Degradation of chloroxylenol by CoSx activated peroxomonosulfate: Role of cobalt-sulfur ratio

Jiayi Guo Liangxiong Ling Qinwei Lu Yi Zhou Xubiao Luo Yanbo Zhou

Citation:  Jiayi Guo, Liangxiong Ling, Qinwei Lu, Yi Zhou, Xubiao Luo, Yanbo Zhou. Degradation of chloroxylenol by CoSx activated peroxomonosulfate: Role of cobalt-sulfur ratio[J]. Chinese Chemical Letters, 2025, 36(4): 110380. doi: 10.1016/j.cclet.2024.110380 shu

Degradation of chloroxylenol by CoSx activated peroxomonosulfate: Role of cobalt-sulfur ratio

English

  • In recent years, there is a continuous growth of the demand for personal care products with antibacterial effects. Thus, various germicidal ingredients are widely added to hand sanitizers, soaps, and body lotions [1, 2]. As a typical alternative to traditional bactericides (such as triclosan and triclocarban), the antiviral efficacy of chloroxylenol (PCMX) against the novel coronavirus (COVID-19) has been demonstrated, which was significant used in recent years [3, 4]. Notably, PCMX has stable chemical properties, and could reach a certain cumulative concentration in the environment. PCMX (at concentrations of 1.46–5.59 µg/L) has been reported to be detected in influent water from wastewater treatment plants [5, 6]. Despite the low concentrations detected, PCMX can cause significant harm to the aquatic environment and human health due to its biological activity [7].

    Advanced oxidation processes (AOPs) that generate strong oxidizing free radicals (e.g., SO4•- and OH) have received much attention in degrading of organic pollutants in water [8-10]. Transition metals (Co, Fe, Mn, Cu, etc.) are widely used as catalysts for oxidative degradation due to their variable valence states, which active oxidizing agent mainly by single electron transfer [11, 12]. Cobalt-based catalysts are considered to be the most effective catalysts among various transition metal [13, 14]. Reduced sulfur species can accelerate the M(II)/M(III) cycle, thereby further improving the production and conversion efficiency of related reactive oxygen species (ROS) and improving catalytic performance [15, 16].

    The efficient degradation of organic pollutants in wastewater by cobalt sulfide-activated peroxymonosulfate (PMS) has attracted much attention [17, 18]. Previous studies have shown that different crystal types of CoSx affect the catalytic degradation performance, and the design and modulation of cobalt-based catalysts with abundant active sites is an effective strategy to obtain efficient cobalt-based catalysts [19], [20]. However, there are few reports investigating the effects of different cobalt-sulfur molar ratios on cobalt sulfide catalysts with tunable morphology and crystal structure, while the relationship between the physical properties of different CoSx and their catalytic reaction efficiencies is still poorly understood. Therefore, the aim of this work is to investigate the effect of precursor molar ratio on the physical properties of CoSx and the structure-performance conformational relationship for PMS activation.

    In this work, CoSx with three different cobalt-sulfur molar ratios were fabricated and characterized by a one pot method. The degradation performance of PCMX by CoSx activated PMS was systematically evaluated, including process parameters and reaction mechanism combined with density functional theory (DFT), the effect of cobalt sulfide ratio on catalyst reactivity were systematically studied.

    All detailed chemicals, experimental and characterization methods are provided in Text S1 (Supporting information).

    Three kinds of cobalt sulfide were simply synthesized by changing the ratio of cobalt to sulfur in the synthesis process. The crystal structure of CoSx was characterized by X-ray diffractometer (XRD). The XRD patterns were well matched with CoS2 (PDF #89-3056), Co3S4 (PDF #02-1361), and Co9S8 (PDF #75-2023), respectively. It indicates that CoSx with different crystal structures were successfully synthesized by facile varying cobalt-sulfur ratio (Figs. 1ac). The crystal structure of CoS2, Co3S4, and Co9S8 is shown in Fig. S1 (Supporting information).

    Figure 1

    Figure 1.  XRD patterns, FESEM images, and TEM images of (a, d, e) CoS2, (b, f, g) Co3S4, and (c, h, i) Co9S8.

