Next Article in Journal
Probiotic Spores of Shouchella clausii SF174 and Displayed Bromelain Show Beneficial Additive Potential
Previous Article in Journal
GraphPhos: Predict Protein-Phosphorylation Sites Based on Graph Neural Networks
Previous Article in Special Issue
The Facile Production of p-Chloroaniline Facilitated by an Efficient and Chemoselective Metal-Free N/S Co-Doped Carbon Catalyst
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Heterogeneous Activation of NaClO by Nano-CoMn2O4 Spinel for Methylene Blue Decolorization

School of Urban Construction, Yangtze University, Jingzhou 434023, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2025, 26(3), 940; https://doi.org/10.3390/ijms26030940
Submission received: 4 December 2024 / Revised: 12 January 2025 / Accepted: 20 January 2025 / Published: 23 January 2025
(This article belongs to the Special Issue Advanced Catalytic Materials (Second Edition))

Abstract

:
In this study, the nano-spinel CoMn2O4 was synthesized by coprecipitation pyrolysis and employed to heterogeneously activate hypochlorite (NaClO) for the oxidative decolorization of methylene blue (MB). The crystal structure, elemental composition, surface morphology, and microstructure of the prepared CoMn2O4 nano-spinel were analyzed using a series of characterization techniques. The pyrolysis temperature was screened on the basis of MB decolorization efficiency and the leaching of metal ions during the reaction. The MB decolorization efficiency was compared using different catalysts and process. The impacts of CoMn2O4 dosage, effective chlorine dose, MB concentration, and initial pH on MB decolorization were explored. The catalytic mechanism of MB oxidation was elucidated through quenching experiments combined with radical identification. The degradation pathway of MB was preliminarily proposed based on the detection of the intermediates. The reusability of recycled CoMn2O4 was finally investigated. The results revealed that maximal MB oxidation efficiency and minimal leaching of Co and Mn ions were achieved at the calcination temperature of 600 °C. Complete oxidative decolorization of MB within 40 min was obtained at an initial MB concentration of 50 mg/L, a CoMn2O4 dosage of 1 g/L, an effective chlorine dose of 0.1%, and an initial pH of 4.3. Superoxide radical (O2•−) was found to be dominantly responsible for MB decolorization according to the results of radical scavenging experiments and electron paramagnetic resonance. The CoMn2O4 spinel can be recycled for five cycles with the MB removal in the range of 90.6~98.7%.

