3D layer-by-layer amorphous MoSx assembled from [Mo3S13]2- clusters for efficient removal of tetracycline: Synergy of adsorption and photo-assisted PMS activation

Yue Li Minghao Fan Conghui Wang Yanxun Li Xiang Yu Jun Ding Lei Yan Lele Qiu Yongcai Zhang Longlu Wang

Citation:  Yue Li, Minghao Fan, Conghui Wang, Yanxun Li, Xiang Yu, Jun Ding, Lei Yan, Lele Qiu, Yongcai Zhang, Longlu Wang. 3D layer-by-layer amorphous MoSx assembled from [Mo3S13]2- clusters for efficient removal of tetracycline: Synergy of adsorption and photo-assisted PMS activation[J]. Chinese Chemical Letters, 2024, 35(9): 109764. doi: 10.1016/j.cclet.2024.109764 shu

3D layer-by-layer amorphous MoSx assembled from [Mo3S13]2- clusters for efficient removal of tetracycline: Synergy of adsorption and photo-assisted PMS activation

English

  • World Health Organization has identified tetracycline (TC) as a key antibiotic. It is disturbing that the abuse of antibiotics poses a serious threat to ecosystems [1,2]. Generally, adsorption technology can efficiently enrich contaminants by pore-filling, hydrogen bonding, π-π interaction and electrostatic interactions [35]. However, by using only the adsorption technology, the pollutants cannot be degraded. Advanced oxidation processes (AOPs) based on peroxymonosulfate (PMS) activation has been applied to remove refractory organic pollutants [610]. When the O—O bonds of PMS break, the sulfate radicals (SO4•‒) with high oxidation potential (2.5–3.1 eV) and long half-life t1/2 (30–40 µs) can be generated to oxidize the organic pollutants [11]. Meanwhile, further introduction of illumination to assist in activating PMS can produce more active species to improve the degradation efficiency [1216]. Nevertheless, the contribution of pollutant adsorption by catalysts was often overlooked when using photo-assisted PMS activation to degrade contaminants. Researchers demonstrated that the adsorption process can narrow the distance between pollutants and catalysts, which is the crucial step for the PMS activation process [17]. Therefore, considering the complementarity between adsorption and photo-assistant PMS activation, the synthesis of bifunctional materials is a feasible strategy to efficiently remove antibiotics.

    Recently, molybdenum disulfide (MoS2) has been extensively investigated in the field of adsorption and PMS activation due to the typical two-dimensional sandwich structure of S-Mo-S. Li et al. [18] fabricated composite adsorbent of Co3S4–MoS2 with a hollow structure by hydrothermal method. The good TC adsorption performance of Co3S4–MoS2 mainly depended on multiple interactions caused by abundant functional groups, including surface complexation, π-π interaction and hydrophobic effect. Su et al. [19] prepared the MoS2 nanosheets vertically aligned on biochar material, which can efficiently activate PMS. The vertical structure can expose abundant catalytic sites, accelerating the production of reactive species. Although these reported MoS2 based composites have increased number of functional groups and active sites to improve the adsorption and catalytic capabilities, their complex preparation processes and high costs make them unpractical. In addition, the basal planes of MoS2 are inert, resulting in the very limited adsorption and catalytic active sites [20]. Therefore, the construction of single MoSx with bifunction by regulating the coordination environments of Mo-S bonds, which can generate maximum adsorption and catalytic sites, is a tremendous challenge.

    Herein, we developed a simple oil bath heating approach to prepare a 3D layer-by-layer amorphous MoSx (a-MoSx) superstructure by assembling basic unit [Mo3S13]2- clusters. The unique channels and disordered arrangements in a-MoSx can maximize the access to reactants and the exposure of active sites, making it an efficient bifunctional catalyst for synergy of adsorption and photo-assisted PMS activation. As a result, the prepared a-MoSx exhibits far greater adsorption capacity (106.91 mg/g) for TC, in comparison with 1T/2H MoS2 (47.92 mg/g). The TC adsorption mechanism of a-MoSx was explored by several characterization methods, mainly involving pore filling, π-π interaction and hydrogen bonding. Afterwards, the synergistic degradation process was comprehensively studied by triggering the PMS system, and possible degradation pathways of TC were proposed. Moreover, density functional theory (DFT) calculations were employed to explore the adsorption energy of TC and PMS on a-MoSx and corresponding electron transfer. In this work, the construction of bifunctional catalyst with strong adsorption and catalytic abilities provides new insights for future water treatment.

    The chemicals, synthesis method, characterization, processes of adsorption and activation of PMS and DFT calculation are listed in Texts S1-S5 and Eqs. S1-S4 (Supporting information). The standard curve of TC was shown in Fig. S1 (Supporting information).