    The scanning electron microscope (SEM) and transmission electron microscope (TEM) images show that CoS2 appears as nanospheres (Figs. 1d and e), Co3S4 appears as hollow nanospheres (Figs. 1f and g), and Co9S8 exhibits nanoflowers morphology (Figs. 1h and i). The results show that the controllable preparation of CoSx morphology can be achieved by adjusting cobalt-sulfur ratio. Inductive coupled plasma-optical emission spectrometry (ICP-OES) (Table S1 in Supporting information) turns out that the true cobalt-sulfur molar ratios of CoS2, Co3S4, and Co9S8 are 1:2.0, 1:1.5, and 1:1.1, respectively. The Brunauer-Emmett-Teller (BET) surface area and pore structure of catalysts were analyzed using N2 adsorption-desorption isotherms and BJH pore size distribution (Fig. S2 in Supporting information), which indicates the mesoporous structures of both three catalysts. The above results conform to their theoretical ratios and provide further evidence for the successful preparation of CoSx nanocatalyst as well. BET surface area, pore volume, and pore size of the three catalysts are shown in Table S2 (Supporting information). Obviously, the controllable preparation of CoSx morphology and the corresponding crystal structure can be realized by adjusting the ratio of cobalt(II) acetate tetrahydrate and thiourea.

    Fig. 2a shows the effects of different catalysts CoSx on PCMX degradation via PMS activation. The adsorption ability of the pure CoS2, Co3S4, or Co9S8 on PMCX was negligible (3.1%, 3.9%, and 1.3%, respectively). This finding is consistent with those of Lu et al. and Li et al. who found that pure PMS and pure CoSx had a weak adsorption capacity for pollutants [21, 22]. The degradation efficiencies of CoS2, Co3S4, or Co9S8 on PCMX by PMS activation were 67.7%, 88.7% and 100%, respectively, which indicated that Co9S8 exposed more PMS activation sites than Co3S4 and CoS2 under the same catalyst dosage conditions. It has been reported that Co2+surf. exposed on the surface of cobalt-based catalysts could participate in the activation of PMS, which significantly enhanced the degradation efficiency [23].

    Figure 2

    Figure 2.  (a) PCMX removal curves in different systems. (b) The pseudo-first-order kinetic rate plots of PCMX degradation in CoS2/PMS, Co3S4/PMS, and Co9S8/PMS systems. (c) PMS residual concentration in each system. (d) Utilization efficiency of different cobalt sulfides. Conditions: [Cat.]0 = 0.20 g/L, [PCMX]0 = 20 mg/L, [PMS]0 = 1.00 mmol/L, pH0 = 6.6, T = 20 ± 1 ℃).

    The kinetic simulation results showed that the degradation behavior of PCMX was better fitted with the pseudo-first-order kinetic model (Table S3 in Supporting information). The degradation rate of PCMX was obviously accelerated with the increase of cobalt-sulfur ratio (Fig. 2b). The reaction rate constants of PCMX degradation in CoS2/PMS, Co3S4/PMS, and Co9S8/PMS were 0.0386 min-1, 0.1619 min-1 and 2.3026 min-1, respectively. The decomposition efficiency of PMS by CoSx with time was in agreement with that of PCMX degradation in CoSx/PMS systems, as shown in Fig. 2c. CoS2 had a relatively weak activation ability for PMS, with a decomposition efficiency of 57.6% at 30 min, while Co3S4 and Co9S8 were able to decompose PMS almost completely at 30 min and 5 min, respectively. It may be attributed to the lower sulfation degree of Co9S8, which facilitates the exposure of its surface active sites (Co2+surf.) to promote the decomposition of PMS.

    To further determine the catalytic activity of the materials, the catalyst utilization efficiency (ηCat.), the molar mass of pollutant degraded per gram of catalyst per minute, was calculated (Eq. 1). The calculation results are shown in Fig. 2d. The CoSx nanocatalysts prepared in this study showed high degradation activity for PCMX compared to previously reported cobalt-based catalysts (Table S4 in Supporting information). Among the CoSx, Co9S8 could completely remove PCMX within 3 min, and its ηCat. reached 1.4703 mmol g-1 min-1.

    (1)

    In order to further explore the reactivity of catalysts with different crystal forms caused by cobalt-sulfur ratio, DFT calculations were performed on reaction energy levels and electronic properties [24]. By calculating (energy level gap between catalyst and PMS), it can be seen that the energy level difference between CoS2 and PMS is narrower (), which proves that PMS can more easily receive electrons transferred from CoS2 (Fig. 3a) [25-27]. This (HOMO-LUMO gap) is similar to , and is significantly lower than CoS2, indicating that the increase of cobalt to sulfur ratio makes the catalyst's reactivity significantly enhanced.

