be observed at 2950 cm−1 and 2880 cm−1. The absorption peaks at 1650 cm−1 and 1546 cm−1 correspond to the C=O vibrational absorption from amides and carboxyl groups, respectively [45,46]. Meanwhile, these corresponding peaks are also observed in the composite material without significant changes. Additionally, XPS of the composite material shows the peaks of S 2p (161.5 eV), Mo 3d3/2 (230.6 eV), and Mo 3d5/2 (226.9 eV) (Figure 2E), further confirming the successful loading of MoS2 onto CF. Compared with the pristine MoS2, we observed the significant enhancement in the hydrophilicity of the composite material, as evidenced by water contact angle measurements, which is beneficial to the adsorption of U(VI) by MoS2-CF [33] (Figure 2F). Moreover, FT-IR and PXRD of MoS2-CF after soaking solutions with different pH values (pH 1–12) do not show significant changes compared to the untreated material, indicating that MoS2-CF exhibits excellent acid-base stability [7] (Figure S4).

Zeta potential measurements shed light on the compositing mechanism. After ball milling, the single CF exhibits positive Zeta potential (1.0 mV), while MoS2 shows negative Zeta potential (−30.6 mV) (Figure 2G). The electrostatic interaction between them facilitates the self-assembly of MoS2 and CF into composite material. The negative Zeta potential of MoS2-CF (−28.7 mV) suggests that it can enhance the adsorption of U (VI) through Coulomb interaction [12].

Adsorption of uranium

Effect of composite ratio

In the prepared transition metal dichalcogenides and CF composite materials, we chose MoS2-CF as the representative to conduct a detailed study on the U adsorption performance. Considering the influence of ball milling experimental conditions on the uranium adsorption performance of composite materials, we investigated the effects of zirconium bead diameter, mass, and ball milling time on the uranium adsorption performance of composite materials. It was found that, the composite material prepared with 60 mg of 2 mm diameter zirconium beads had the highest uranium adsorption performance at a ball milling time of 16 hours, reaching 85.3 mg g−1 (Figure S5). Therefore, such ball milling conditions were chosen for the preparation of composite materials. The ratio of components in a composite material significantly influences its adsorption capabilities, prompting us to investigate how the ratio of MoS2 and CF affects U(VI) adsorption (E). MoS2 and CF exhibit relatively low adsorption capacities for U(VI), at 12.6 mg g−1 and 34.8 mg g−1, respectively. However, the U(VI) adsorption capacity of MoS2-CF with the different ratio of MoS2 and CF (ranging from 4:1 to 1:8) surpasses that of the pristine MoS2 and CF, which may be derived from the improved hydrophilicity of MoS2-CF and the accelerated mass transfer process of the fibrous sample of composite. These findings validate the effectiveness of our composite strategy in improving the adsorption capabilities of transition metal chalcogenides. As the proportion of MoS2 within the composite material increases, so does the number of available adsorption sites provided by MoS2 for U(VI). This increase in available sites is instrumental in enhancing the adsorption capacity of the composite material for U(VI), resulting in a gradual rise in U(VI) removal capacity. Notably, when the ratio of MoS2 to CF is 1:5, the adsorption capacity reaches its maximum at 85.3 mg g−1, which is 6.7 times greater than pristine MoS2. Then, an increase in MoS2 proportion results in a decrease in U(VI) adsorption capacity of the MoS2-CF composite. This decline may stem from an excess of MoS2, which obscures adsorption sites, thereby reducing the overall adsorption capacity. Taking the material cost and adsorption capacity into account, we ultimately selected a composite material with a ratio of 1:5 between MoS2 and CF for subsequent experiments.

Effect of pH value

As is well known, pH value can significantly impact the speciation of uranium, which in turn affects its adsorption by various adsorbents [47,48]. Consequently, it is crucial to examining the adsorption behavior of U(VI) under different pH conditions. At low pH (pH < 5.0), UO22+ ions occupy the dominant position. However, at higher pH values, uranium is less likely to remain as uranyl cations and may instead form hydrolytic precipitates [25,39]. With this in mind, we focused our investigation on the adsorption of UO22+ at pH values less than 5. At a pH of 1, the adsorption capacity of MoS2-CF is only 3.33 mg g−1, which may be due to a large amount of H+ competing with U(VI) for adsorption sites in the solution. Subsequently, as the pH increases, the number of hydrogen ions competing with U(VI) for adsorption sites in the solution decreases, leading to the removal capacity of MoS2-CF for U(VI) gradually increases, reaching 83.5 mg g−1 at pH of 5 (Figure 3B). Therefore, pH of 5 was selected for subsequent experiments.

thumbnail Figure 3

Adsorption of MoS2-CF for U(VI). (A) The effect of composite ratio of MoS2 and CF on U(VI) adsorption (C0U = 150 ppm, pH = 5, t = 12 h, m/V = 1 g L−1). (B) The pH effect on adsorption of MoS2-CF (C0U = 200 ppm, t = 12 h, m/V = 1 g L−1). (C) Adsorption kinetics simulation (C0U = 150 ppm, pH = 5, m/V = 1 g L−1). (D) Simulation of adsorption isotherms (C0U = 25–550 ppm, pH = 5, t= 12 h, m/V = 1 g L−1). (E) The effect of different molar ratios coexisting Na+, K+, Mg2+, and Ca2+ on U(VI) adsorption of MoS2-CF (C0U = 10 ppm, pH = 5, t = 12 h, m/V = 1 g L−1). (F) U(VI) adsorption of MoS2-CF in actual water samples (Taken from Fuzhou, Fujian) (C0U = 150 ppm, pH = 5, t = 12 h, m/V = 1 g L−1).