    The detail preparation process of a-MoSx was shown in Fig. S2 (Supporting information). For comparison, 1T/2H MoS2 sample was synthesized by hydrothermal method. Fig. 1a illustrates the model of the 1T/2H MoS2 flower-like structure. For the prepared a-MoSx, it exhibits an obvious 3D layer-by-layer superstructure formed by assembling [Mo3S13]2− clusters (Fig. 1b). The unique structure has large interlayer spacing, allowing for maximized accessibility to reactants and exposing abundant active sites. For each monolayer, it is assembled of [Mo3S13]2− clusters (Fig. 1c). The disordered polymer chain can produce more unsaturated S ligands (show in the red arrow) to greatly increase the active sites. Fig. 1d and Fig. S3 (Supporting information) demonstrate that the synthesized 1T/2H MoS2 sample with distinct nanoflower structure and 1T/2H mixed phases. The morphology of the prepared a-MoSx is shown by SEM (Fig. 1e) and TEM (Fig. 1f) images. Obviously, the a-MoSx exhibits a wormlike assembling structure, which consists of face-to-face monolayer nanosheet. The corresponding HRTEM images in Figs. 1g and h demonstrate the highly disordered feature of a-MoSx. Notably, the lattice spacing of 0.67 nm comes from the layer-by-layer stacking domains in a-MoSx [21]. The EDS mappings indicate that the elements of Mo and S are presented in the a-MoSx (Fig. S4 in Supporting information). Furthermore, the Raman and XRD characterization results also confirm the successful preparation of 1T/2H MoS2 and a-MoSx (Fig. 1i, Fig. S5 and Text S6 in Supporting information) [2123].

    Figure 1

    Figure 1.  Illustration of (a) 1T/2H MoS2 nanoflower structure, (b) a-MoSx worm-like superstructure and (c) molecular structure of monolayer a-MoSx. SEM of (d) 1T/2H MoS2 and (e) a-MoSx. (f) TEM image of a-MoSx. (g) HRTEM image of a-MoSx and (h) magnified image of white box in (g). (i) Raman patterns of a-MoSx and 1T/2H MoS2.

    The adsorption and catalytic degradation of TC by 1T/2H MoS2 and a-MoSx were compared by adsorption and photo-assistant PMS activation experiments (Fig. 2a). The TC adsorption capacity (23.6%) of 1T/2H MoS2 was mediocre and only 74.6% of TC was eliminated by photo-assistant PMS activation after 60 min. For a-MoSx, the adsorption capacity and removal efficiency of TC was significantly improved to 53.2% and 98.8%, respectively, which may be attributed to the larger layer spacing in a-MoSx and the effective exposure of active sites due to the abundant unsaturated S atom. In addition, it was found that pH 5 was the optimal adsorption condition (Fig. S6 in Supporting information).

    Figure 2

    Figure 2.  (a) Adsorption and photo-assistant PMS activation curves of a-MoSx and 1T/2H MoS2. (b) Adsorption kinetics of different catalysts. (c) N2 adsorption-desorption isotherms of different catalysts, inset shows the pore-size distribution curves. (d) FTIR spectra of a-MoSx, TC-a-MoSx (a-MoSx after adsorption of TC) and 1T/2H MoS2. XPS spectra of (e) Mo 3d and (f) S 2p of a-MoSx before and after TC adsorption. (g) Zeta potentials of a-MoSx at different pH and maximum adsorption capacity of TC. (h) Adsorption energies of TC on the surface of a-MoSx and 2H MoS2. Conditions: [Catalyst] = 0.050 g/L, pH 5.0, [TC] = 10 mg/L.

    The TC adsorption capacities of 1T/2H MoS2 and a-MoSx were shown in Fig. 2b. It can be observed that the a-MoSx presents a superior adsorption capacity (106.91 mg/g) compared to 1T/2H MoS2 (47.92 mg/g). The kinetic models and the isotherms models were used to explore the adsorption process of TC. The maximum adsorption capacity and the related parameters were shown in Tables S1 and S2 in Supporting information. The adsorption process of TC is more suitably described by the pseudo−second−order mode and Langmuir isotherm, suggesting that chemisorption and homogeneous molecular adsorption are the dominant adsorption behavior (Fig. 2b and Fig. S7 in Supporting information) [24]. In addition, the intraparticle diffusion model find that adsorption process is related to multiple steps including film diffusion and intraparticle diffusion (Fig. S8 and Table S3 in Supporting information).