    Figure 3

    Figure 3.  (a) LUMO, HOMO and LUMO-HOMO gap energy levels of CoSx and PMS. (b) Electrostatic potential distribution of CoSx; Electron density differences of different PMS adsorption models. Isosurface contour is 0.005 e/bohr3. (c) Density of states of CoSx. (d) The corresponding reaction energy barriers for PMS activation.

    Electrostatic potential distribution (ESP) shows that there are great differences in the charge distribution on the surface of the catalyst with different cobalt-sulfur ratios (Fig. 3b). The results show that this cobalt-sulfur ratio modulation increases the number of electronic sites on the catalyst surface, which can effectively control the charge distribution on the catalyst surface and create more sites for PMS adsorption and activation [28-30]. Furthermore, the electron density difference indicates an obvious charge transfer between CoSx and PMS molecules can be found, which indicates that the coordination adsorption and charge transfer between PMS and CoSx contribute to the activation of PMS [28, 31].

    Besides, density of states analysis reveals covalent hybridization between Co and S orbitals (Fig. 3c) [32], with the d orbitals of CoSx being the main component of the valence ban. By simulating the reaction energy barriers for the CoSx-catalyzed PMS activation process, the results shown in Fig. 3d indicate that with the increase of cobalt-sulfur ratio, the catalyst surface is more likely to adsorb the PMS molecules to complete the electron transfer in the initial state [31, 33]. And the energy required for the peroxy bond breaking of PMS catalyzed by CoS2 is obviously larger than that of Co3S4 and Co9S8, so the CoS2/PMS system has the worst reactivity. Overall, the increase of cobalt-sulfur ratio was more favorable to expose adsorption-catalytic sites and activate the PMS process.

    In the following discussion, using the Co9S8/PMS system as a representative, the effect of other process parameters on the degradation of PCMX by CoSx activated PMS will be investigated. The effects of PMS dosage and initial pH on the catalytic performance of Co9S8/PMS system were investigated (Fig. S3 in Supporting information). With the dosage of 1.00 mmol/L PMS, PCMX was degraded below the detection limit within 5 min. It indicates that there were enough active sites on the catalyst surface, and PMS was activated to generate enough ROS to degrade PCMX (Fig. S3a in Supporting information) [34]. When the initial pH value is in the range of 3.0–9.0, the pH changes little in the degradation process and has a certain stability (Fig. S3b in Supporting information).

    Quenching experiments and EPR techniques revealed the reactive species in the Co9S8/PMS system (Fig. 4). Fig. 4a showed that 200 mmol/L of MeOH and TBA inhibited approximate 87.5% and 17.4% of PCMX degradation, respectively, compared to the control. It indicates that a large amount of SO4•- was produced in the system, while a certain amount of OH was present. CF on the Co9S8/PMS system was negligible, indicating that the content of O2- in the system is very small, or the generated O2- immediately transforms into other oxygen forms [35]. FFA can quench SO4•-, OH and 1O2 simultaneously [36, 37]. The degradation of PCMX could be inhibited almost completely by adding 10 mmol/L FFA, providing evidence to the presence of a large amount of 1O2 in the system. DMPO and TEMP spin traps were used for EPR analysis, and distinct signal peaks of 1:2:2:1 and 1:1:1:1:1:1 were detected in the Co9S8/PMS system, and the 1O2 triplet peaks were significantly stronger than that in PMS (Figs. 4bd). In agreement with the quenching experiments, the presence of SO4•-, 1O2 and OH in the Co9S8/PMS system was further verified.

    Figure 4

    Figure 4.  (a) The effect of scavengers on PCMX degradation. EPR spectra of (b) DMPO-SO4•- and DMPO-OH, (c) DMPO-O2-, and (d) TEMP-1O2 in Co9S8/PMS system. Conditions: [Co9S8]0 = 0.20 g/L, [PCMX]0 = 20 mg/L, [PMS]0 = 1.00 mmol/L, pH0 = 6.6, T = 20 ± 1 ℃.

