Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Efficient peroxymonosulfate activation by magnesium-doped Co3O4 for thiacloprid degradation: regulation of Co2+/Co3+ ratios and degradation mechanism

Hui Fui*ab, Xinran Mab, Yiping Huangb, Shiyao Xib, Zhandong Renab and Yuchan Zhuab
aHubei Province Key Laboratory of Agricultural Waste Resource Utilization, Wuhan Polytechnic University, Wuhan 430023, China. E-mail: feihui509@163.com
bSchool of Chemistry and Environmental Engineering, Wuhan Polytechnic University, Wuhan 430023, China

Received 10th June 2025 , Accepted 21st August 2025

First published on 28th August 2025


Abstract

AS a low-cost and high-performance catalyst, spinel cobalt oxide (Co3O4) has two different catalytic active sites (tetrahedral Co2+ and octahedral Co3+) to drive the activation of peroxymonosulfate (PMS) through Co2+/Co3+ redox cycle. Tuning Co2+/Co3+ atomic ratio on the surface of Co3O4 for the construction of a synergy in the Co2+/Co3+ redox cycle might be an effective way to further boost PMS activation performance of Co3O4 catalyst. Herein, we suggested a metal-doping strategy to regulate Co2+/Co3+ atomic ratio of Co3O4 by partially substituting Co2+ with inert Mg2+ and formed a series of Mg doped Co3O4 (MCO) catalysts. Structural characterizations and experimental investigations demonstrated that Mg doping did not change Co3O4 host lattice and particle morphology, but could manipulate surface Co2+/Co3+ atomic ratio of Co3O4 for an improved PMS activation. The optimal MCO catalysts (MCO-0.2) with the suitable Co2+/Co3+ atomic ratios (1.13) exhibited the excellent thiacloprid (THIA) degradation performance through PMS activation, and the apparent degradation rate constant (0.2835 min−1) was highly outperformed that of pure Co3O4 (0.09555 min−1) and other similar cobalt-based catalysts. The optimal THIA degradation conditions might be: catalyst dose 100 mg L−1, PMS concentration 0.8 mM, pH 7 and THIA concentration 20 mg L−1. Quenching experiments and electron paramagnetic resonance (EPR) characterizations suggested SO4˙, HO˙ and 1O2 were all involved in THIA degradation during the MCO-0.2/PMS process. Furthermore, the steady-state concentrations of these reactive species and their relative contributions to THIA degradation were also calculated by combining a kinetic model and a series of probe compound-based experiments. The results indicated that SO4˙ and HO˙ were generated at lower steady-state concentrations than that of 1O2, but they dominated THIA abatement during the MCO-0.2/PMS process. This study presented new insights into the construction of efficient PMS activator and a mechanistic understanding for PMS-mediated reaction.


1. Introduction

Refractory organic pollutants (e.g., pesticides, antibiotics and industrial chemicals) in wastewater constitute a serious threat to the ecosystem and human health.1–3 For eliminating these organic pollutants from wastewater, peroxymonosulfate-based advanced oxidation processes (PMS-AOPs) have been extensively studied and recognized as a promising approach4–6 during the past few years. In PMS-AOPs, reactive species (RS) originated from PMS activation is crucial to accelerate these organic pollutants degradation.7–9 Therefore, many catalysts, including some transition metals (Co, Fe, Cu and Mn),10–13 carbonaceous materials and their composites,14,15 have been developed for efficient PMS activation during the past few years. Of all these catalysts, Co-based heterogeneous catalysts (e.g., metal ions, oxides, hydroxides and Co-containing single atomic catalysts)16–20 have drawn extensive attention because they are amongst the most active PMS activator and can be recycled to minimize environmental impact.21,22

Among various Co-based catalysts, Co3O4 stands out as a low-cost catalyst and was widely used in different oxidation reaction systems.23,24 Normal Co3O4 catalysts with two different Co sites in a spinel structure, where Co2+ is bonded to four neighboring oxygen atoms at tetrahedral sites and Co3+ is bonded to six neighboring oxygen atoms at the octahedral sites, have been demonstrated to be efficient for PMS activation.25,26 Previous studies27,28 have revealed that both tetrahedral Co2+ and octahedral Co3+ were the efficient active sites of PMS activation in PMS-AOPs over the Co3O4 catalysts. Specifically, two reactions (the reductive reaction of PMS and the oxidative reaction of PMS) should take place simultaneously to generate radicals continuously through a Co2+/Co3+ redox cycle (eqn (1) and (2)).27,28

The oxidative reaction of PMS:

 
Co2+ + HSO5 → Co3+ + SO4˙ + OH (1)

The reductive reaction of PMS:

 
Co3+ + HSO5 → Co2+ + SO5˙ + H+ (2)

Undoubtedly, the different density of Co2+/Co3+ on the surface of Co3O4 catalysts might be resulted in the different capacity for the circulation of Co2+/Co3+ and further vigorously affect the activity of Co3O4 catalysts.29 Moreover, the different density of Co2+/Co3+ on the surface could also change atomic arrangements and electronic structures of Co3O4 to favour Co2+/Co3+ redox cycle.30,31 Thus, tuning the atomic ratios of Co2+/Co3+ exposed on the surface of Co3O4-based catalytic materials could be a reasonable way to construct a synergistic effect of Co2+ and Co3+ in the Co2+/Co3+ redox cycle for an improved PMS activation.

Metal doping was a simple and efficient approach to engineer surface tetrahedral Co2+ and octahedral Co3+ of Co3O4 for manipulating the ratio of Co2+/Co3+ of Co3O4 by the substitution of tetrahedral Co2+ and octahedral Co3+ with the corresponding valence states of inactive metal.32,33 For example, Dong's group17 reported the synthesis of Al doped Co3O4 catalysts by incorporating inert Al3+ ion into the lattice of Co3O4. They found that Al incorporation could partly replace octahedral Co3+ and modify the ratio of Co2+/Co3+ in Co3O4 for an improved performance of PMS activation. While previous studies34 have demonstrated octahedral Co3+ possessed a high standard reduction potential (E0(Co3+/Co2+) = 1.92 V) and was notably more active than tetrahedral Co2+ for OER. The significant roles for pollutant abatement of octahedral Co3+ were also confirmed during the PMS-AOPs.35 Therefore, the engineering of tetrahedral Co2+ by the doping of divalent metal ion (Zn2+ and Mg2+) might be an effective way to regulate the atomic ratio of Co2+/Co3+ of Co3O4 for an improved PMS activation.10,20 While Mg2+ has a similar ionic radius to that of Co2+ (0.72 Å vs. 0.74 Å).28 The precise substitution of tetrahedral Co2+ with inert Mg2 could be easily realized for the regulation of Co2+/Co3+ ratio in the spinel structure of Co3O4, but might not change Co3O4 host lattice,28,29 which facilitated us to judge the important role of Co2+/Co3+ ratio in PMS activation, however it has not been reported yet.

Hence, a magnesium doping strategy was developed to manipulate Co2+/Co3+ ratio in the cobalt spinel for an enhanced PMS activation, and obtained a series of Mg doped Co3O4 (MCO) catalysts. Optimal MCO catalysts (MCO-0.2) with suitable Co2+/Co3+ atomic ratios possessed superior activities for THIA degradation during PMS-AOPs, and the rate constant (0.2835 min−1) was 2.97 folds faster than that of pure Co3O4 (0.0955 min−1). To further elucidate the degradation mechanisms, the formation of reactive species in the MCO-0.2/PMS process was verified by scavenger tests and EPR characterizations. Moreover, the steady-state concentrations of these RS were quantified through a kinetic model and several probe experiments. Using the newly measured kinetic data, the relative contributions of these RS to THIA degradation were thus determined. This study might present valuable design guide of cobalt-based catalysts with regulated Co2+/Co3+ ratio for environmental applications.