Effect of time

We explored the effect of adsorption time on the capture of U(VI). U(VI) in the solution was rapidly adsorbed by MoS2-CF within the first 2 hours. As the adsorption process continues, available sites on MoS2-CF are gradually occupied, resulting in a decelerating adsorption rate and ultimately reaching adsorption equilibrium at 12 hours with the adsorption capacity of MoS2-CF for U(VI) being 95.7 mg g−1 (Figure 3C) [48]. Therefore, we chose 12 hours as the subsequent adsorption time. Fitting the adsorption kinetics curve using different models (Table S1), we found that the R2 of pseudo-second-order kinetic model (R2 = 0.985) was higher than pseudo-first-order kinetic model (R2 = 0.945) and intra-particle diffusion model (R2 = 0.576) (Table S1), which indicating the adsorption process of MoS2-CF is more in line with pseudo-second-order kinetic model. This result indicates that the adsorption process is chemical adsorption.

Effect of U concentration

In order to obtain the saturated adsorption capacity of MoS2-CF, we systematically measured the adsorption isotherms of adsorbents through sequencing batch experiments (Figure 3D). The fitting of the adsorption isotherm curve revealed that it aligns more closely with the Langmuir adsorption isotherm model (R2 = 0.997) compared to Freundlich adsorption isotherm model (R2 = 0.972) (Table S2), indicating the adsorption process is a single-layer adsorption. The saturated adsorption capacity of MoS2-CF for U(VI) was calculated to be 301.1 mg g−1 by Langmuir adsorption isotherm model (Figure 3D), which is much higher than the reported 2H-MoS2 (45.7 mg g−1) and sodium dodecyl sulfate modified molybdenum disulfide composites (98.4 mg g−1)[18,35]. This indicated that MoS2-CF possesses superior U(VI) adsorption performance, which validates the effectiveness of our composite strategy in enhancing the adsorption capacity for U(VI).

Effect of competitive ions and adsorption in actual water samples

In actual water environments, the presence of competing ions can interfere with the adsorption performance of materials. Therefore, it is crucial to explore the adsorption selectivity in the presence of other coexisting ions [7]. Our studies reveal that MoS2-CF maintains adsorption selectivity even under with high concentrations of monovalent ions including Na+ and K+. Even in the presence of individual Na+ ions at a 1000-fold molar excess to U, the distribution ratio (KdU) of MoS2-CF can remain at 2.8 × 103 mg L−1 (Figure S6). Similarly, MoS2-CF maintains high adsorption performance in the presence of high concentrations of individual K+ ions with the KdU of 2.3 × 103 mg L−1 with K+ ions at the 1000-fold molar excess to U (Figure S7). Under the coexisting divalent Mg2+ and Ca2+ ions, MoS2-CF also maintains high adsorption performance. Under 1000-fold molar excess of Mg2+ ions, the KdU of the MoS2-CF can still be as high as 3.5 × 103 mg L−1 (Figure S8), while the KdU of MoS2-CF can remain at 2.3 × 103 mg L−1 in the condition of Ca/U molar ratio more than 1000 times (Figure S9). These results indicate that MoS2-CF has excellent resistance to interference from Na+, K+, Mg2+ and Ca2+. Furthermore, when 10-fold excess of Na+, K+, Mg2+, and Ca2+ are present simultaneously, the KdU of MoS2-CF can still maintain 3.7 × 103 mg L−1. This demonstrates that the composite material possesses excellent selectivity [49] (Figure 3E). Adsorption experiments on labeled U(VI) in tap water, lake water, and seawater confirm that MoS2-CF can effectively capture U(VI) in actual water environment (Figure 3F). The adsorption performance of MoS2-CF in actual water samples is sea water > lake water > tap water. Notably, even in seawater, which contains complex competitive ions, the adsorption capacity of MoS2-CF for U(VI) still reaches 112.2 mg g−1, likely due to the highest pH of seawater.

Recycle performance of MoS2-CF

The recycling performance of the material is also crucial. After MoS2-CF has completed U(VI) adsorption (MoS2-CF-U), the adsorbed U(VI) can be removed by soaking the MoS2-CF-U in 0.1 M sodium carbonate solution, and then the desorbed composite material can be reused again for U(VI) adsorption. In the first cycle, the uranium adsorption capacity for U(VI)of MoS2-CF is 81.2 mg g−1. After the five cycles, there was no significant change in the adsorption performance for U(VI) with the adsorption capacity being 79.4 mg g−1, indicating MoS2-CF has good cycling performance (Figure S10a). Moreover, no significant changes in the absorption peak were found in the FT-IR after cycling, indicating that the composite material has good recycling stability (Figure S10b).