    The changes of element valance of Co9S8 before and after the degradation reaction were analyzed (after 5 cycles) (Figs. 5a and b). Table S5 (Supporting information) shows the changes in the valence content of Co and S before and after the degradation reaction. After the reaction, the content of Co2+ decreased by 23.0%, indicating that the main active site of PMS activation was Co2+surf., which was oxidized to Co3+surf. after the reaction (Eq. 2) [17]. At the same time, Co3+surf. generated by oxidation would be reduced to Co2+surf. by PMS (Eq. 3) [38]. OH in the system was generated by the reaction of SO4•- with H2O or OH- (Eqs. 4 and 5) [39]. The content of S2- decreased by 12.4% after the reaction. S2- has been reported to reduce Co3+surf. to Co2+surf. (Eq. 6), thereby enhancing the Co2+surf./Co3+surf. cycle efficiency and promoting the improvement of the catalytic efficiency [23]. A small fraction of 1O2 in the system is produced by self-decomposition of PMS (Eq. 7) [21]. The majority of 1O2 is formed by the reaction of accumulated SO5•- with H2O (Eq. 8) [40]. Finally, ROS such as SO4•-, OH, and 1O2 were produced in the system to oxidatively degrade PCMX to generate CO2, H2O, and degradation intermediates (Eq. 9).

    (2)

    (3)

    (4)

    (5)

    (6)

    (7)

    (8)

    (9)

    Figure 5

    Figure 5.  XPS spectra of Co9S8 for (a) Co 2p and (b) S 2p before and after use. (c) Schematic illustration of mechanism for PCMX removal via Co9S8/PMS system.

    The above discussion indicates that the main ROS in Co9S8/PMS system are SO4•-, OH, and 1O2 (Fig. 5c). The main active site of the catalyst is Co2+surf.. The reduced sulfur species S2- promotes the regeneration of Co2+surf. on the catalyst surface, improving the catalytic efficiency.

    To study the transformation and evolution of PCMX in the Co9S8/PMS system, the active site of contaminant PCMX susceptible to ROS attack was verified by DFT calculations (Fig. S4 in Supporting information) [40]. The HOMO and LUMO values of PCMX are -6.110 eV and -0.097 eV, respectively. the value of the energy gap (ΔE = ELUMO - EHOMO) between HOMO and LUMO is 6.013 eV. The Fukui function (f +, f -, f 0) can be introduced to better investigate the oxidizing ability of each atom towards persulfate, catalysts and pollutants [41]. Based on the above calculations, it is considered that ROS such as SO4•-, 1O2, OH are more likely to undergo ring opening reaction in directly attacking the benzene ring at the position of C3-C4/C6-C7 versus C4-C5/C5-C6. Also, ROS can undergo side chain oxidation at C8 and C9. Combined with LCMS results (Fig. S5 in Supporting information), the degradation path of PCMX by Co9S8/PMS system was analyzed, as shown in Fig. S6 (Supporting information).

    In order to reveal the effect of cobalt-sulfur ratio on the catalytic efficiency of the CoSx/PMS system, the relationship between n(Co/S) and BET surface area was fitted. The result indicates that n(Co/S) was positively correlated with BET surface area (Fig. 6a). Meanwhile, BET surface area was in a good linear relationship with catalytic reaction rate constant and the correlation coefficient R2 was 0.999 (Fig. 6b), indicating that the cobalt-sulfur ratio plays a key role in the PMS activation performance of cobalt sulfides, and the catalytic efficiency of cobalt sulfides increased with the cobalt-sulfur ratio. The difference in cobalt content on the surface of catalyst was analyzed by XPS to further explore the effect of BET surface area on the catalytic efficiency of catalysts. Fig. 6c shows the Co 2p spectra of CoS2, Co3S4, and Co9S8 before reaction. The peaks at approximate 778.3, 781.0, 793.4, and 797.1 eV were assigned to be Co3+ 2p3/2, Co2+ 2p3/2, Co3+ 2p1/2, and Co2+ 2p1/2 [42], respectively.

    Figure 6

    Figure 6.  (a) The relationship between n(Co/S) and BET surface area of CoSx. (b) The relationship between BET surface area and reaction rate constant k of CoSx. (c) Co 2p spectra of CoS2, Co3S4, and Co9S8 before reaction. (d) The relationship between content of Co2+ and reaction rate constant k of CoSx. (e) Schematic graph of the effect of cobalt-sulfur ratio on CoSx/PMS system.