2. Experimental procedures

2.1. Chemicals and materials

Thiacloprid (THIA), sodium hydroxide (NaOH), atrazine (ATZ), cobalt nitrate hexahydrate (Co(NO3)2·6H2O), sodium bicarbonate (NaHCO3), methanol (MeOH), magnesium nitrate dihydrate (Mg(NO3)2·2H2O), chloramphenicol (CAP) and tert-butyl alcohol (TBA) were obtained from Sinopharm Chemical Reagent Co. Ltd. Potassium peroxymonosulfate (PMS), sodium chloride (NaCl), sulfuric acid (H2SO4), sodium nitrate (NaNO3), disodium hydrogen phosphate (Na2HPO4), sodium thiosulfate (Na2S2O3), furfuryl alcohol (FFA), methyl phenyl sulfone (PMSO2), metronidazole (MTZ), methyl phenyl sulfoxide (PMSO), commercial nano-Co3O4 and MgCo2O4 were purchased from Aladdin Company, China. All chemical reagents were used without further purification. Deionized water (DI water) was used throughout the whole experiments.

2.2. Catalyst synthesis

The synthesis of MCO catalysts involved the pre-synthesis of Co-based precursors and subsequently calcining them under air atmosphere. The obtained catalysts were labelled MCO-X, where X indicated the nominal molar ratio of Mg/(Co + Mg) in the catalysts. Taking MCO-0.2 as an example, 1.5 mmol Mg(NO3)2·2H2O and 6 mmol Co(NO3)2·6H2O were mixed in 100 mL DI water. Subsequently, the solution pH was adjusted to 11 by dropwise addition of 0.1 M NaOH solution under continuous stirring. The above solution was centrifugated at 10[thin space (1/6-em)]000 rpm for 10 min, and the obtained precipitate was resuspended in a small volume of DI water, repeating the procedure until the suspension pH reached 8. After rinsing with ethanol several times, the precipitate was dried at 80 °C for 24 h, and then thermally treated at 350 °C for 3 h in air atmosphere to obtain the final product. Similarly, Co3O4 (Co(NO3)2·6H2O: 7.5 mmol), MCO-0.1 (Mg(NO3)2·2H2O: 0.75 mmol, Co(NO3)2·6H2O: 6.75 mmol), MCO-0.3 (Mg(NO3)2·2H2O: 2.25 mmol, Co(NO3)2·6H2O: 5.25 mmol) were prepared as above using different doses of Mg(NO3)2·2H2O and Co(NO3)2·6H2O. For comparison, ZnCo2O4 was prepared through the above procedure with appropriate doses of Zn(NO3)2·2H2O and Co(NO3)2·6H2O. β-Co(OH)2 were synthesized with same method as that of Co3O4 without thermal treatment.

2.3. Catalytic performance

The activities of Co3O4 and MCO catalysts were checked in 100 mL THIA solution with mechanically agitating. Specifically, the catalysts were first spiked into THIA solution and further stirred until adsorption/desorption equilibrium, followed by the initiation of the reaction with the introduction of 0.2 mM PMS. Periodically, 2 mL samples were filtered through a 0.22 μm filter and immediately quenched with Na2S2O3 solution, followed by component analysis. The pH value of the reaction system was under control utilizing 0.1 M NaOH or H2SO4. The presence of reactive species (RS) in THIA degradation was evaluated by scavenger tests, where TBA, MeOH, DMSO and FFA were used as RS quenchers and spiked into the system before PMS addition. In addition, the depletion experiments of several probe compounds (including CAP, ATZ and MTZ) were performed as the same conditions as that of THIA abatement to measure RS exposures during the MCO-0.2/PMS process. Moreover, the reusability of MCO-0.2 catalysts was investigated through recycle tests. After each run, MCO-0.2 water and then reused under the same experimental conditions (details provided in Text S1).

2.4. Analytical methods

The crystalline structures of the synthetized catalysts were detected by a Bruker D8 Advance X-ray diffractometer (XRD) using Cu Ka radiation (λ = 1.5418 Å) with 2-theta range of 10–80° and scanning rate of 10° s−1. The micro-morphologies of pure Co3O4 and MCO-0.2 catalysts were measured by scanning electron microscope (SEM, Zeiss Gemini 300) equipped with energy dispersive spectrometer (EDS, Oxford X-MAX). The surface elements and chemical valences of pure Co3O4, MCO-0.1 and MCO-0.2 catalysts were analysed by X-ray photoelectron spectroscopy (XPS, Thermo Scientific ESCALAB 250, USA) with Al Ka radiation, and the binding energies were calibrated with the residual C 1s peak (284.8 eV). Zeta potentials of MCO-0.2 catalysts were measured by Malvern Zeta sizer Nano ZS90 through the method described in Text S2. Electron paramagnetic resonance (EPR) signals of active species were monitored on a Bruker A300 EPR spectrometer using TEMP or DMPO as capture agents. In addition, the concentration of THIA and probe compounds were determined by high-performance liquid chromatography (HPLC, Agilent 1260), and the details of the measurements were illustrated in Text S3. Cobalt dissolution amount was measured by inductively coupled plasma-mass spectrometry (ICP-MS, Optima 5300 DV, USA). The residual concentrations of PMS were analyzed through the method described in Text S4.

3. Results and discussion

3.1. Characterization of MCO catalysts

MCO catalysts were prepared through a co-precipitation method followed by the calcination in air (Fig. 1a), in which the precipitate was recovered by the centrifugation at 10[thin space (1/6-em)]000 rpm and then resuspended in DI water, repeating the sequence several times to obtain the mono-dispersed catalysts. By adjusting the amounts of Mg ions in the precursor, the Mg doping degree could be controlled and a series of MCO catalysts were obtained. SEM images of Co3O4 and MCO-0.2 catalysts were presented in Fig. 1b and c, suggesting Co3O4 exhibited the granular structure with an average diameter around 40 nm. Notably, MCO-0.2 exhibited a similar size and shape to that of Co3O4, which indicated that Mg doping did not destroy the sample morphology. Moreover, EDS analysis (Fig. 1d and e) demonstrated the presence of Co, O and Mg elements in MCO-0.2, and Mg/Co atomic ratio was about 1[thin space (1/6-em)]:[thin space (1/6-em)]23, suggesting partial Co could be substituted by Mg atoms. These results suggested that MCO catalysts with similar morphology and different Mg contents were successfully synthesised, which facilitate us to judge the important role of Mg doping in PMS activation.
image file: d5ra04080a-f1.tif
Fig. 1 (a) Schematic illustration of the preparation of MCO-X catalysts, SEM images of (b) Co3O4 and (c) MCO-0.2; EDS spectra for (d) Co3O4 and (e) MCO-0.2, respectively.