Adsorption mechanism

Studying the adsorption mechanism of MoS2-CF for U(VI) is crucial for understanding the adsorption behavior of U(VI) and providing guidance for the preparation of more efficient adsorption materials. From the SEM images of MoS2-CF-U, it is found that there is no significant change in the surface morphology before and after U(VI) adsorption, suggesting the good stability of MoS2-CF (Figure 4A). The PXRD of MoS2-CF and MoS2-CF-U are consistent (Figure S11a). The presence of U is confirmed by elemental mapping and EDS analysis, indicating successfully U(VI) adsorption by MoS2-CF (Figure S11b) [50]. XPS analysis provides further evidence for U(VI) adsorption. The peak at 381.8 eV in XPS of MoS2-CF-U is attributed to U 4f7/2, confirming the presence of U (Figure 4B and 4C). High-resolution XPS spectra of N 1s, O 1s and S 2p measured for MoS2-CF-U are presented in Figure 4. The N 1s peak for MoS2-CF observed at 399.33 eV shifts to 399.57 eV for MoS2-CF-U, indicating that N coordinates with U (Figure 4E) [7,13]. Similarly, after completing the adsorption of U(VI), the O 1s peaks shifts from 530.75 eV (C–O) and 530.99 eV (C=O) to 531.20 eV and 531.43 eV, respectively, suggesting coordination between carboxyl groups and U (Figure 4F) [46,51]. In addition, we found that the S 2p peaks of MoS2-CF observed at 160.66 eV, 161.84 eV, 162.11 eV, and 163.26 eV belonging to S–H 2 p1/2 and S–H 2 p3/2 and Mo–S 2 p1/2 and Mo–S 2 p3/2, respectively. These peaks of MoS2-CF-U shifted to 160.81 eV, 161.96 eV, 162.30 eV, and 163.45 eV, respectively (Figure 4D). This indicates that the abundant S atoms from the surface of MoS2 also participate in the interaction for uranium [7]. FT-IR further supports successful uranium adsorption (Figure 4G), with a characteristic UO22+ absorption peak appearing at 912 cm−1 [13,51]. Additionally, the carboxyl group’s vibration absorption peak shifts from 1546 cm−1 to 1535 cm−1 after UO22+ absorption, consistent with literature reports of such shifts upon coordination with uranium [13,51]. Based on these findings, we think that the abundant sulfur sites on the surface of MoS2 and the amino and carboxyl groups from CF work synergistically to coordination with uranium, jointly completing the capture of uranium from aqueous solution.

thumbnail Figure 4

Characterization of MoS2-CF after U(VI) adsorption. (A) SEM and elemental distribution mappings of MoS2-CF-U. (B) XPS of MoS2-CF and MoS2-CF-U. (C) High-resolution XPS spectra of U 4f of MoS2-CF-U. (D) High-resolution XPS spectra of S 2p of MoS2-CF and MoS2-CF-U. (E) High-resolution XPS spectra of N 1s of MoS2-CF and MoS2-CF-U. (F) High-resolution XPS spectra of O 1s of MoS2-CF and MoS2-CF-U. (G) FT-IR of MoS2-CF and MoS2-CF-U.

Verification of strategy universality

To demonstrate the versatility of our compositing strategy, we selected commercially available transition metal dichalcogenides (MoSe2, WS2, and WSe2) as validation materials for the synthesis of other composite materials and applied them to uranium adsorption. Given MoS2-CF’s optimal performance at a 1:5 ratio, we applied the same ratio between transition metal dichalcogenides and CF for the synthesis of MoSe2-CF, WS2-CF, and WSe2-CF. The photo of the MoSe2-CF, WS2-CF, and WSe2-CF shows that their color closely resembles their pristine dichalcogenide counterparts (Figure S12). Comprehensive characterizations (including SEM, Mapping, EDS, PXRD, FT-IR, and XPS) of MoSe2, WS2, WSe2, MoSe2-CF, WS2-CF, and WSe2-CF (Figures S13–19) confirm the successful fabrication of the composites. Zeta-potential of MoSe2, WS2, WSe2 MoSe2-CF, WS2-CF, and WSe2-CF were −11.9, −34.0, −24.8, −18.6, −21.6, and −34.4 eV, respectively, suggesting that the assembly mechanism of MoSe2-CF, WS2-CF, and WSe2-CF are electrostatic self-assembly, and their surface negative potential is advantageous for U(VI) adsorption (Figure 5A). Meanwhile, we anticipate that the hydrophilic properties and fibrous structure of the composite materials will enhance their U(VI) adsorption capabilities compared to the pristine dichalcogenides (Figure 5B). Adsorption experiments on these composite materials confirmed that their U(VI) removal capacity exceeds that of the unmodified dichalcogenides, highlighting the improved adsorption performance after compositing (Figure 5C). After the U(VI) adsorption, SEM image of the MoSe2-CF, WS2-CF, and WSe2-CF showed no significant change in the morphology (Figure 5D, 5E, and 5F), but also the appearance of uranium elements in EDS and elemental mapping indicated the successful adsorption of uranium (Figure S20). XPS analysis of the MoSe2-CF-U (380.9, and 391.8 eV), WS2-CF-U (381.2, and 392.0 eV), and WSe2-CF-U (380.7, and 391.5 eV) also showed U 4f absorption peaks, further confirming U(VI) capture (Figure 5G). Their FT-IR also display uranyl absorption peaks around 916 cm−1 (Figure 5H), providing additional evidence of successful U(VI) adsorption [7,13,51]. Therefore, our experiments demonstrate that the compositing strategy is universal and can be effectively applied to prepare a series of MQ2-CF (M = Mo, W; Q = S, Se) composite materials with enhanced uranium adsorption performance.