    We found that the content of Co2+ in the catalyst was increasing with the cobalt-sulfur ratio, while the content of Co3+ was decreasing. The proportions of Co2+ in CoS2, Co3S4, and Co9S8 were 41.1%, 42.9%, and 57.5%, respectively. However, the proportions of Co3+ were 41.4%, 32%, and 12.8%, respectively. The relationship between the content of Co2+ and reaction rate constant k of each catalytic system was further determined at the same time. Fig. 6d reveals that the content of Co2+ is positively related to reaction rate constant k with a linear relationship. Therefore, the enhanced catalytic activity of cobalt sulfides can be attributed to the increase of Co2+ content on catalysts.

    According to Fig. 6e, the cobalt-sulfur ratio can affect the physical properties of cobalt sulfides, such as morphology, BET surface area and surface Co2+ content, thus determining the catalytic activity of cobalt sulfides. With the increase of cobalt-sulfur ratio, the morphology of cobalt sulfides changes from nanoparticle to nanoflower, and the BET surface area of the catalyst increases gradually. The active site Co2+ on the catalyst surface thus can be more fully exposed, so as to improve the catalytic activity.

    In this work, a series of CoSx with different cobalt-sulfur ratio were successfully fabricated via one plot method for PCMX removal. The cobalt-sulfur ratio is a key factor in regulating the catalytic activity of CoSx through the changes in charge properties and the exposure of active sites. With the reduction of the degree of vulcanization, the prepared cobalt sulfide has a larger BET surface area, and the active site (Co2+surf.) on its surface can be more fully exposed. The results shows that the first-order kinetic constant of PCMX degradation by Co9S8 was 2.30 min-1, which is much higher than that of CoS2 (0.039 min-1) and Co3S4 (0.16 min-1), and the value of k was highly linearly related to the specific surface area of CoSx and the Co2+ content. DFT theoretical calculations also verified that the increase of cobalt-sulfur ratio can expose more active sites and reduce the energy barrier of catalytic oxidation reaction. This work provides both theoretical and technical support for the design of new cobalt-based catalysts for PMS activation.

    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.

    Jiayi Guo: Writing – review & editing, Writing – original draft, Software, Data curation, Conceptualization. Liangxiong Ling: Writing – original draft, Conceptualization. Qinwei Lu: Writing – review & editing, Software. Yi Zhou: Writing – review & editing. Xubiao Luo: Writing – original draft. Yanbo Zhou: Writing – review & editing, Funding acquisition, Conceptualization.

    This work was supported by the National Natural Science Foundation of China (No. 52370168) and the Program of Shanghai Outstanding Technology Leaders (No. 20XD1433900).

    Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.cclet.2024.110380.


    1. [1]

      K. Lv, L. Ling, Q. Lu, et al., Sep. Purif. Technol. 344 (2024) 127207. doi: 10.1016/j.seppur.2024.127207

    2. [2]

      J. Lu, Y. Zhou, L. Ling, et al., Chem. Eng. J. 446 (2022) 137067. doi: 10.1016/j.cej.2022.137067

    3. [3]

      J. Tan, H. Kuang, C. Wang, et al., Sci. Total Environ. 786 (2021) 147524. doi: 10.1016/j.scitotenv.2021.147524

    4. [4]

      M. Ijaz, K. Whitehead, V. Srinivasan, et al., Am. J. Infect. Control. 48 (2020) 972-973. doi: 10.1016/j.ajic.2020.05.015

    5. [5]

      W. Li, H.G. Guo, C. Wang, et al., Chem. Eng. J. 390 (2020) 124610. doi: 10.1016/j.cej.2020.124610

    6. [6]

      Y. Zhang, T. Wang, Q. Lu, et al., Chem. Eng. J. 482 (2024) 149002. doi: 10.1016/j.cej.2024.149002

    7. [7]

      C. Au, K. Chan, W. Chan, et al., J. Hazard. Mater. 445 (2023) 130550. doi: 10.1016/j.jhazmat.2022.130550

    8. [8]

      J. Lu, Q. Liu, Y. Zhang, et al., Chin. Chem. Lett. 35 (2024) 109406. doi: 10.1016/j.cclet.2023.109406

    9. [9]