XRD measurements were employed to verify the crystal characteristics of Co3O4 and MCO-X catalysts. As can be seen in Fig. 2a, XRD pattern of Co3O4 exhibited the diffraction peaks at 2θ values of 18.9°, 31.3°, 36.9°, 38.4°, 44.8°, 55.7°, 59.2°, and 65.2° were attributed to (111), (220), (311), (222), (400), (422), (511) and (440) lattice planes of the spinel-type Co3O4 (space group Fd[3 with combining macron]m, JCPDS card no. 42-1467).9,13 For MCO catalysts, only peaks of Co3O4 can be observed, indicating that Mg atoms should incorporate into the Co3O4 crystal lattice and the introduction of Mg did not change the Co3O4 host lattice, due to similar ionic radius of Mg2+ to that of Co2+ (0.72 Å of Mg2+ vs. 0.74 Å of Co2+).17,28 After all, Mg2+ has a little smaller ionic radius to that of Co2+, Mg doping might lead to a slight change of peak positions and intensities for these catalysts.28,29 Specially, a slight shift of the peak representing the (311) plane to lower angle can be observed by expanding the abscissa (Fig. 2b) of the XRD patterns for the MCO-X catalysts in comparison with that of the pure Co3O4. Since the (311) lattice plane of MgCo2O4 is lower to that of Co3O4 by about 2θ = 0.047°,13 the shift should be caused by the substitution of tetrahedral Co2+ with Mg2+. As Mg/Co atomic ratio increased, a further shift to the lower angle appeared on the series of MCO-X catalysts, indicating more tetrahedral Co2+ could be substituted by Mg2+. This structure change might lead to the improved chemical properties and the higher activity.13

XPS analysis was performed to probe surface elements and chemical states of pure Co3O4, MCO-0.1 and MCO-0.2 catalysts. The survey spectra (Fig. 2c) clearly confirmed the presence of Co, O and Mg elements in MCO-0.1 and MCO-0.2 catalysts, and the corresponding Mg 2p spectra (Fig. 2d) exhibited a characteristic peaks of Mg 2p, testifying the Mg atoms might be incorporated into the crystal structure of Co3O4, which agreed well with XRD measurements. In the Co 2p spectra (Fig. 2e), two main peaks at 779.4 and 794.5 eV should be attributed to Co 2p3/2 and Co 2p1/2, respectively, along with their corresponding satellite peaks (denoted as “sat.”) at 788.1 and 803.1 eV. The Co 2p3/2 spectra were fitted into Co2+ and Co3+ constituents at 779.7 and 781 eV, while Co 2p1/2 were also separated into the same components at 725.8 and 727.45 eV respectively.36,37 The ratio of their fitted peak area revealed that Co2+/Co3+ atomic ratio gradually decreased with the increment of Mg content, verifying Mg doping could be a potential strategy for manipulating surface metal state of Co3O4 toward improved activities. Meanwhile, the O 1s spectra (Fig. 2f) displayed three peaks, which ascribed to lattice oxygen (OL, ∼529.5 eV), surface hydroxyl groups (OOH, ∼530.3 eV), and oxygen vacancies (Ov, ∼531.6 eV).38 Notably, the peak area ratio of deficient oxygen gradually increased with increased Mg content, confirming Mg doping might create new oxygen vacancies.39 The finding revealed that due to the lower Co2+/Co3+ atomic ratio induced by Mg doping, the neighbouring oxygen atom might be more easily oxidized by octahedral Co3+ and squeezed out of the crystalline structure, thus forming MCO-0.2 with enriched oxygen vacancies at the surface.40


image file: d5ra04080a-f2.tif
Fig. 2 (a) XRD patterns and (b) the diffraction peak at 36–37.5° expanded in the abscissa; (c) XPS survey spectra, high-resolution XPS spectra of (d) Mg 2s, (e) Co 2p, (f) O 1s of Co3O4, MCO-0.1 and MCO-0.2 catalysts, respectively.

3.2. Catalytic activity of MCO catalysts

The catalytic activities of MCO catalysts were evaluated in thiacloprid (THIA) degradation via PMS activation. Fig. 3a showed negligible THIA removal was observed when each catalyst or PMS was used separately. Optimization experiments for the catalysts were also determined. Fig. 3b showed a volcano-like relationship between the catalytic performance and Mg doping amount. MCO-0.2 exhibited the best catalytic performance for THIA degradation and the rate constants was 0.28347 min−1, which was about 2.97, 2.02 and 1.35 times higher than that of Co3O4 (0.09555 min−1), MCO-0.1 (0.14045 min−1) and MCO-0.3 (0.20932 min−1), respectively. On the other hand, PMS decomposition (Fig. 3d and e) on the surface of different catalyst was in line with THIA degradation, which could be attributed to the fact that the catalyst accelerated the PMS activation to produce more ROS, thus resulting in significant THIA degradation. The depletion rates constants of PMS in MCO-0.3, MCO-0.2, MCO-0.1 and Co3O4 systems were 0.02597, 0.03648, 0.01145 and 0.01447 min−1, respectively. The above results suggested the important function of Mg doping in tuning the Co2+/Co3+ ratio for enhanced THIA degradation. MCO-0.2 might have the rational Co2+/Co3+ ratio, which facilitate the Co2+/Co3+ recycle, thus promoting PMS activation toward a significantly improved THIA degradation. However, the excessive substitution of Co2+ with Mg in MgCO-0.3 might deteriorating the Co2+/Co3+ recycle, ultimately resulting in the decrease of degradation efficiency.17
image file: d5ra04080a-f3.tif
Fig. 3 (a) THIA removal and (d) PMS depletion over Co3O4 and MCO-X catalysts, (b) and (e) the corresponding first-order rate constants; (c) THIA removal curves over MCO-0.2 and other similar catalysts, (f) the corresponding Arrhenius plots. Experiment conditions: pH = 7, [THIA] = 20 mg L−1, [catalyst] = 100 mg L−1, [PMS] = 0.4 mM.

To further verify the excellent activation performance of MCO-0.2 catalyst toward PMS, Similar catalysts such as β-Co(OH)2 and ZnCo2O4 were used to activate PMS for THIA degradation. As reported,35,41 the majority of cobalt was Co2+ ions at the tetrahedral sites of β-Co(OH)2, while most of cobalt was Co3+ ions at the octahedral sites of ZnCo2O4. Under the same conditions, the activation performance of these two catalysts (Fig. 3c and S1c) was considerably lower than that of MCO-0.2, with THIA removals of 78.7.2% and 48.2%, respectively, which suggested that the rational Co2+/Co3+ ratio in MCO-0.2 catalysts could favour Co2+/Co3+ recycle and boost the PMS activation toward a rapid THIA degradation. Furthermore, in comparison with commercial Co3O4 (denoted as C-Co3O4) and MgCo2O4 under the identical conditions (Fig. 3c and S1c), it was observed that the MCO-0.2 exhibited the higher THIA removal rates, which also highly outperformed many other reported cobalt-based catalysts (Table S2), indicating that MCO-0.2 had the extremely superior ability of PMS activation and great application potential. More interestingly, the activation energies (Ea) derived from a series of kinetic experiments also exhibited the similar trend (Fig. 3f and S2). The activation energy (Ea) of the MCO-0.2/PMS system was calculated to 41.6 kJ mol−1 through Arrhenius formula, which was considerably lower than that of Co(OH)2/PMS (46.7 kJ mol−1) and ZnCo2O4/PMS (51.2 kJ mol−1) systems, indicating that the rational Co2+/Co3+ ratio in MCO-0.2 catalyst could notably promote THIA degradation by lowering Ea. Thus, we selected MCO-0.2, which displayed the superior PMS activation capability, as the catalyst for further experiments.