thumbnail Figure 5

Characterization of U(VI) adsorption by MoSe2-CF, WS2-CF, and WSe2-CF. (A) Zeta potential of MoSe2, WS2, WSe2, MoSe2-CF, WS2-CF, and WSe2-CF after ball milling. (B) Water contact angle of MoSe2-CF, WS2-CF, and WSe2-CF. (C) Adsorption of uranium by MoSe2, WS2, WSe2, MoSe2-CF, WS2-CF (C0U = 150 ppm, pH = 5, t = 12 h, m/V = 1 g L−1). SEM and elemental distribution mappings of MoSe2-CF-U (D), WS2-CF-U (E) and WSe2-CF-U (F). (G) XPS of MoSe2-CF-U, WS2-CF-U, and WSe2-CF-U. (H) FT-IR of MoSe2-CF-U, WS2-CF-U, and WSe2-CF-U.

CONCLUSIONS

In this study, we propose a universal compositing strategy for the preparation of transition metal dichalcogenides composites with CF, resulting in a series of MQ2-collagen fiber (M = Mo, W; Q = S, Se) composite materials designed for efficient uranium adsorption. The composite materials exhibit enhanced hydrophilicity and leverage the fibrous structure of CF to improve uranium adsorption, outperforming the pristine transition metal dichalcogenides. Zeta potential analysis illuminated the self-assembly mechanism between transition metal dichalcogenides and CF, which occurs through electrostatic interactions. Characterization techniques including SEM, EDS, elemental mapping, FT-IR, and XPS revealed the U(VI) adsorption mechanism of the composite materials. The findings suggest that sulfur sites on the surface of MoS2, along with amino and carboxyl groups from CF, work in synergy to coordinate and capture uranium from aqueous solutions. Our study proposes a straightforward method for fabricating transition metal dichalcogenides composite materials, offering a significant boost to U(VI) adsorption performance, and paves the way for the reuse of leather industry waste as functional materials.

MATERIALS AND METHODS

Materials

All chemicals are commercially available products with analytical grade and used without purification. All solutions were prepared using Milli-Q water (18.2 MΩ cm). MoS2 was synthesized based on literature. CF are the rough product from the leather factory (Shenzhen Yongshengye Textile Products Trade Co., Ltd.). (UO2)(NO3)2⋅6H2O (99%, Hubei Jusheng Technology Co., Ltd.), NaCl (AR, Sinopharm Chemical Reagent Co., Ltd), MgCl2 (99%, Adamas Reagent Co., Ltd), KCl (AR, General-Reagent), CaCl2⋅6H2O (98%, Adamas Reagent Co., Ltd), Na2CO3 (Shanghai Hongguang Industry Co., Ltd), 1-butyl-3-methylimidazolium thiocyanate ([BMIM]SCN) (98%, Adamas Reagent Co., Ltd), (NH4)6Mo7O24⋅4H2O (98%, Adamas Reagent Co., Ltd), MoSe2 (99.99%, Adamas Reagent Co., Ltd), WS2 (99.9%, Adamas Reagent Co., Ltd), and WSe2 (99.999%, Adamas Reagent Co., Ltd). MoSe2, WS2, and WSe2 are 2-H phase as verified by powder X-ray diffraction (PXRD) (Figure S27).

Preparation of MoS2 and MQ2-CF composite materials (M = Mo, W; Q = S, Se)

MoS2 was synthesized based on literature [42]. First, 3.50 mmol [BMIM]SCN, 0.05 mmol (NH4)6Mo7O24·4H2O, and 100 μL HCl were dissolved in 10 mL deionized water under vigorous stirring for 30 min. The obtained homogeneous solution was transferred to a 25 mL Teflon-lined autoclave. Then, the autoclave was sealed and heated at 200 °C for 24 h. The obtained black powder sample was washed with deionized water and ethanol for several times, and then dried in a vacuum at 65 °C for 6 h. The black powder sample is 1T phase of MoS2 as proved through scanning electron microscopy (SEM), elemental distribution mappings, energy dispersive spectroscopy (EDS), Raman spectroscopy, and PXRD (Figure S2). The synthesized 1T phase of MoS2 after calcination at 400 °C last for 4 h at N2 atmosphere can transform into 2H phase of MoS2 (Figure S2d).