      L. Di, T. Wang, Q. Lu, et al., Sep. Purif. Technol. 339 (2024) 126740. doi: 10.1016/j.seppur.2024.126740

    10. [10]

      J. Lu, Q. Lu, L. Di, et al., Chin. Chem. Lett. 34 (2023) 108357. doi: 10.1016/j.cclet.2023.108357

    11. [11]

      Y. Li, H. Dong, L. Li, et al., Water Res. 192 (2021) 116850. doi: 10.1016/j.watres.2021.116850

    12. [12]

      P. Wang, H. Zhang, Z. Wu, et al., Chin. Chem. Lett. 34 (2023) 108722. doi: 10.1016/j.cclet.2023.108722

    13. [13]

      Y. Zhang, J. Liu, A. Moores, et al., Environ. Sci. Technol. 54 (2020) 4631-4640. doi: 10.1021/acs.est.9b07113

    14. [14]

      G. Anipsitakis, D. Dionysiou, Environ. Sci. Technol. 38 (2004) 3705-3712. doi: 10.1021/es035121o

    15. [15]

      J. Xu, D. Wang, D. Hu, et al., Front. Env. Sci. Eng. 18 (2023) 37.

    16. [16]

      Y. Zhou, L. Zhou, Y. Zhou, et al., Appl. Catal. B Environ. 279 (2020) 119365. doi: 10.1016/j.apcatb.2020.119365

    17. [17]

      Y. Han, Y. Yang, W. Liu, et al., Front. Env. Sci. Eng. 18 (2024) 9. doi: 10.1007/s11783-024-1769-6

    18. [18]

      J. Xu, D. Wang, D. Hu, et al., Front. Env. Sci. Eng. 18 (2024) 37. doi: 10.1007/s11783-024-1797-2

    19. [19]

      Q. Yan, J. Zhang, M. Xing, Cell Rep. Phys. Sci. 1 (2020) 100149. doi: 10.1016/j.xcrp.2020.100149

    20. [20]

      L. Song, S. Deng, C. Bian, et al., Front. Env. Sci. Eng. 17 (2023) 96. doi: 10.1007/s11783-023-1696-y

    21. [21]

      J. Lu, Y. Zhou, Y.B. Zhou, Chem. Eng. J. 422 (2021) 130126. doi: 10.1016/j.cej.2021.130126

    22. [22]

      W. Li, S. Li, Y. Tang, et al., J. Hazard. Mater. 389 (2020) 121856. doi: 10.1016/j.jhazmat.2019.121856

    23. [23]

      F. Wang, H. Fu, F. Wang, et al., J. Hazard. Mater. 423 (2022) 126998. doi: 10.1016/j.jhazmat.2021.126998

    24. [24]

      P. Zhang, P. Zhou, J. Peng, et al., Water Res. 219 (2022) 118626. doi: 10.1016/j.watres.2022.118626

    25. [25]

      G. Pan, J. Wei, M. Xu, et al., J. Hazard. Mater. 445 (2023) 130479. doi: 10.1016/j.jhazmat.2022.130479

    26. [26]

      Q. Wu, Y. Zhang, H. Liu, et al., Water Res. 224 (2022) 119022. doi: 10.1016/j.watres.2022.119022

    27. [27]

      R. Zheng, Q. Lin, L. Meng, et al., Sep. Purif. Technol. 298 (2022) 121619. doi: 10.1016/j.seppur.2022.121619

    28. [28]

      J. Yang, D. Zeng, J. Li, et al., Chem. Eng. J. 404 (2021) 126376. doi: 10.1016/j.cej.2020.126376

    29. [29]

      L. Wang, H. Dong, Z. Guo, et al., J. Phys. Chem. C. 120 (2016) 17427-17434. doi: 10.1021/acs.jpcc.6b04639

    30. [30]

      M. Feng, J.C. Baum, N. Nesnas, et al., Environ. Sci. Technol. 53 (2019) 2695-2704. doi: 10.1021/acs.est.8b06535

    31. [31]

      X. Li, X. Huang, S. Xi, et al., J. Am. Chem. Soc. 140 (2018) 12469-12475. doi: 10.1021/jacs.8b05992

    32. [32]

      S. Pang, C. Zhou, Y. Sun, et al., J. Clean Prod. 414 (2023) 137569. doi: 10.1016/j.jclepro.2023.137569

    33. [33]