    Subsequently, the Brunauer-Emmett-Teller (BET), Fourier transform infrared spectroscopy (FTIR), X-ray photoelectron spectroscopy (XPS) and zeta potential were used to further explore the adsorption mechanism. As depicted in Fig. 2c and Table S4 (Supporting information), the a-MoSx demonstrates larger surface area, pore volume and average pore size than those of 1T/2H MoS2, which may be attributed to the larger interlayer spacing in a-MoSx. It should be noted that the value of pore size, pore volume and specific surface area all decrease to some extent after adsorbing TC molecule on a-MoSx, suggesting that pore filling is one of the adsorption mechanisms [25]. Fig. 2d and Table S5 (Supporting information) show the FTIR spectra of 1T/2H MoS2, a-MoSx and TC adsorbed a-MoSx (TC-a-MoSx). It can be obviously observed that the positions of O—H, Mo=O and S-H bonds of a-MoSx change after adsorption of TC, resulting from the formation of hydrogen bonding between a-MoSx and TC [25,26]. In addition, the peaks of cumulated double bond, C=O and Mo-S are also shifted to some extent, suggesting that the presence of π-π stacking interaction [26,27]. Furthermore, the chemical valences of a-MoSx and TC-a-MoSx were investigated by XPS analysis and the related parameters were shown in Table S6 (Supporting information). In the full XPS spectra (Fig. S9a in Supporting information), the higher intensity of C 1s and O 1s were observed in TC-a-MoSx, demonstrating that TC were successfully adsorbed on a-MoSx [18]. In Mo 3d XPS spectrum (Fig. 2e), the area percentage of S 2s increase from 13.2% to 14.7% and the positions shift as well after loading of TC, suggesting that there may be interactions between S atoms and TC molecules [18]. Notably, in S 2p spectrum, the positions and area percentages of bridging sulfur S22− (bri-S22−)/apical sulfur S2− (api-S2−) and terminal sulfur S22− (ter-S22−)/unsaturated S2− (uns-S2−) also change after TC adsorption (Fig. 2f), demonstrating that the existence of π-π stacking interactions between the benzene ring of TC molecule and the outer S atoms of a-MoSx [28]. Moreover, in C 1s spectrum (Fig. S9b in Supporting information), the peak position of C=O and C—O shift slightly after TC adsorption, illustrating that oxygen-containing groups are involved in the adsorption process between TC and a-MoSx [28]. In O 1s spectrum (Fig. S9c in Supporting information), the peaks of C-O, H-O and Mo-O shift with amplitudes of 0.17, 0.33 and 0.12 eV and all of their area percentages also change, which may be attributed to the existence of hydrogen bonding between oxygen-containing groups in a-MoSx and TC molecules [28]. Solution pH was a critical environmental factor to alter the species distribution of TC and surface charging state of adsorbent. As can be seen in Fig. 2g and Fig. S10 (Supporting information), there is almost no electrostatic interaction occurring between the negatively charged forms of a-MoSx and TC at different pH values, suggesting that non-electrostatic interactions is the dominant driving force (the detailed information in Text S7 in Supporting information) [28]. DFT calculations were used to verify the adsorption capacity of TC on a-MoSx (Fig. 2h). The adsorption energy of TC on a-MoSx (−1.09 eV) is 1.91 times stronger than that on 1T/2H MoS2 surface (−0.57 eV), demonstrating that the interaction forces and interlayer spacing presented in a-MoSx are more conducive to the adsorption of TC. Above all, the superior adsorption capacity of a-MoSx for TC depends on pore filling, π-π interaction, hydrogen bonding and high adsorption energy, providing strong driving forces for effective adsorption of TC.

    Adsorption technology only transfers contaminants to the surface of catalyst, but cannot effectively degrade pollutants. Therefore, further exploration of synergy of adsorption and phototo-assistant PMS activation by a-MoSx is necessary. As depicted in Fig. 3a, under the synergy of adsorption and photo-assistant PMS activation, the removal efficiency and rate constant k of a-MoSx could reach 96.6% and 0.2245 min−1 within 20 min. These values are much higher than those of only adsorption (43.9% or 0.0334 min−1) and only photo-assistant PMS activation (92.9% or 0.0914 min−1) within 20 min, as shown in Fig. 3b. The faster degradation efficiency and less degradation time are beneficial to saving sewage treatment costs. Moreover, the synergy ability of a-MoSx was obviously higher than that of 1T/2H MoS2, which may be attributed to the presence of more active sites in a-MoSx. To further optimize the experimental parameters, the influence of oxidant kinds, oxidant concentration, catalyst dosage and pollutant on the synergistic system was investigated (Figs. S11-S14 and Text S8 in Supporting information). Fig. 3c shows that 0.5 mmol/L PMS, 0.05 g/L a-MoSx and 10 mg/L TC are the optimized parameters and selected for the subsequent experiments. Table S7 showed other reported bifunctional catalysts with synergistic effects of adsorption and catalysis. It could be found that our prepared a-MoSx had excellent degradability capacity.