3.3. Versatile applicability and reusability of MCO-0.2

To investigate the universal applicability of MCO-0.2, THIA degradation in the MCO-0.2/PMS system were examined under different reaction conditions (e.g., initial pH, PMS dosage, THIA concentration and co-existing anions). Fig. 4a showed the activity of MCO-0.2 was susceptible to the changes of pH value. The neutral pH condition achieved a higher THIA removal efficiency, while either acidic conditions or alkaline condition achieved a relatively low THIA degradation. Specifically, THIA removal efficiency was increased from 67.9% to 94.4% in 15 min by increasing the initial pH from 3.0 to 7.0. While THIA removal efficiency significantly declined from 94.4% to 41.9% with the further increase of pH from 7.0 to 11.0. As reported, pKa1 and pKa2 of PMS were about 0 and 9.4, respectively, which indicated that the main form of PMS was HSO5 under the acidic conditions and neutral conditions, while the main form was SO52− under alkaline conditions.42,43 Zero potential of MCO-0.2 was 6.91 (Fig. S3), demonstrating that the catalyst surface was mainly positively charged under negatively charged under alkaline conditions. As a result, pH has significant impact on the adsorption and existence form of PMS, thus notably affecting PMS activation. In the acidic solution, HSO5 and RS could be severely scavenged by excessive H+, therefore hampering PMS activation and THIA degradation. Under alkaline conditions, the main form of PMS was SO52− under alkaline conditions, which could be more difficultly activated to produce RS than that one (HSO5) under acidic condition and neutral condition. Furthermore, the electrostatic repulsion between the negatively charged catalyst and SO52− also greatly restrained the adsorption of PMS on the catalyst surface, thus future blocked PMS activation.
image file: d5ra04080a-f4.tif
Fig. 4 Effects of (a) initial pH, (b) PMS concentration, (c) coexisting ions and (d) and (e) THIA concentration on THIA degradation in the MCO-0.2/PMS system; (f) THIA degradation in the recycle tests of MCO-0.2 catalysts. Reaction conditions: [THIA] = 20 mg L−1, [PMS] = 0.4 mM, [catalyst] = 100 mg L−1, T = 298 K, pH = 7, [Cl]0 = [NO3]0 = [HCO3]0 = [SO42−]0 = 100 mM.

As for PMS dosage (Fig. 4b), the degradation efficiency of THIA was sharply enhanced as PMS dosage increased from 0.2 to 0.8 mM. However, the enhancement was negligible with PMS dosage further increasing. The optimal dosage of PMS was 0.8 mM with respect to practical cost. Apparently, more PMS could generate more ROS in the reaction system and accelerate THIA degradation. However, excessive PMS might quench the active radicals in the reaction system and deteriorate THIA degradation. Similar results were also reported by Li and Gong's groups.44,45 Fig. 4d showed THIA removal efficiency decreased from 94.4% to 40.3% as the concentration of THIA increased from 20 to 120 mg L−1, and the rate constants (Fig. S4b) decreased from 0.2835 to 0.04475 min−1, which could be attributed to insufficient RS produced from constant PMS to pollutants. Notable, although the reaction rates were different with the higher initial concentration of THIA (80 mg L−1 vs. 120 mg L−1), the final removal amount (C0C) was almost the same (Fig. 4e), suggesting that the removal of one THIA might consumed approximately two PMS molecules for these concentrations.

Some anions (including Cl, HCO32−, NO3, and SO42−) might be naturally present in actual water environment and affect THIA degradation.46,47 Therefore, the impact of these co-existing anions on THIA degradation was investigated in the MCO-0.2/PMS process. As shown in Fig. 4c, NO3 and SO42− possessed slight negative impact on THIA degradation with degradation efficiencies from 94.4% to 89.6% and 89.1% in 15 min, respectively, manifesting NO3 and SO42− might react with reactive species (SO4˙ and/or ˙OH) to generate the radicals with weaker redox, thus inhibit THIA degradation.46 While THIA degradation efficiencies increased from 94.4% to 95.8% after Cl was spiked the system, implying the reaction of Cl and reactive species might convert into powerful radicals (such as Cl2˙ and ClO˙),48 and future boost THIA degradation. Additionally, HCO3 caused a significant restraint (https://fanyi.so.com/) on THIA removal with degradation efficiencies decline to 17.8% in 15 min, which might be attributed to the fact that HCO3 could easily scavenge reactive species (SO4˙ and/or ˙OH) SO4˙ and/or ˙OH to generate weak oxidant (CO3˙), which attacks THIA more slowly.49 THIA degradation in various water conditions (such as deionized water, tap water and lake water) was also investigated. As shown in Fig. S9, similar THIA removal rates were witnessed in deionized water (DI water) and tap water. In lake water, THIA removal rate exhibited an apparent inhibition but still exceeded 85%, suggesting excellent practical application prospects.

Finally, the reusability of MCO-0.2 catalysts was evaluated through successive THIA degradation tests. Fig. 4f showed that the degradation rates of THIA declined slightly and were still above 85% after five cycles. THIA degradation rate for the five recycles was 94.4%, 92.8%, 91.2, 88.8% and 87.3% in 15 min, and the rate constant (k) was estimated to be 0.2835, 0.2668, 0.2427, 0.2121 and 0.19297 min−1 (Fig. S5b), demonstrating superior reusability of MCO-0.2. Simultaneously, the leaching of cobalt ion from MCO-0.2 catalysts was tracked during the successive processes. The dissolved cobalt ions for the five recycles was 0.082, 0.065, 0.062, 0.057 and 0.056 mg L−1 (Fig. S5c), which were all far below the permissible discharge quantity for surface water in China.3 In addition, THIA degradation was carried out in dissolved Co ions (Co2+ 0.08 mg L−1)/PMS system (Fig. S6) and THIA degradation efficiency was only 7.17%, indicating the leached cobalt ions contributed little to THIA removal. The above results prove MCO-0.2 catalysts have the excellent and stable catalytic activity just with a slight decrease after five cycles. The negligible decrease of the catalyst activity might be ascribed to the accumulation of products or intermediates on the catalyst during the THIA degradation process and cobalt leaching from the catalyst in the four recycling processes.

3.4 Mechanistic insights into PMS activation by MCO-0.2

3.4.1 Identification of reactive species. While it is generally agreed that various RS (including SO4˙, ˙OH, 1O2) were involved in the PMS-AOPs process, but their relative importance for pollutant abatement still remained significant controversies.10 Originally, sulfate radical (SO4˙) and its secondary radical (hydroxy radical, ˙HO) were proposed as the primary reactive intermediates for pollutant abatement during the persulfate-based process.50–52 Recently, a series of increasing finding demonstrated that 1O2 was the dominant reactive intermediates and played an even more important role than SO4˙ and ˙OH for pollutant abatement, especially during PMS-AOPs activated with carbon-based materials and carbon-metal composites.53–55 However, latest studies56,57 showed high-valent cobalt-oxo (Co(IV)[double bond, length as m-dash]O) species were successfully detected in the PMS-AOPs by the fact that Co(IV)[double bond, length as m-dash]O could easily oxidize DMSO to the corresponding sulfone product PMSO2, and contributed greatly to pollutant abatement. Thus, it is crucial to delve deeper into the components of the MCO-0.2/PMS system, which dominated the reaction process.