      Y. Zou, J. Hu, B. Li, et al., Appl. Catal. B Environ. 312 (2022) 121408. doi: 10.1016/j.apcatb.2022.121408

    34. [34]

      D. Gao, Y. Lu, Y. Chen, et al., Appl. Catal. B Environ. 309 (2022) 121234. doi: 10.1016/j.apcatb.2022.121234

    35. [35]

      Q. Yi, J. Ji, B. Shen, et al., Environ. Sci. Technol. 53 (2019) 9725-9733. doi: 10.1021/acs.est.9b01676

    36. [36]

      Y. Zhou, J. Jiang, Y. Gao, et al., Environ. Sci. Technol. 49 (2015) 12941-12950. doi: 10.1021/acs.est.5b03595

    37. [37]

      X. Cheng, H.G. Guo, Y.L. Zhang, et al., Water Res. 113 (2017) 80-88. doi: 10.1016/j.watres.2017.02.016

    38. [38]

      X. Dong, X. Duan, Z. Sun, et al., Appl. Catal. B Environ. 261 (2020) 118214. doi: 10.1016/j.apcatb.2019.118214

    39. [39]

      Q. Lu, L. Di, Y. Zhou, et al., Sep. Purif. Technol. 353 (2025) 128099. doi: 10.1016/j.seppur.2024.128099

    40. [40]

      Y. Gu, T. Gao, F. Zhang, et al., Chin. Chem. Lett. 33 (2022) 3829-3834. doi: 10.1016/j.cclet.2021.12.004

    41. [41]

      R. Parr, W. Yang, J. Am. Chem. Soc. 106 (1984) 4049-4050. doi: 10.1021/ja00326a036

    42. [42]

      R. Guo, Y. Chen, Y. Yang, et al., Chin. Chem. Lett. 34 (2023) 107837. doi: 10.1016/j.cclet.2022.107837

  • Figure 1  XRD patterns, FESEM images, and TEM images of (a, d, e) CoS2, (b, f, g) Co3S4, and (c, h, i) Co9S8.

    Figure 2  (a) PCMX removal curves in different systems. (b) The pseudo-first-order kinetic rate plots of PCMX degradation in CoS2/PMS, Co3S4/PMS, and Co9S8/PMS systems. (c) PMS residual concentration in each system. (d) Utilization efficiency of different cobalt sulfides. Conditions: [Cat.]0 = 0.20 g/L, [PCMX]0 = 20 mg/L, [PMS]0 = 1.00 mmol/L, pH0 = 6.6, T = 20 ± 1 ℃).

    Figure 3  (a) LUMO, HOMO and LUMO-HOMO gap energy levels of CoSx and PMS. (b) Electrostatic potential distribution of CoSx; Electron density differences of different PMS adsorption models. Isosurface contour is 0.005 e/bohr3. (c) Density of states of CoSx. (d) The corresponding reaction energy barriers for PMS activation.

    Figure 4  (a) The effect of scavengers on PCMX degradation. EPR spectra of (b) DMPO-SO4•- and DMPO-OH, (c) DMPO-O2-, and (d) TEMP-1O2 in Co9S8/PMS system. Conditions: [Co9S8]0 = 0.20 g/L, [PCMX]0 = 20 mg/L, [PMS]0 = 1.00 mmol/L, pH0 = 6.6, T = 20 ± 1 ℃.

    Figure 5  XPS spectra of Co9S8 for (a) Co 2p and (b) S 2p before and after use. (c) Schematic illustration of mechanism for PCMX removal via Co9S8/PMS system.

    Figure 6  (a) The relationship between n(Co/S) and BET surface area of CoSx. (b) The relationship between BET surface area and reaction rate constant k of CoSx. (c) Co 2p spectra of CoS2, Co3S4, and Co9S8 before reaction. (d) The relationship between content of Co2+ and reaction rate constant k of CoSx. (e) Schematic graph of the effect of cobalt-sulfur ratio on CoSx/PMS system.

  • 加载中
计量
  • PDF下载量:  1
  • 文章访问数:  111
  • HTML全文浏览量:  6
文章相关
  • 发布日期:  2025-04-15
  • 收稿日期:  2024-05-22
  • 接受日期:  2024-08-28
  • 修回日期:  2024-07-23
  • 网络出版日期:  2024-08-29
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