    Figure 3

    Figure 3.  (a) Performance of a-MoSx and 1T/2H MoS2 in the adsorption and in synergy of adsorption and photo-assistant PMS activation. (b) Comparison of rate constants k. (c) The reaction rate constants k of optimized processes. The calculated Eads of PMS on (d) 2H MoS2 and (f) a-MoSx. The charge density difference diagrams of PMS adsorbed on (e) 2H MoS2 and (g) a-MoSx. The blue represents gaining electrons and the green represents losing electrons. Conditions: [Catalyst] = 0.050 g/L, pH 5.0, [PMS] = 0.5 mmol/L, [TC] = 10 mg/L.

    Theoretical calculations can be used to further explore the better activation of PMS on a-MoSx than 1T/2H MoS2. As presented in Figs. 3d and f, the adsorption energy of PMS on a-MoSx surface (−1.29 eV) is stronger than that on 1T/2H MoS2 surface (−0.73 eV). Meanwhile, the O–O bond length (lOO) of PMS (1.333 Å) is stretched to 1.388 Å for 1T/2H MoS2 and 1.469 Å for a-MoSx, respectively. The higher adsorption energy and longer lOO length indicate that a-MoSx is more conducive to the dissociation of PMS. Moreover, the transfer of 0.65 and 0.81 electrons from 1T/2H MoS2 and a-MoSx to PMS can be seen from the charge density difference images (Figs. 3e and g), manifesting that a-MoSx with abundant unsaturated S coordination has stronger electron transfer capacity. All the above results are consistent with excellent TC degradation efficiency of a-MoSx.

    In order to explore the valence state change and catalytic mechanism, XPS analysis of a-MoSx before and after catalyzing TC degradation was performed (Fig. S15 in Supporting information). As shown in Fig. S15b, the amount of Mo4+ decreases from 69.69% to 55.38% and the contents of Mo5+ and Mo6+ increase from 17.16% to 33.19% after catalyzing TC degradation, suggesting that the existence of valence cycle of Mo4+/Mo5+/Mo6+, which contributes to the activation process of PMS [29]. Fig. S15c shows that the content of S atom also decreases after catalyzing TC degradation, especially for the bri-S22−. This indicates that more bri-S22− are captured by H+ in the solution, which may be due to the higher catalytic activity of bri-S22−. In addition, the increase in S=O bond content may be due to the participation of S species with reduction ability in the regeneration process of Mo4+ [30]. The detail mechanism will be discussed later. A series of free radical scavenging experiments were conducted to find out the main reactive oxygen species in TC degradation process. Tert-butyl alcohol (TBA) and methanol (MeOH) were commonly selected as the quenching agents of OH and OH + SO4•‒ due to the fast second-order reaction rates (TBA: k•OH = 3.8–7.6 × 108 L mol−1 s−1; MeOH: kSO4•− = 2.5 × 107 L mol−1 s−1, k•OH = 9.7 × 107 L mol−1 s−1) [31]. The L-histidine (L-his), p-benzoquinone (p-BQ), ammonium oxalate (AO) and AgNO3 were used as scavengers for 1O2, O2•−, h+ and e [3235], respectively. The highest inhibition of TC removal at scavenger concentrations of 200 mmol/L MeOH, 200 mmol/L TBA, 1 mmol/L PBQ, 1 mmol/L AO, 1 mmol/L AgNO3 and 20 mmol/ L-his were shown in Fig. S16 (Supporting information) and Fig. 4a. The TC degradation efficiency was slightly reduced to 66.8% and 72.1% after the addition of 200 mmol/L MeOH and TBA, indicating that the contribution of OH is higher than that of SO4•‒. The slight inhibition by AO and AgNO3 manifests that h+ and e contribute little to the TC degradation. On the contrary, the sluggish degradation efficiency after adding PBQ and L-his demonstrate that the dominant role of O2•− and 1O2. EPR experiments were performed to further confirm the reactive species generated in the synergy of adsorption and photo-assistant PMS activation (Figs. 4b-d). The characteristic peaks of DMPO-SO4•‒, DMPO-OH, DMPO-O2•− and TEMP-1O2 were observed after 10 min [3640]. It can be seen that DMPO-SO4•‒ exhibits weak peak intensity, indicating the generation of a small amount of SO4•‒. Furthermore, the DMPO-O2•− and TEMP-1O2 display obviously stronger peak intensity than DMPO-OH and DMPO-SO4•‒, which demonstrates convincingly that the O2•− and 1O2 are the main active species. Electron migration in TC degradation was analyzed by it curves (Fig. 4e). The current output of a-MoSx increased much more sharply compared with that of 1T/2H MoS2 after the introduction of PMS, indicating that more efficient electron transfer occurs from a-MoSx to PMS. Further improvement of the current was observed after the injection of TC, manifesting that the electrons of TC are transferred to high redox potential region [41]. The smaller arc radius of a-MoSx than 1T/2H MoS2 in electrochemical impedance spectroscopy (EIS) curves (Fig. 4f) also indicates the faster electron transfer of a-MoSx, which was consistent with the results of the DFT calculations [42,43].