First, quenching experiments were firstly conducted to prove the presence of ROS during the MCO-0.2/PMS process. MeOH was used as the scavengers of sulfate radical (SO4˙) and hydroxy radical (˙HO). While TBA, FFA and DMSO were applied to scavenge ˙OH, 1O2 and Co(IV) [double bond, length as m-dash]O, respectively. As reported, MeOH can synchronously scavenge ˙OH or SO4˙ (k(MeOH, ˙OH) = 9.7 × 108 M−1 s−1; k(MeOH, SO4˙) = 3.2 × 106 M−1 s−1), but TBA can merely efficiently capture ˙OH (k(TBA, ˙OH) = 6 × 108 M−1 s−1; and k(TBA, SO4˙) = 8.4 × 105 M−1 s−1).58 Fig. 5a and b showed TBA and MeOH could apparently inhibited THIA degradation and the degradation rate decreased from 94.3% to 81.21% and 48.21% in 15 min, respectively, confirming the generation of SO4˙ and ˙HO in THIA degradation. While FFA could almost completely inhibited THIA degradation, and the corresponding rate constant (k) decreased to 0.0122, accounting for a 95.69% decrease, which confirmed the generation of 1O2 radicals. Further, the introduction of DMSO displayed a remarkably promoted inhibition for THIA degradation, and reduced the reaction rate by 63.76%. However, the formation of PMSO2 could not catch in the process. Since the reaction of PMSO with ˙OH or SO4˙ could also result in a vigorous inhibitory for THIA abatement without the formation of PMSO2,59 it is reasonable to speculate that Co(IV)[double bond, length as m-dash]O was almost absent in the reaction system.


image file: d5ra04080a-f5.tif
Fig. 5 (a) Removal efficiencies of THIA on MCO-0.2 catalysts in the presence of 200 mM quenchers, and (b) the corresponding first-order rate constants; EPR spectra of (c) DMPO and (d) TEMP as the trapping agent in the MCO-0.2/PMS system; (e) abatement of probe compounds on MCO-0.2 catalysts, and (f) the corresponding kinetic curves. Reaction conditions: [catalysts] = 100 mg L−1, [PMS] = 0.4 mM, [THIA] = 20 mg L−1, [CAP] = [ATZ] = [MTZ] = 0.08 mM, pH = 7, T = 298 K.

Furthermore, EPR tests were performed to verify the formation of the above-mentioned ROS in MCO-0.2/PMS system. Fig. 5c and d showed no distinct characteristic peaks were detected when only PMS was introduced in the THIA solution, indicating that the generation of ROS from PMS self-decomposition was negligible. However, when PMS and MCO-0.2 were added simultaneously in the system, characteristic peaks of different ROS could be clearly observed. To be specific, Fig. 5c showed the four well-defined peaks with intensity ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]2[thin space (1/6-em)]:[thin space (1/6-em)]2[thin space (1/6-em)]:[thin space (1/6-em)]1 were ascribed to the characteristic signals of DMPO-˙OH adducts, while the rest peaks were accordant with typical character signals of DMPO-SO4˙ adducts, signifying the formation of ˙OH and SO4˙ during the MCO-0.2/PMS process. In Fig. 5d, the three-line signals with an intensity ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1 were observed, ascribing to the characteristic signals of 1O2, confirmed the presence of 1O2 in the THIA degradation process. Therefore, EPR measurement further identified the involvement of ˙OH, SO4˙ and 1O2 in THIA degradation, agreeing well with the results of quenching tests.

3.4.2. Relative contributions of various reactive species. The above results proved that SO4˙, HO˙ and 1O2 were the reactive oxidants in the MCO-0.2/PMS system. Therefore, THIA degradation kinetic in the system could be simulated as follows:60
 
image file: d5ra04080a-t1.tif(3)

Integrating eqn (3) and yields eqn (4).

 
image file: d5ra04080a-t2.tif(4)
 
image file: d5ra04080a-t3.tif(5)
where [THIA]0 and [THIA] were the concentration of THIA at time 0 and t, respectively. image file: d5ra04080a-t4.tif, k˙OH,THIA and k1O2,THIA were the second-order rate constant for the reaction of THIA with SO4˙, HO˙ and 1O2, respectively (Table S2). kTHIA was the pseudo-first-order rate constant of THIA degradation in the MCO-0.2/PMS system. [SO4˙]ss, [˙OH]ss and [1O2ss] represented the steady-state concentrations of SO4˙, HO˙ and 1O2, respectively, and could be measured by several probe experiments.

In the probe experiments, probe compounds (including CAP, ATZ and MTZ) were added in the MCO-0.2/PMS system under identical conditions, and the abatement kinetic of these probes could be expressed as follows:

 
image file: d5ra04080a-t5.tif(6)
 
image file: d5ra04080a-t6.tif(7)
 
image file: d5ra04080a-t7.tif(8)
where kCAP, kATZ and kMTZ were the pseudo-first-order rate constants of the degradation of CAP, ATZ and MTZ probes in the MCO-0.2/PMS system, respectively, which were obtained from the kinetic plots in Fig. 5e and f. Moreover, image file: d5ra04080a-t8.tif, k˙OH,CAP, image file: d5ra04080a-t9.tif, image file: d5ra04080a-t10.tif, k˙OH,ATZ, k1O2,ATZ, image file: d5ra04080a-t11.tif, k˙OH,MTZ and k1O2,ATZ represented the second-order reaction rate constants of SO4˙, HO˙ and 1O2 with CAP, ATZ and MTZ, respectively (Table S1). By using the eqn (6)–(8), the steady-state concentrations of 1O2, ˙OH and SO4˙ in the MCO-0.2/PMS system were thus readily calculated to be 2.321 × 10−10, 1.079 × 10−11 and 4.408 × 10−11 M, respectively.

Further, the relative contributions of SO4˙, HO˙, 1O2 and other potential reactive species were defined as the ratios of THIA degradation rate induced by one specific reactive intermediate to the total degradation rate of THIA (kTHIA),16 and could be expressed as follows:

 
image file: d5ra04080a-t12.tif(9)
 
image file: d5ra04080a-t13.tif(10)
 
image file: d5ra04080a-t14.tif(11)
 
image file: d5ra04080a-t15.tif(12)

Using the newly measured kinetic data and the eqn (6)–(8), the relative contributions of SO4˙, HO˙, 1O2 and other potential reactive species for image file: d5ra04080a-t16.tif degradation in the MCO-0.2/PMS system were quantified to be 17.2%, 74.1%, and 3.5%, respectively. The measured results implied the contribution of ˙OH, SO4˙ and 1O2 to THIA abatement relied on both their reactivity and exposures. As reported,61 image file: d5ra04080a-t17.tif and k˙OH,THIA were several orders of magnitude larger than k1O2,THIA. Therefore, although the exposures of ˙OH/˙SO4 are about two or more times lower than that of 1O2, they dominated THIA abatement in the system. Further, since ˙OH exposures are 2.9 times higher than the exposures of SO4˙, it contributed predominantly to THIA abatement in the MCO-0.2/PMS system given the similar value of image file: d5ra04080a-t18.tif and k˙OH,THIA.

Combined with the above analysis, a possible activation mechanism of PMS on MCO-0.2 catalysts was proposed (Fig. S8). Initially, PMS was adsorbed onto the catalyst surface and activated by the active Co2+, accompanying with the generation of Co3+ and SO4˙; meanwhile Co3+ ions withdraw electrons from PMS to form Co2+ and SO5˙, driving a Co2+/Co3+ cycle. Subsequently, partial SO4˙ further react with H2O or OH of the solution to produce HO˙. SO4˙ and HO˙ would also react with PMS to produce SO5˙, which could be decomposed into 1O2. Consequently, upon a rapid Co2+/Co3+ cycle, these generated RS could efficiently degraded THIA molecules over MCO-0.2 catalysts, where HO˙ might contributed mainly to THIA degradation given the exposures of these RS and their reactivity with THIA.