All MQ2-CF composite materials (M = Mo, W; Q = S, Se) are prepared according to the following steps. In a typical synthesis, 1T phase of MoS2 and CF (a total of 0.2 g) were placed into PTFE ball mill, then added with 2 mL of water and 60 g of zirconium beads with a diameter of 2 millimeters. The ball milling program was 450 r/min and lasted for 16 hours. Then the products were freeze-dried for 24 hours to obtain the MoS2-CF. MoSe2-CF, WS2-CF, and WSe2-CF were prepared through the same method. To demonstrate the assembly mechanism of composite materials, 0.2 g MQ2 (M = Mo, W; Q = S, Se), CF and the mixture of MQ2 and CF were ball milled separately with 2 mL water. The Zeta potential of these samples after ball milling was measured.

Characterization

The surface morphology, EDS, and elemental distribution mappings were characterized by scanning electron microscopy (JEOL JSM-6700F). Fourier transform infrared spectroscopy (FTIR) was conducted in the 400–4000 cm−1 range by using ATR on a Bruker VERTEX 70 FTIR spectrophotometer. Powder X-ray diffraction (PXRD) patterns were recorded by Rigaku Miniflex-II diffractometer at room temperature. X-ray photoelectron spectroscopy (XPS) was performed on a Thermo Fisher ESCALAB 250Xi XPS Spectrometer system. Inductively coupled plasma-mass spectrometry (ICP-MS) was performed on an XSeries II and inductively coupled plasma-optical emission spectrometry (ICP-OES) was studied with a Thermo 7400. The pH values of all solutions were tested by Shanghai Leich E-201F. Contact angle images are recorded on a Zhongchen JC2000DS contact angle meter equipped with a CCD camera. Zeta potential was measured by Zetasizer Nano ZS90 DLS spectrometer (Malvern Inc., UK).

Batch adsorption experiments

In a typical adsorption experiment, composite material was transferred to a 20 mL glass bottle containing aqueous solution of UO2(NO3)2⋅6H2O (m/V = 1 g L−1). The adsorption efficiency (E), adsorption capacity (qe), and distribution ratio (Kd) can be calculated by equations (1), (2), and (3) in supporting information, respectively. The pseudo-first-order kinetic model, pseudo-second-order kinetic model, and intra-particle diffusion model can be calculated by equations (4), (5), and (6) in supporting information, respectively. The Langmuir and Freundlich isotherm adsorption models can be calculated by equations (7) and (8) in supporting information, respectively. The mixture was shaken in a shaker for 12 h at 25 °C and then centrifuged for separation. The composite material after U(VI) adsorption (MoS2-CF-U) was obtained. The supernatant was filtered and diluted for ICP-MS or ICP-OES testing and the solid was washed several times with deionized water drying. The effect of composite ratio, pH, time, U(VI) concentration, coexisting ions, and actual water sample on adsorption were investigated in the study. Moreover, the reuse experiment of MoS2-CF was also researched. The detailed adsorption experiments were shown in SI. At the same time, two parallel experiments were carried out in each experiment to ensure the repeatability of the experiments.

Data availability

The original data are available from corresponding authors upon reasonable request.

Acknowledgments

We are grateful for the technical support from the Comprehensive Training Platform of the Specialized Laboratory (College of Chemistry, Sichuan University) and Analytical & Testing Center of Sichuan University.

Funding

This work was supported by the National Natural Science Foundation of China (U21A20296, 22376149, 22325605, 22176136, U1867205 and 22076185), and the Key Research and Development Program of Sichuan Province, China (2020YFS0070 and 2021YFH0170).

Author contributions

C.L.: Conceptualization, Methodology, Validation, Formal analysis, Investigation, Writing - Original draft, Visualization. L.P.L.: Formal analysis, Methodology, Writing - Review & Editing. J.H.H.: Methodology, Investigation. L.Y.: Methodology. J.T.L.: Methodology. C.Q.X.: Supervision, Project administration, Funding acquisition, Investigation. X.Y.H.: Conceptualization, Supervision, Writing - Review & Editing, Project administration, Funding acquisition. M.L.F.: Conceptualization, Supervision, Writing - Review & Editing, Project administration, Funding acquisition, Resources. All authors have given approval to the final version of the manuscript.

Conflict of interest

The authors declare no conflict of interest.

Supplementary information

Supplementary files provided by the authors. Access here

The supporting information is available online at https://doi.org/10.1360/nso/20240029. The supporting materials are published as submitted, without typesetting or editing. The responsibility for scientific accuracy and content remains entirely with the authors.