    Figure 4

    Figure 4.  (a) Effects of radical scavengers on TC degradation by a-MoSx/PMS/light in synergistic system. EPR spectra of (b) SO4•‒ and OH, (c) O2•− and (d) 1O2 in synergy of adsorption and photo-assistant PMS activation by a-MoSx. (e) i−t curves of a-MoSx and 1T/2H MoS2 with adding PMS and TC. (f) EIS curves of a-MoSx and 1T/2H MoS2. (g) Possible degradation pathway of TC by a-MoSx/PMS/light. Conditions for (a): [a-MoSx] = 0.050 g/L, pH 5.0, [PMS] = 0.5 mmol/L, [TC] = 10 mg/L, [TBA] = [MeOH] = 200 mmol/L, [PBQ] = [AO] = [AgNO3] = 1.0 mmol/L, [L-his] = 20 mmol/L.

    Based on the LC-MS spectra (Fig. S17 in Supporting information), the degradation pathway of TC is proposed and presented in Fig. 4g. First, TC was transformed to product P1 (m/z 427) induced by dehydration reaction on C8 site [44]. Then, P2 (m/z 410) was derived from the C—N bond cleavage at N2 site [45]. Meanwhile, P2 generated another intermediate P3 (m/z 318) through the fourth ring break and -OH, -N-CH3 and -CH3 group removal [46]. The C18, C19 and O7 sites were attacked and the molecular chain was further broken to form P4 (m/z 274) [46]. Thereafter, P5 (m/z 230) formed via dihydroxylation at C8 and decarbonylation at C16 [46]. Eventually, the intermediates were further mineralized to CO2 and H2O. Simultaneously, the TOC removal efficiency can reach 56.8% within 30 min in synergistic system (Fig. S18 in Supporting information).

    Overall, as illuminated in Fig. 5 and Text S9 (Supporting information), the degradation mechanism of TC by a-MoSx was proposed, involving the synergy of adsorption and photo-assistant PMS activation process. For the adsorption process, due to its larger layer spacing and more interaction forces, a-MoSx exhibits excellent adsorption performance for TC through pore filling, hydrogen bonding and π-π interactions. Simultaneously, PMS was rapidly attracted by a-MoSx due to the stronger Eads of PMS on a-MoSx. For the photo-assistant PMS process, the numerous unsaturated S atoms in a-MoSx combined with protons in aqueous solution to generate H2S, resulting in the exposure of Mo4+. Then, the adsorbed PMS reacted with Mo4+ to generate SO4•‒ and Mo5+/Mo6+, while Mo5+/Mo6+ were turned back to the Mo4+ with the aid of PMS (Eqs. S4-S6 in Supporting information). Notably, we deduce that reduced S22− can also react with Mo6+ or Mo5+ to promote the regeneration of metal active sites of Mo4+ (Eqs. S7-S10 in Supporting information). After further introduction of illumination, h+ and e can be formed. O2 in solution will react with the e in the conduction band to generate O2•− (Eq. S11 in Supporting information), while h+ reacts with water to generate OH (Eq. S12 in Supporting information). SO4•‒ can also react with H2O to produce OH (Eq. S13 in Supporting information). 1O2 can be produced by the recombination of OH and the hydrolysis of O2•− (Eqs. S14 and S15 in Supporting information). Ultimately, both free and non-free radical pathways participate in the degradation of TC.

    Figure 5

    Figure 5.  Schematic diagram of the TC removal mechanism by a-MoSx in synergy of adsorption and photo-assistant PMS activation.