4. Conclusions

Mg-doped Co3O4 with the regulated ratios of Co2+/Co3+ were prepared by incorporating Mg dopants into the lattice of Co3O4 and used as PMS activator for THIA degradation. Structural characterizations and experimental investigations confirmed that Mg doping did not change the Co3O4 host lattice and particle morphology, but could manipulate surface metal state of Co3O4 for an improved PMS activation. The optimized sample (MCO-0.2) with the suitable Co2+/Co3+ atomic ratios (1.13) exhibited efficient THIA degradation, and the rates constants (0.28347 min−1) highly outperformed that of pure Co3O4, β-Co(OH)2 and ZnCo2O4. A systematic analysis of the influences of reaction parameters indicated that the maximum degradation might be achieved under the conditions: catalyst dose 100 mg L−1, PMS concentration 0.8 mM, pH 7 and THIA concentration 20 mg L−1. Quenching experiment and ERP tests confirmed that 1O2, SO4˙ and ˙OH were all involved in MCO-0.2/PMS System. According to a kinetic model and a series of probe compound-based experiments, the steady-state concentrations of SO4˙, HO˙ and 1O2 were determined to be 1.079 × 10−11, 4.408 × 10−11 and 2.321 × 10−10 M in MCO-0.2/PMS system, and their corresponding contributions to THIA degradation were also quantified to 17.2%, 74.1%, and 3.5%. This work may inform a new approach for constructing efficient Fenton-like catalysis.

Conflicts of interest

There are no conflicts to declare.

Data availability

All data included in this study are available upon request by contact with the corresponding author.

Additional details on analytical methods of PMS and various organic pollutants, SI degradation results and the parameters for various organic pollutants, and other associated SI tables and figures. See DOI: https://doi.org/10.1039/d5ra04080a.

Acknowledgements

This work was supported by Hubei Key Laboratory of Animal Nutrition and Feed Science, Wuhan Polytechnic University (No. 201911), the building project of Hubei Province for Technology Innovation and Entrepreneurship Service Capacity (No. 2018BEC466).