References

  • Parsons J, Buongiorno J, Corradini M, et al. A fresh look at nuclear energy. Science 2019; 363: 105. [Article] [Google Scholar]
  • Rhodes R. More nuclear power can speed CO2 cuts. Nature 2017; 548: 281 [CrossRef] [PubMed] [Google Scholar]
  • Abney CW, Mayes RT, Saito T, et al. Materials for the recovery of uranium from seawater. Chem Rev 2017; 117: 13935-14013. [Article] [CrossRef] [PubMed] [Google Scholar]
  • Chen B, Zhang G, Chen L, et al. Visible light driven photocatalytic removal of uranium(VI) in strongly acidic solution. J Hazard Mater 2022; 426: 127851. [Article] [Google Scholar]
  • Liu C, Hsu PC, Xie J, et al. A half-wave rectified alternating current electrochemical method for uranium extraction from seawater. Nat Energy 2017; 2: 17007 [CrossRef] [Google Scholar]
  • Mei D, Liu L, Yan B. Adsorption of uranium (VI) by metal-organic frameworks and covalent-organic frameworks from water. Coord Chem Rev 2023; 475: 214917. [Article] [Google Scholar]
  • Yang L, Zeng X, Tang JH, et al. Rapid and selective uranium adsorption by a low-cost, eco-friendly, and in-situ prepared nano-ZnS/alkali-activated collagen fiber composite. Separation Purification Tech 2024; 333: 125856. [Article] [Google Scholar]
  • Wang S, Ran Y, Lu B, et al. A review of uranium-induced reproductive toxicity. Biol Trace Elem Res 2019; 196: 204-213. [Article] [Google Scholar]
  • Sen N, Darekar M, Sirsat P, et al. Recovery of uranium from lean streams by extraction and direct precipitation in microchannels. Separation Purification Tech 2019; 227: 115641. [Article] [Google Scholar]
  • Fu M, Ao J, Ma L, et al. Uranium removal from waste water of the tailings with functional recycled plastic membrane. Separation Purification Tech 2022; 287: 120572. [Article] [Google Scholar]
  • Chen T, Liu B, Li M, et al. Efficient uranium reduction of bacterial cellulose-MoS2 heterojunction via the synergistically effect of Schottky junction and S-vacancies engineering. Chem Eng J 2021; 406: 126791. [Article] [NASA ADS] [CrossRef] [Google Scholar]
  • Liang P, Yuan L, Du K, et al. Photocatalytic reduction of uranium(VI) under visible light with 2D/1D Ti3C2/CdS. Chem Eng J 2021; 420: 129831. [Article] [CrossRef] [Google Scholar]
  • Xin Q, Wang Q, Luo K, et al. Mechanism for the seleikctive adsorption of uranium from seawater using carboxymethyl-enhanced polysaccharide-based amidoxime adsorbent. Carbohydrate Polyms 2024; 324: 121576. [Article] [Google Scholar]
  • Estes SL, Powell BA. Enthalpy of uranium adsorption onto hematite. Environ Sci Technol 2020; 54: 15004-15012. [Article] [NASA ADS] [CrossRef] [PubMed] [Google Scholar]
  • Wang J, Zhou W, Shi Y, et al. Uranium sorption on oxyhydroxide minerals by surface complexation and precipitation. Chin Chem Lett 2022; 33: 3461-3467. [Article] [Google Scholar]
  • Liao J, Liu P, Xie Y, et al. Metal oxide aerogels: Preparation and application for the uranium removal from aqueous solution. Sci Total Environ 2021; 768: 144212. [Article] [CrossRef] [PubMed] [Google Scholar]
  • Roberts HE, Morris K, Law GTW, et al. Uranium(V) incorporation mechanisms and stability in Fe(II)/Fe(III) (oxyhydr)oxides. Environ Sci Technol Lett 2017; 4: 421-426. [Article] [Google Scholar]
  • Liu Y, Lu Y, Zhang S, et al. Amphiphilic ligand in situ assembly of uranyl active sites and selective interactions of molybdenum disulfide. J Hazard Mater 2023; 442: 130089. [Article] [NASA ADS] [CrossRef] [PubMed] [Google Scholar]
  • Tang JH, Sun HY, Ma W, et al. Recent progress in developing crystalline ion exchange materials for the removal of radioactive ions. Chin J Struct Chem 2020; 39: 2157–2171 [Google Scholar]
  • Tang J, Feng M, Huang X. Metal chalcogenides as ion-exchange materials for the efficient removal of key radionuclides: A review. Fundamental Res 2024; [Article] [Google Scholar]
  • Jun BM, Lee HK, Park S, et al. Purification of uranium-contaminated radioactive water by adsorption: A review on adsorbent materials. Separation Purification Tech 2021; 278: 119675. [Article] [Google Scholar]
  • Lebed PJ, Savoie JD, Florek J, et al. Large pore mesostructured organosilica-phosphonate hybrids as highly efficient and regenerable sorbents for uranium sequestration. Chem Mater 2012; 24: 4166-4176. [Article] [CrossRef] [Google Scholar]
  • Xie Y, Chen C, Ren X, et al. Emerging natural and tailored materials for uranium-contaminated water treatment and environmental remediation. Prog Mater Sci 2019; 103: 180-234. [Article] [Google Scholar]
  • Rani L, Srivastav AL, Kaushal J, et al. Significance of MOF adsorbents in uranium remediation from water. Environ Res 2023; 236: 116795. [Article] [NASA ADS] [CrossRef] [PubMed] [Google Scholar]
  • Ahmad M, Chen J, Yang K, et al. Preparation of amidoxime modified porous organic polymer flowers for selective uranium recovery from seawater. Chem Eng J 2021; 418: 129370. [Article] [Google Scholar]
  • Yuan D, Wang Y, Qian Y, et al. Highly selective adsorption of uranium in strong HNO3 media achieved on a phosphonic acid functionalized nanoporous polymer. J Mater Chem A 2017; 5: 22735-22742. [Article] [Google Scholar]