    Finally, practical application tests were conducted to investigate the adaptability of the synergistic system of a-MoSx/PMS/light. The influence of various anions on TC degradation is shown in Fig. S19 (Supporting information). In the presence of different concentrations of Cl, Br HCO3 and HPO42−, the TC removal efficiency has only a limited drop, indicating that a-MoSx has a strong anti-interference ability. Furthermore, humic acid (HA) was chosen to study the effect of natural organic matters (NOM) on TC degradation (Fig. S20 in Supporting information). The synergistic system still maintains a high degradation efficiency as the concentration of HA increases, suggesting the efficient and stable catalytic process. The cycle experiment in Fig. S21 (Supporting information) shows the removal efficiency of TC maintained at about 95.2% in fifth recycling, manifesting the acceptable stability and reusability of a-MoSx. The concentration of Mo ions dissolved in the solution was measured after the synergistic degradation under the optimized condition. Notably, the dissolved Mo and S of 1T/2H MoS2 are 4.2 mg/L and 10.8 mg/L, while those of a-MoSx are 0.8 mg/L and 8.4 mg/L, respectively (Fig. S22 in Supporting information). The results indicate that a-MoSx with 3D layer-by-layer superstructure has great stability and will not produce noteworthy metal pollution. When applied in different water environments (Longzi lake, Xiliu lake and Yellow river), the removal efficiency of TC in 60 min still reached over 80.6% (Fig. S23 in Supporting information). For the degradation of other organic pollutants, such as TC, norfloxacin (NOR) and bisphenol A (BPA), a-MoSx also exhibits excellent synergistic adsorption and catalytic degradation performance (Fig. S24 in Supporting information). The COD removal efficiency of TC, NOR and BPA solutions reached 78.9%, 61.2%, 64.1% within 60 min, respectively (Fig. S25 in Supporting information). Raman patterns of a-MoSx before and after catalyzing TC degradation (Fig. S26 in Supporting information) display no remarkable change, further confirming its outstanding stability.

    In summary, a-MoSx with 3D layer-by-layer superstructure formed by assembling [Mo3S13]2- clusters was designed and prepared. The unique structure with large interlayer spacing and abundant active sites endows a-MoSx to have an excellent adsorption capacity to TC (106.91 mg/g). As a result, the TC degradation efficiency reached 96.6% within 20 min in the synergistic system. Various characterizations and DFT calculations were used to explore the mechanism of adsorption and PMS activation by a-MoSx. The quenching experiments and EPR analysis discovered that 1O2 and O2•− play the dominatant role while OH and SO4•‒ have a minor role in synergistic system. Furthermore, the degradation pathway of TC was explored by HPLC−MS. Meanwhile, the synergy system exhibits great potential in piratical application and the COD removal efficiency of TC solution could reach 78.9% in 60 min. This study provides ideas for catalyst design with outstanding synergistic adsorption and photo-assistant PMS activation in environment remediation.

    The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

    This work was financially supported by the National Natural Science Foundation of China (Nos. 52370073, 12274115), Program for Science and Technology Innovation Team in Universities of Henan Province (No. 24IRTSTHN017), Natural Science Foundation of Henan Province (No. 212300410336), Program for Science and Technology Innovation Talent in Universities of Henan Province (No. 23HASTIT027), Key Scientific and Technological Project of Henan Province (No. 222102320188), Key Project of Science and Technology Research of Henan Provincial Department of Education (No. 21A430008).

    Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.cclet.2024.109764.


    1. [1]

      Y. Li, B. Yu, B. Liu, et al., Chem. Eng. J. 452 (2023) 139542. doi: 10.1016/j.cej.2022.139542

    2. [2]

      T. Liu, Y. Li, H. Sun, et al., Chin. J. Struct. Chem. 41 (2022) 2206055.

    3. [3]

      A. Mei, W. Chen, Z. Yang, et al., Angew. Chem. Int. Ed. 62 (2023) e202301440. doi: 10.1002/anie.202301440

    4. [4]

      T. Wang, J. He, J. Lu, et al., Chin. Chem. Lett. 33 (2022) 3585–3593. doi: 10.1016/j.cclet.2021.09.029

    5. [5]

      P. Muang-Non, C. Richardson, N.G. White, Angew. Chem. Int. Ed. 62 (2023) e202212962. doi: 10.1002/anie.202212962

    6. [6]

      J. Song, N. Hou, X. Liu, et al., Adv. Mater. 35 (2023) 2209552. doi: 10.1002/adma.202209552

    7. [7]

      J. Yang, M. Zhang, M. Chen, et al., Adv. Mater. 35 (2023) 2209885. doi: 10.1002/adma.202209885

    8. [8]

      M. Wang, F. Wang, P. Wang, et al., Sep. Purif. Technol. 326 (2023) 124806. doi: 10.1016/j.seppur.2023.124806

    9. [9]

      F. Wang, Y. Gao, H. Chu, et al., ACS ES&T Eng. 4 (2024) 153–165.

    10. [10]

      Z.C. Zhang, F.X. Wang, C.C. Wang, et al., Sep. Purif. Technol. 327 (2023) 124944. doi: 10.1016/j.seppur.2023.124944

    11. [11]

      X. Zhang, J. Zhang, B. Hou, et al., Appl. Surf. Sci. 643 (2024) 158663. doi: 10.1016/j.apsusc.2023.158663

    12. [12]