Notes and references

  1. A. Alsbaiee, B. J. Smith, L. L. Xiao, Y. H. Ling, D. E. Hebling and W. R. Dichtel, Rapid removal of organic micropollutants from water by a porous beta-cyclodextrin polymer, Nature, 2016, 529, 190–194 CrossRef CAS PubMed.
  2. S. D. Richardson and T. A. Ternes, Water analysis: emerging contaminants and current issues, Anal. Chem., 2014, 86, 2813–2848 Search PubMed.
  3. Y. L. Luo, W. S. Guo, H. H. Ngo, L. D. Nghiem, F. I. Hai, J. Zhang, S. Liang and X. C. Wang, A review on the occurrence of micropollutants in the aquatic environment and their fate and removal during wastewater treatment, Sci. Total Environ., 2014, 473–474, 619 CrossRef CAS PubMed.
  4. M. S. Han, W. K. Zhu, M. S. A. Hossain, J. You and J. Kim, Recent progress of functional metal-organic framework materials for water treatment using sulfate radicals, Environ. Res., 2022, 211, 112956 CrossRef CAS PubMed.
  5. X. Wang, Z. Xiong, H. Shi, Z. Wu, B. Huang, H. Zhang, P. Zhou, Z. Pan, W. Liu and B. Lai, Switching the reaction mechanisms and pollutant degradation routes through active center size-dependent Fenton-like catalysis, Appl. Catal., B, 2023, 329, 122569 CrossRef CAS.
  6. S. Mao, G. Y. Yao, P. Liu, C. Liu, Y. Wu, Z. T. Ding, C. Ding, M. Z. Xia and F. Y. Wang, Construction of an enhanced built-in electric field in S-doped g-C3N4/NiCo2O4 for boosting peroxymonosulfate activation, Chem. Eng. J., 2023, 470, 144250 CrossRef CAS.
  7. Y. Bao, C. Lian, K. Huang, H. R. Yu, W. Y. liu, J. L. Zhang and M. Y. Xing, Generating High-valent Iron-oxo≡FeIV[double bond, length as m-dash]O Complexes in Neutral Microenvironments through Peroxymonosulfate Activation by Zn-Fe Layered Double Hydroxides, Angew. Chem., Int. Ed., 2022, 61, e202209542 CrossRef CAS PubMed.
  8. C. L. Song, Q. Zhan, F. Liu, C. Wang, H. C. Li, X. Wang, X. F. Guo, Y. C. Cheng, W. Sun, L. Wang, J. S. Qian and B. C. Pan, Overturned loading of inert CeO2 to active Co3O4 for unusually improved catalytic activity in Fenton-like reactions, Angew. Chem., Int. Ed., 2022, 61, e202200406 CrossRef CAS PubMed.
  9. J. Wang and S. Wang, Activation of persulfate (PS) and peroxymonosulfate (PMS) and application for the degradation of emerging contaminants, Chem. Eng. J., 2018, 334, 1502–1517 CrossRef CAS.
  10. Y. K. Pan, J. Z. Cao, M. Y Xing and Y. Y. Zhang, Current mechanism of peroxymonosulfate activation by cobalt-based heterogeneous catalysts in degrading organic compounds, ACS ES&T Eng., 2024, 4, 19–46 Search PubMed.
  11. J. Xie, S. H. Wu, C. Y. Luo, J. C. Zou, Y. Lin, S. Y. He and C. P. Yang, Modulating electronic structure of active sites on iron-based nanoparticles enhances peroxymonosulfate activation, Appl. Catal., B, 2024, 354, 124138 CrossRef CAS.
  12. F. Li, Z. Lu, T. Li, P. Zhang and C. Hu, Origin of the Excellent Activity and Selectivity of a Single-Atom Copper Catalyst with Unsaturated Cu-N2 Sites via Peroxydisulfate Activation: Cu(III) as a Dominant Oxidizing Species, Environ. Sci. Technol., 2022, 56(12), 8765–8775 CrossRef CAS PubMed.
  13. F. Wang, M. Xiao, X. Ma, S. Wu, M. Ge and X. Yu, Insights into the transformations of Mn species for peroxymonosulfate activation by tuning the Mn3O4 shapes, Chem. Eng. J., 2021, 404, 127097 CrossRef CAS.
  14. J. J. Jiang, Z. Q. Zhao, J. Y. Gao, T. R. Li, M. Y. Li, D. D. Zhou and S. S. Dong, Nitrogen Vacancy-Modulated Peroxymonosulfate Nonradical Activation for Organic Contaminant Removal via High-Valent Cobalt-Oxo Species, Environ. Sci. Technol., 2022, 56(9), 5611–5619 CrossRef CAS PubMed.
  15. J. J. Pei, K. X. Fu, Y. K. Fu, X. L. Liu, S. L. Luo, K. Yin and J. M. Luo, Manipulating High-Valent Cobalt-Oxo Generation on Co/N Codoped Carbon Beads via PMS Activation for Micropollutants Degradation, ACS ES&T Eng., 2023, 3(11), 1997–2007 Search PubMed.
  16. Y. L. Cao, Z. Wang, S. X. He, L. X. Shi, K. H. Guo and J. Y. Fang, Reinvestigation on High-Valent Cobalt for the Degradation of Micropollutants in the Co(II)/Peroxymonosulfate System: Roles of Co(III), Environ. Sci. Technol., 2024, 58(7), 3564–3575 CAS.
  17. H. T. Li, Q. Gao, G. S. Wang, B. Han, K. S. Xia, J. P. Wu, C. G. Zhou and J. Dong, Postsynthetic incorporation of catalytically inert Al into Co3O4 for peroxymonosulfate activation and insight into the boosted catalytic performance, Chem. Eng. J., 2021, 426, 131292 CrossRef CAS.
  18. W. Li, S. Li, Y. Tang, X. Yang, W. Zhang, X. Zhang, H. Chai and Y. Huang, Highly efficient activation of peroxymonosulfate by cobalt sulfide hollow nanospheres for fast ciprofloxacin degradation, J. Hazard. Mater., 2020, 389, 121856 CrossRef CAS PubMed.
  19. H. Fui, S. M. Gao, X. R. Ma and Y. P. Huang, Facile fabrication of CoAl-LDH nanosheets for efficient rhodamine B degradation via peroxymonosulfate activation, RSC Adv., 2023, 13, 29695 RSC.
  20. S. H. Wu, Z. W. Yang, Z. Y. Zhou, X. Li, Y. Lin, J. J. Cheng and C. P. Yang, Catalytic activity and reaction mechanisms of single-atom metals anchored on nitrogen-doped carbons for peroxymonosulfate activation, J. Hazard. Mater., 2023, 459, 132133 CrossRef CAS PubMed.
  21. J. Di, R. Jamakanga, Q. Chen, J. Li, X. Gai, Y. Li, R. Yang and Q. Ma, Degradation of rhodamine B by activation of peroxymonosulfate using Co3O4-rice husk ash composites, Sci. Total Environ., 2021, 784, 147258 CrossRef CAS PubMed.
  22. S. Shao, L. Qian, X. Zhan, M. Wang, K. Lu, J. Peng, D. Miao and S. Gao, Transformation and toxicity evolution of amlodipine mediated by cobalt ferrite activated peroxymonosulfate: effect of oxidant concentration, Chem. Eng. J., 2020, 382, 123005 CrossRef CAS.
  23. M. S. Wu, B. Xu and C. Y. Ouyang, Manipulation of spin–flip in Co3O4: a first principles study, J. Mater. Sci., 2016, 51, 4691–4696 CrossRef CAS.
  24. S. D. Liu, D. X. Ni, H. F. Li, K. N. Hui, C. Y. Ouyang and S. C. Jun, Effect of cation substitution on the pseudocapacitive performance of spinel cobaltite MCo2O4 (M = Mn, Ni, Cu, and Co) dagger, J. Mater. Chem. A, 2018, 6, 10674–10685 RSC.
  25. Y. Wei, J. Miao, J. Ge, J. Lang, C. Yu, L. Zhang, P. J. J. Alvarez and M. Long, Ultrahigh peroxymonosulfate utilization efficiency over CuO nanosheets via heterogeneous Cu(III) formation and preferential electron transfer during degradation of phenols, Environ. Sci. Technol., 2022, 56, 8984–8992 CrossRef CAS PubMed.
  26. J. Li, M. Xu, G. Yao and B. Lai, Enhancement of the degradation of atrazine through CoFe2O4 activated peroxymonosulfate (PMS) process: kinetic, degradation intermediates, and toxicity evaluation, Chem. Eng. J., 2018, 348, 1012–1024 CrossRef CAS.
  27. E. T. Yun, H. Y. Yoo, H. Bae, H. I. Kim and J. Lee, Exploring the role of persulfate in the activation process: radical precursor versus electron acceptor, Environ. Sci. Technol., 2017, 51, 10090–10099 CrossRef CAS PubMed.
  28. P. Niu, C. H. Li, D. Q. Wang, C. X. Jia, J. Zhao, Z. M. Liu, X. L. Zhang and L. L. Geng, Electronic modulation of fiber-shaped-CoFe2O4 via Mg doping for improved PMS activation and sustainable degradation of organic pollutants, Appl. Surf. Sci., 2022, 605, 154732 CrossRef CAS.
  29. X. P. Han, G. W. He, Y. He, J. F. Zhang, X. R. Zheng, L. L. Li, C. Zhong, W. B. Hu, Y. D. Deng and T. Y. Ma, Engineering Catalytic Active Sites on Cobalt Oxide Surface for Enhanced Oxygen Electrocatalysis, Adv. Energy Mater., 2018, 8, 1702222 CrossRef.
  30. Q. F. Liu, Z. P. Chen, Z. Yan, Y. Wang, E. Wang, S. Wang, S. D. Wang and G. Q. Sun, Crystal-Plane-Dependent Activity of Spinel Co3O4 Towards Water Splitting and the Oxygen Reduction Reaction, ChemElectroChem, 2018, 5, 1080–1086 CrossRef CAS.
  31. C. P. Plaisance and R. A. van Santen, Structure Sensitivity of the Oxygen Evolution Reaction Catalyzed by Cobalt(II,III) Oxide, J. Am. Chem. Soc., 2015, 137, 14660–14672 CrossRef CAS PubMed.