  • Cheng G, Zhang A, Zhao Z, et al. Extremely stable amidoxime functionalized covalent organic frameworks for uranium extraction from seawater with high efficiency and selectivity. Sci Bull 2021; 66: 1994-2001. [Article] [Google Scholar]
  • Yusran Y, Guan X, Li H, et al. Postsynthetic functionalization of covalent organic frameworks. Natl Sci Rev 2020; 7: 170-190. [Article] [CrossRef] [PubMed] [Google Scholar]
  • Hao M, Chen Z, Liu X, et al. Converging cooperative functions into the nanospace of covalent organic frameworks for efficient uranium extraction from seawater. CCS Chem 2022; 4: 2294-2307. [Article] [CrossRef] [Google Scholar]
  • Yang S, Hua M, Shen L, et al. Phosphonate and carboxylic acid co-functionalized MoS2 sheets for efficient sorption of uranium and europium: Multiple groups for broad-spectrum adsorption. J Hazard Mater 2018; 354: 191-197. [Article] [CrossRef] [PubMed] [Google Scholar]
  • Radisavljevic B, Radenovic A, Brivio J, et al. Single-layer MoS2 transistors. Nat Nanotech 2011; 6: 147-150. [Article] [NASA ADS] [CrossRef] [PubMed] [Google Scholar]
  • Rahman A, Khan MM. Chalcogenides as photocatalysts. New J Chem 2021; 45: 19622-19635. [Article] [CrossRef] [Google Scholar]
  • Ji DB, Hao SX, Fan XQ, et al. A strategy for the preparation of super-hydrophilic molybdenum disulfide composites applied to remove uranium from wastewater. Dalton Trans 2024; 53: 5020-5033. [Article] [CrossRef] [PubMed] [Google Scholar]
  • Kou Y, Zhang L, Liu B, et al. Phosphonate modified MoS2 composite material for effective adsorption of uranium (VI) in aqueous solution. J Radioanal Nucl Chem 2019; 323: 641-649. [Article] [Google Scholar]
  • Wang J, Yang S, Cheng G, et al. The adsorption of europium and uranium on the sodium dodecyl sulfate modified molybdenum disulfide composites. J Chem Eng Data 2020; 65: 2178-2185. [Article] [Google Scholar]
  • Xiao H, Wang Y, Hao B, et al. Collagen fiber-based advanced separation materials: Recent developments and future perspectives. Adv Mater 2022; 34: e2107891. [Article] [CrossRef] [Google Scholar]
  • Li J, Xiao P, Xu Y, et al. Collagen fibril-assembled skin-simulated membrane for continuous molecular separation. ACS Appl Mater Interfaces 2022; 14: 7358-7368. [Article] [Google Scholar]
  • Cheng YM, Sun X, Liao X, et al. Adsorptive recovery of uranium from nuclear fuel industrial wastewater by titanium loaded collagen fiber. Chin J Chem Eng 2011; 19: 592-597. [Article] [Google Scholar]
  • Deng G, Zhang Y, Luo X, et al. Direct extraction of U(VI) from a simulated saline solution by alkali-activated collagen fiber. J Radioanal Nucl Chem 2018; 318: 1109-1118. [Article] [NASA ADS] [CrossRef] [Google Scholar]
  • Tang Y, Zhou J, Luo W, et al. In situ chemical oxidation-grafted amidoxime-based collagen fibers for rapid uranium extraction from radioactive wastewater. Separation Purification Tech 2023; 307: 122826. [Article] [Google Scholar]
  • Liao H, Yu J, Zhu W, et al. Nano-zero-valent Fe/Ni particles loaded on collagen fibers immobilized by bayberry tannin as an effective reductant for uranyl in aqueous solutions. Appl Surf Sci 2020; 507: 145075. [Article] [Google Scholar]
  • Yang J, Xu Q, Zheng Y, et al. Phase engineering of metastable transition metal dichalcogenides via ionic liquid assisted synthesis. ACS Nano 2022; 16: 15215-15225. [Article] [Google Scholar]
  • Yu Y, Nam GH, He Q, et al. High phase-purity 1T′-MoS2- and 1T′-MoSe2-layered crystals. Nat Chem 2018; 10: 638-643. [Article] [CrossRef] [PubMed] [Google Scholar]
  • Han Y, Ma J, Liu D, et al. Microenvironment-modulating adsorption enables highly efficient lithium extraction under natural pH conditions. ACS Nano 2024; 18: 9071-9081. [Article] [Google Scholar]
  • Gao D, Cheng Y, Wang P, et al. An eco-friendly approach for leather manufacture based on P(POSS-MAA)-aluminum tanning agent combination tannage. J Cleaner Production 2020; 257: 120546. [Article] [Google Scholar]
  • Peng L, Zhang X, Guo L, et al. Rapid removal and easy separation recovery of Cs+ and Sr2+ by zirconium molybdopyrophosphate-functionalized collagen fibers. J Cleaner Production 2023; 414: 137653. [Article] [Google Scholar]
  • Lv C, Fu J, Wang SL, et al. Separation and recovery of Th(IV) from rare earth and other cation solutions using pH-responsive ionic liquids at high acidity condition of 1 M HNO3. Hydrometallurgy 2022; 214: 105953. [Article] [Google Scholar]
  • Lv C, Guo SJ, Chen H, et al. Ionic liquid-melamine foam composites for capture of thorium under high acidity conditions. Separation Purification Tech 2024; 328: 125020. [Article] [Google Scholar]
  • Feng ML, Sarma D, Qi XH, et al. Efficient removal and recovery of uranium by a layered organic–inorganic hybrid thiostannate. J Am Chem Soc 2016; 138: 12578-12585. [Article] [CrossRef] [PubMed] [Google Scholar]
  • Yuan Y, Zhao S, Wen J, et al. Rational design of porous nanofiber adsorbent by blow‐spinning with ultrahigh uranium recovery capacity from seawater. Adv Funct Mater 2018; 29: 1805380. [Article] [Google Scholar]
  • Zhou M, Wang S, Liu F, et al. Effective separation and immobilization of uranium(VI) from wastewater using amino and carboxyl groups functionalized covalent organic frameworks. Separation Purification Tech 2024; 338: 126501. [Article] [Google Scholar]