      J. Yu, X. Li, Z. Jin, et al., Chin. J. Struct. Chem. 41 (2022) 2206001–2206002.

    13. [13]

      Y. Guo, C. Xin, L. Dai, et al., Sep. Purif. Technol. 284 (2022) 120297. doi: 10.1016/j.seppur.2021.120297

    14. [14]

      Q. Tang, W. Tu, Y. Zhou, et al., Chin. J. Struct. Chem. 42 (2023) 100170. doi: 10.1016/j.cjsc.2023.100170

    15. [15]

      X. Yu, C. Zhou, Z. Huang, et al., Appl. Catal. B 325 (2023) 122308. doi: 10.1016/j.apcatb.2022.122308

    16. [16]

      S. Li, C. Wang, K. Dong, et al., Chin. J. Catal. 51 (2023) 101–112. doi: 10.1016/S1872-2067(23)64479-1

    17. [17]

      H. Fang, J. Gao, J. Wang, et al., Sep. Purif. Technol. 314 (2023) 123565. doi: 10.1016/j.seppur.2023.123565

    18. [18]

      S. Li, X. Chen, M. Li, et al., Chem. Eng. J. 441 (2022) 136006. doi: 10.1016/j.cej.2022.136006

    19. [19]

      X. Su, Y. Guo, L. Yan, et al., Sep. Purif. Technol. 282 (2022) 120118. doi: 10.1016/j.seppur.2021.120118

    20. [20]

      C. Gu, T. Sun, Z. Wang, et al., Small Methods 7 (2023) 2201529. doi: 10.1002/smtd.202201529

    21. [21]

      D. Wang, H. Li, N. Du, et al., Chem. Eng. J. 398 (2020) 125685. doi: 10.1016/j.cej.2020.125685

    22. [22]

      J. Chen, Z. Chen, W. Zhao, et al., Inorg. Chem. Commun. 152 (2023) 110657. doi: 10.1016/j.inoche.2023.110657

    23. [23]

      X. Hua, H. Chen, C. Rong, et al., J. Hazard. Mater. 448 (2023) 130951. doi: 10.1016/j.jhazmat.2023.130951

    24. [24]

      J. Zhu, Y. Ma, X. Chen, et al., J. Water Process Eng. 49 (2022) 103089. doi: 10.1016/j.jwpe.2022.103089

    25. [25]

      Z. Zeng, S. Ye, H. Wu, et al., Sci. Total Environ. 648 (2019) 206–217. doi: 10.1016/j.scitotenv.2018.08.108

    26. [26]

      X. Yang, L. Wen, H. Huang, et al., Solid. State Sci. 133 (2022) 107014. doi: 10.1016/j.solidstatesciences.2022.107014

    27. [27]

      S. Ye, M. Yan, X. Tan, et al., Appl. Catal. B 250 (2019) 78–88. doi: 10.1016/j.apcatb.2019.03.004

    28. [28]

      S. Li, W. Huang, P. Yang, et al., Sci. Total Environ. 754 (2021) 141925. doi: 10.1016/j.scitotenv.2020.141925

    29. [29]

      L. Wang, X. Zheng, L. Yan, et al., Sep. Purif. Technol. 317 (2023) 123907. doi: 10.1016/j.seppur.2023.123907

    30. [30]

      M. Huang, X. Wang, C. Liu, et al., J. Hazard. Mater. 404 (2021) 124175. doi: 10.1016/j.jhazmat.2020.124175

    31. [31]

      Z.H. Xie, C.S. He, Y.L. He, et al., Water. Res. 232 (2023) 119666. doi: 10.1016/j.watres.2023.119666

    32. [32]

      W. Liu, Y. Lu, Y. Dong, et al., Chem. Eng. J. 466 (2023) 143161. doi: 10.1016/j.cej.2023.143161

    33. [33]

      J. Yao, N. Wu, X. Tang, et al., Chem. Eng. J. Adv. 12 (2022) 100378. doi: 10.1016/j.ceja.2022.100378

    34. [34]

      Z. Chen, Z. Wei, L. Yang, et al., Chem. Eng. J. 475 (2023) 146236. doi: 10.1016/j.cej.2023.146236

    35. [35]

      Y. Li, B. Yu, H. Li, et al., Chin. Chem. Lett. 34 (2023) 107874. doi: 10.1016/j.cclet.2022.107874

    36. [36]

      C. Xin, B. Wang, J. Yang, et al., Ceram. Int. 49 (2023) 37861–37871. doi: 10.1016/j.ceramint.2023.09.114

    37. [37]

      Z. Chen, G. Liu, S. Yu, et al., Chem. Eng. J. 474 (2023) 145581. doi: 10.1016/j.cej.2023.145581

    38. [38]