  32. X. Li, Z. Wang, B. Zhang, A. I. Rykov, M. A. Ahmed and J. Wang, FexCo3−xO4 nanocages derived from nanoscale metal–organic frameworks for removal of bisphenol A by activation of peroxymonosulfate, Appl. Catal., B, 2016, 181, 788–799 CrossRef CAS.
  33. M. Sun, S. N. Tang, Z. X. Chen, L. F. Zhai, Y. H. Xia and S. B. Wang, Mn-Doping-Induced Co3O4 Structural Distortion and Facet Change for Enhanced Electro-Activation of O2 toward Pollutants Removal, ACS ES&T Eng., 2024, 4(8), 1970–1980 Search PubMed.
  34. Y. Xu, F. C. Zhang, T. Sheng, T. Ye, D. Yi, Y. J. Yang, S. J. Liu, X. Wang and J. N. Yao, Clarifying the controversial catalytic active sites of Co3O4 for the oxygen evolution reaction, J. Mater. Chem. A, 2019, 7, 23191–23198 RSC.
  35. P. Wu, L. Xia, Y. Liu, J. Wu, Q. Chen and S. Song, Simultaneous Sorption of Arsenate and Fluoride on Calcined Mg–Fe–La Hydrotalcite-Like Compound from Water, ACS Sustain. Chem. Eng., 2018, 6(12), 16287–16297 CrossRef CAS.
  36. A. F. Syeda and M. N. Khan, Spin State of Cobalt and Electrical Transport Mechanism in MgCo2O4 System, J. Supercond. Novel Magn., 2018, 31(11), 3545–3551 CrossRef CAS.
  37. W. Li, X. He, B. Li, B. Zhang, T. Liu, Y. Hu and J. Ma, Structural tuning of multishelled hollow microspheres for boosted peroxymonosulfate activation and selectivity: role of surface superoxide radical, Appl. Catal., B, 2022, 305, 1210 Search PubMed.
  38. M. Li, S. You, X. Duan and Y. Liu, Selective formation of reactive oxygen species in peroxymonosulfate activation by metal-organic framework-derived membranes: a defect engineering-dependent study, Appl. Catal., B, 2022, 312, 121419 CrossRef CAS.
  39. L. Zhuang, L. Ge, Y. Yang, M. Li, Y. Jia, X. Yao and Z. Zhu, Ultrathin iron-cobalt oxide nanosheets with abundant oxygen vacancies for the oxygen evolution reaction, Adv. Mater., 2017, 29(17), 1606793 CrossRef PubMed.
  40. W. Zhang, S. Zhang, C. Meng and Z. Zhang, Nanoconfined catalytic membranes assembled by cobalt-functionalized graphitic carbon nitride nanosheets for rapid degradation of pollutants, Appl. Catal., B, 2023, 322, 122098 CrossRef CAS.
  41. B. E. Channab, M. E. Ouardi, S. E. Marrane, O. A. Layachi, A. E. Idrissi, S. Farsad, D. Mazkad, A. BaQais, M. Lasri and H. A. Ahsaine, Alginate@ZnCO2O4 for efficient peroxymonosulfate activation towards effective rhodamine B degradation: optimization using response surface methodology, RSC Adv., 2032, 13, 20150–20163 RSC.
  42. C. Tan, X. Lu, X. Cui, X. Jian, Z. Hu, Y. Dong, X. Liu, J. Huang and L. Deng, Novel activation of peroxymonosulfate by an easily recyclable VC@Fe3O4 nanoparticles for enhanced degradation of sulfadiazine, Chem. Eng. J., 2019, 363, 318–328 CrossRef CAS.
  43. H. Li, Q. Gao, G. Wang, B. Han, K. Xia and C. Zhou, Architecturing CoTiO3 overlayer on nanosheets-assembled hierarchical TiO2 nanospheres as a highly active and robust catalyst for peroxymonosulfate activation and metronidazole degradation, Chem. Eng. J., 2020, 392, 123819 CrossRef CAS.
  44. Z. Li, D. Liu, Y. Zhao, S. Li, X. Wei, F. Meng, W. Huang and Z. Lei, Singlet oxygen dominated peroxymonosulfate activation by CuO-CeO2 for organic pollutants degradation: performance and mechanism, Chemosphere, 2019, 233, 549–558 CrossRef CAS PubMed.
  45. C. Gong, F. Chen, Q. Yang, K. Luo, F. Yao, S. Wang, X. Wang, J. Wu, X. Li, D. Wang and G. Zeng, Heterogeneous activation of peroxymonosulfate by Fe-Co layered doubled hydroxide for efficient catalytic degradation of rhodamine B, Chem. Eng. J., 2017, 321, 222–232 CrossRef CAS.
  46. M. Murugalakshmi, K. Govindan, M. Umadevi, C. B. Breslin and V. Muthuraj, Fabrication of a Sm2O3/In2S3 photocatalyst for boosting ciprofloxacin oxidation and the Cr(VI) reduction: process parameters and degradation mechanism, Environ. Sci.:Water Res. Technol., 2023, 9, 1385–1402 RSC.
  47. K. Govindan, D. G. Kim and S. O. Ko, Role of co-existing anions in non-radical and radical processes of carbocatalyzed persulfate activation for acetaminophen degradation, Environ. Sci.:Water Res. Technol., 2022, 8, 2940–2955 RSC.
  48. R. D. Su, N. Li, Z. Liu, X. Y. Song, W. Liu, B. Y. Gao, W. Z. Zhou, Q. Y. Yue and Q. Li, Revealing the Generation of High-Valent Cobalt Species and Chlorine Dioxide in the Co3O4-Activated Chlorite Process: Insight into the Proton Enhancement Effect, Environ. Sci. Technol., 2023, 57(5), 1882–1893 CrossRef CAS PubMed.
  49. P. Devi, U. Das and A. Dalai, In situ chemical oxidation: principle and applications of peroxide and persulfate treatments in wastewater systems, Sci. Total Environ., 2016, 571, 643–657 CrossRef CAS PubMed.
  50. J. Wei, D. Han, J. Bi and J. Gong, Fe-doped ilmenite CoTiO3 for antibiotic removal: electronic modulation and enhanced activation of peroxymonosulfate, Chem. Eng. J., 2021, 423, 130165 CrossRef CAS.
  51. L. Zhang, J. J. Qi, W. X. Chen, X. Y. Yang, Z. M. Fang, J. M. Li, X. Z. Li, S. Y. Lu and L. D. Wang, Constructing Hollow Multishelled Microreactors with a Nanoconfined Microenvironment for Ofloxacin Degradation through Peroxymonosulfate Activation: Evolution of High-Valence Cobalt-Oxo Species, Environ. Sci. Technol., 2023, 57(42), 16141–16151 CrossRef CAS PubMed.
  52. F. Li, Z. Lu, T. Li, P. Zhang and C. Hu, Origin of the Excellent Activity and Selectivity of a Single-Atom Copper Catalyst with Unsaturated Cu-N2 Sites via Peroxydisulfate Activation: Cu(III) as a Dominant Oxidizing Species, Environ. Sci. Technol., 2022, 56(12), 8765–8775 CrossRef CAS PubMed.
  53. A. D. Bokare and W. Choi, Singlet-Oxygen Generation in Alkaline Periodate Solution, Environ. Sci. Technol., 2015, 49(24), 14392–14400 CrossRef CAS PubMed.
  54. Y. Wang, Y. Lin, S. Y. He, S. H. Wu and C. P. Yang, Singlet oxygen: properties, generation, detection, and environmental applications, J. Hazard. Mater., 2024, 461, 132538 CrossRef CAS PubMed.
  55. L. Gao, Y. Guo, J. Huang, B. Wang, S. Deng, G. Yu and Y. Wang, Simulating micropollutant abatement during cobalt mediated peroxymonosulfate process by probe-based kinetic models, Chem. Eng. J., 2022, 441, 135970 CrossRef CAS.
  56. J. S. Song, N. N. Hou, X. C. Liu, M. Antonietti, Y. Wang and Y. Mu, Unsaturated single-atom CoN3 sites for improved fenton-like reaction towards high-valent metal species, Appl. Catal., B, 2023, 325, 122368 CrossRef CAS.
  57. Y. Zong, X. H. Guan, J. Xu, Y. Feng, Y. F. Mao, L. Q. Xu, H. Q. Chu and D. L. Wu, Unraveling the Overlooked Involvement of High-Valent Cobalt-Oxo Species Generated from the Cobalt(II)-Activated Peroxymonosulfate Process, Environ. Sci. Technol., 2020, 54(24), 16231–16239 CrossRef CAS PubMed.
  58. L. Gao, Y. Guo, J. Zhan, G. Yu and Y. Wang, Assessment of the validity of the quenching method for evaluating the role of reactive species in pollutant abatement during the persulfate-based process, Water Res., 2022, 221, 118730 CrossRef CAS PubMed.
  59. Z. Wang, J. Jiang, S. Pang, Y. Zhou, C. Guan, Y. Gao, J. Li, Y. Yang, W. Qiu and C. Jiang, Is Sulfate Radical Really Generated from Peroxydisulfate Activated by Iron(II) for Environmental Decontamination?, Environ. Sci. Technol., 2018, 52, 11276–11284 CrossRef CAS PubMed.
  60. H. J. Wang, L. W. Gao, Y. X. Xie, G. Yu and Y. J. Wang, Clarification of the role of singlet oxygen for pollutant abatement during persulfate-based advanced oxidation processes: Co3O4@CNTs activated peroxymonosulfate as an example, Water Res., 2023, 244, 120480 CrossRef CAS PubMed.
  61. G. Rózsa, L. Szabó, K. Schrantz, E. Takács and L. Wojnárovits, Mechanistic study on thiacloprid transformation: free radical reactions, J. Photochem. Photobiol., A, 2017, 343, 17–25 CrossRef.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.