All Figures

thumbnail Figure 1

(A) Illustration of self-assembly preparation of MoS2-CF. (B) Photos of MoS2, CF, and MoS2-CF; the enlarged image is the SEM of MoS2-CF.

In the text
thumbnail Figure 2

Characterization of MoS2-CF. SEM of MoS2 (A), CF (B) and MoS2-CF (C). (D) Elemental distribution mapping of MoS2-CF. (E) XPS of CF, MoS2, and MoS2-CF. (F) Water contact angle of CF, MoS2, and MoS2-CF. (G) Zeta potential of CF, MoS2, and MoS2-CF after ball milling.

In the text
thumbnail Figure 3

Adsorption of MoS2-CF for U(VI). (A) The effect of composite ratio of MoS2 and CF on U(VI) adsorption (C0U = 150 ppm, pH = 5, t = 12 h, m/V = 1 g L−1). (B) The pH effect on adsorption of MoS2-CF (C0U = 200 ppm, t = 12 h, m/V = 1 g L−1). (C) Adsorption kinetics simulation (C0U = 150 ppm, pH = 5, m/V = 1 g L−1). (D) Simulation of adsorption isotherms (C0U = 25–550 ppm, pH = 5, t= 12 h, m/V = 1 g L−1). (E) The effect of different molar ratios coexisting Na+, K+, Mg2+, and Ca2+ on U(VI) adsorption of MoS2-CF (C0U = 10 ppm, pH = 5, t = 12 h, m/V = 1 g L−1). (F) U(VI) adsorption of MoS2-CF in actual water samples (Taken from Fuzhou, Fujian) (C0U = 150 ppm, pH = 5, t = 12 h, m/V = 1 g L−1).

In the text
thumbnail Figure 4

Characterization of MoS2-CF after U(VI) adsorption. (A) SEM and elemental distribution mappings of MoS2-CF-U. (B) XPS of MoS2-CF and MoS2-CF-U. (C) High-resolution XPS spectra of U 4f of MoS2-CF-U. (D) High-resolution XPS spectra of S 2p of MoS2-CF and MoS2-CF-U. (E) High-resolution XPS spectra of N 1s of MoS2-CF and MoS2-CF-U. (F) High-resolution XPS spectra of O 1s of MoS2-CF and MoS2-CF-U. (G) FT-IR of MoS2-CF and MoS2-CF-U.

In the text
thumbnail Figure 5

Characterization of U(VI) adsorption by MoSe2-CF, WS2-CF, and WSe2-CF. (A) Zeta potential of MoSe2, WS2, WSe2, MoSe2-CF, WS2-CF, and WSe2-CF after ball milling. (B) Water contact angle of MoSe2-CF, WS2-CF, and WSe2-CF. (C) Adsorption of uranium by MoSe2, WS2, WSe2, MoSe2-CF, WS2-CF (C0U = 150 ppm, pH = 5, t = 12 h, m/V = 1 g L−1). SEM and elemental distribution mappings of MoSe2-CF-U (D), WS2-CF-U (E) and WSe2-CF-U (F). (G) XPS of MoSe2-CF-U, WS2-CF-U, and WSe2-CF-U. (H) FT-IR of MoSe2-CF-U, WS2-CF-U, and WSe2-CF-U.

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.