      T. Ying, W. Liu, L. Yang, et al., Sep. Purif. Technol. 330 (2024) 125272. doi: 10.1016/j.seppur.2023.125272

    39. [39]

      M. Cai, Y. Liu, K. Dong, et al., Chin. J. Catal. 52 (2023) 239–251. doi: 10.1016/S1872-2067(23)64496-1

    40. [40]

      S. Zhang, G. Hu, M. Chen, et al., Appl. Catal. B 330 (2023) 122584. doi: 10.1016/j.apcatb.2023.122584

    41. [41]

      A. Wang, J. Ni, W. Wang, et al., Appl. Catal. B 319 (2022) 121926. doi: 10.1016/j.apcatb.2022.121926

    42. [42]

      L. Xie, L. Wang, X. Liu, et al., Angew. Chem. Int. Ed. (2023), doi: 10.1002/anie.202316306.

    43. [43]

      S. Li, K. Dong, M. Cai, et al., eScience 4 (2024) 100208. doi: 10.1016/j.esci.2023.100208

    44. [44]

      P. Gao, Z. Li, L. Feng, et al., Chem. Eng. J. 429 (2022) 132398. doi: 10.1016/j.cej.2021.132398

    45. [45]

      Q. Si, W. Guo, H. Wang, et al., Appl. Catal. B 299 (2021) 120694. doi: 10.1016/j.apcatb.2021.120694

    46. [46]

      X. Ge, G. Meng, B. Liu, J. Mol. Liq. 364 (2022) 120078. doi: 10.1016/j.molliq.2022.120078

  • Figure 1  Illustration of (a) 1T/2H MoS2 nanoflower structure, (b) a-MoSx worm-like superstructure and (c) molecular structure of monolayer a-MoSx. SEM of (d) 1T/2H MoS2 and (e) a-MoSx. (f) TEM image of a-MoSx. (g) HRTEM image of a-MoSx and (h) magnified image of white box in (g). (i) Raman patterns of a-MoSx and 1T/2H MoS2.

    Figure 2  (a) Adsorption and photo-assistant PMS activation curves of a-MoSx and 1T/2H MoS2. (b) Adsorption kinetics of different catalysts. (c) N2 adsorption-desorption isotherms of different catalysts, inset shows the pore-size distribution curves. (d) FTIR spectra of a-MoSx, TC-a-MoSx (a-MoSx after adsorption of TC) and 1T/2H MoS2. XPS spectra of (e) Mo 3d and (f) S 2p of a-MoSx before and after TC adsorption. (g) Zeta potentials of a-MoSx at different pH and maximum adsorption capacity of TC. (h) Adsorption energies of TC on the surface of a-MoSx and 2H MoS2. Conditions: [Catalyst] = 0.050 g/L, pH 5.0, [TC] = 10 mg/L.

    Figure 3  (a) Performance of a-MoSx and 1T/2H MoS2 in the adsorption and in synergy of adsorption and photo-assistant PMS activation. (b) Comparison of rate constants k. (c) The reaction rate constants k of optimized processes. The calculated Eads of PMS on (d) 2H MoS2 and (f) a-MoSx. The charge density difference diagrams of PMS adsorbed on (e) 2H MoS2 and (g) a-MoSx. The blue represents gaining electrons and the green represents losing electrons. Conditions: [Catalyst] = 0.050 g/L, pH 5.0, [PMS] = 0.5 mmol/L, [TC] = 10 mg/L.

    Figure 4  (a) Effects of radical scavengers on TC degradation by a-MoSx/PMS/light in synergistic system. EPR spectra of (b) SO4•‒ and OH, (c) O2•− and (d) 1O2 in synergy of adsorption and photo-assistant PMS activation by a-MoSx. (e) i−t curves of a-MoSx and 1T/2H MoS2 with adding PMS and TC. (f) EIS curves of a-MoSx and 1T/2H MoS2. (g) Possible degradation pathway of TC by a-MoSx/PMS/light. Conditions for (a): [a-MoSx] = 0.050 g/L, pH 5.0, [PMS] = 0.5 mmol/L, [TC] = 10 mg/L, [TBA] = [MeOH] = 200 mmol/L, [PBQ] = [AO] = [AgNO3] = 1.0 mmol/L, [L-his] = 20 mmol/L.

    Figure 5  Schematic diagram of the TC removal mechanism by a-MoSx in synergy of adsorption and photo-assistant PMS activation.

  • 加载中
计量
  • PDF下载量:  2
  • 文章访问数:  793
  • HTML全文浏览量:  31
文章相关
  • 发布日期:  2024-09-15
  • 收稿日期:  2023-12-06
  • 接受日期:  2024-03-11
  • 修回日期:  2024-02-20
  • 网络出版日期:  2024-03-13
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

/

返回文章