1. Introduction

Fast-paced industrial advancement has significantly eased human activities and daily life. Nevertheless, the emergence of novel wastewater types due to industrial development has presented formidable challenges to conventional water treatment methodologies. These traditional techniques frequently prove ineffective in managing emerging pollutants and fall short of meeting the increasingly stringent discharge standards [1]. In recent years, environmental scientists have carried out comprehensive research and discovered that among the array of pollution remediation technologies (biological, physical, chemical, etc.), advanced oxidation processes (AOPs) hold the greatest potential, demonstrating high efficiency in treating recalcitrant or difficult-to-mineralize organic pollutants [2,3,4,5]. In homogeneous/heterogeneous AOP reactions, highly oxidative species (such as HO, SO4•−, O2•−, 1O2) attack target pollutants, mineralizing them into small organic molecules and even carbon dioxide and water. AOPs provide advantages in terms of treatment speed, efficiency, and environmental friendliness [6,7]. To date, various AOP technologies have received widespread attention, including Fenton oxidation [8,9], photocatalytic oxidation [10], ozone oxidation [11], and persulfate oxidation [12]. However, homogeneous AOPs have limitations, such as poor catalyst availability and recovery difficulties. Some of these AOP technologies are costly (e.g., high energy input, expensive chemicals), which limits their large-scale application [7].
Considering the limitations of existing AOP technologies, an increasing number of researchers are now directing their attention towards the utilization of sodium hypochlorite (NaClO) for wastewater treatment. NaClO, as an oxidant, can also generate nascent oxygen species, which exhibit similar decomposition characteristics to those of H2O2 [13]. NaClO has several notable advantages, including a broader pH range of applicability compared to the Fenton process, greater cost-effectiveness in manufacture than alternative oxidants (with 10% NaClO priced at 82.8 $/t, in contrast to 30% H2O2 at 164.4 $/t and persulfate at 931.6 $/t as of 8 January 2025), and potent disinfection properties [14,15]. Furthermore, the waste produced after NaClO undergoes oxidation is non-toxic sodium chloride (NaCl), which has little effect on the environment at low concentrations [16]. Given these characteristics, the oxidative strength and decolorization efficacy using NaClO in treating wastewater are superior to conventional oxidants, demonstrating significant potential for practical applications. Stephanie et al. demonstrated the oxidation potential of hypochlorous acid in promoting substitution reactions [17]. Existing research indicated that chlorine addition and coupling reactions were the main mechanisms of oxidative degradation of organic pollutants using NaClO [18,19]. Therefore, direct use of NaClO to oxidize pollutants may generate organic chlorides, which could be more harmful than the original pollutants [20]. However, this issue can be addressed by combining NaClO with other processes [18].
Studies have shown that certain metal-based catalysts can effectively enhance the oxidation capacity of sodium hypochlorite (NaClO), thereby improving its oxidative performance [13]. Transition metals (such as manganese, cobalt, iron, nickel, and copper) are easily oxidized by NaClO to higher valence states due to the instability of their valence electrons, gaining strong oxidation capabilities. Thus, they can serve as effective catalysts to promote redox reactions [21]. Meanwhile, some metal-containing oxides are magnetic (such as iron, cobalt, nickel, manganese), which can aid their recovery after water treatment. The oxidative nature of NaClO is attributed to the O-Cl bond, which is easily broken at room temperature [22]. Under the action of transition metal catalysts, the broken O-Cl bond may lead to the production of different states of chlorine and reactive oxygen species, further promoting oxidation reactions (Equation (1)) [23].
2   C l O metals   2 C l + O 2
Based on this characteristic, the active species produced by transition-metal-catalyzed NaClO are capable for decomposition in wastewater. Li et al. [24] demonstrated that the ferrate (VI)–hypochlorite liquid mixture maintained a high decolorization efficiency for initial 10 mg/L Orange II within a broad pH range of 3.0 to 11.0. Meanwhile, Fu et al. [25] achieved a near-100% decolorization rate when treating simulated wastewater containing an initial 400 mg/L Reactive Red concentration 195 using a combination of zero-valent iron/activated carbon, microwave discharge electrodeless lamp, and sodium hypochlorite (ZVI/AC-MDEL/NaClO). In recent years, Mn- and Co-based catalysts have received widespread attention [26]. Cobalt oxides stand out for their abundant oxygen vacancies and oxygen storage capacity. The presence of abundant Co2+ indicates the generation of oxygen vacancies in the catalyst, which is beneficial for promoting catalytic activity [27]. The cobalt oxide catalyst immobilized on graphene oxide (GO) was able to activate peroxymonosulfate (PMS), resulting in the complete degradation of 0.2 mM Orange II in water within 6 min [28]. Manganese oxides are widely used in catalysts due to their multiple oxidation states (Mn4+, Mn3+, Mn2+). According to the Mars-van Kreveen model, a large amount of Mn4+ is beneficial for increasing the ratio of lattice oxygen (Olatt) to adsorbed oxygen (Oads), with lattice oxygen playing a positive role in enhancing catalytic activity [29,30]. Zhi et al. [13] prepared MnO2–mullite–cordierite composite particles and utilized NaClO as the oxidant for the decolorization of methylene blue (MB). The experimental results indicated that this system could achieve a near-100% decolorization rate for 100 mg/L MB.
Nano-spinel (XY2O4, X/Y are transition metal ions) catalysts have attracted widespread attention due to the excellent synergistic effect between the two metal ions and the low leaching of metal ions in aqueous solutions [31,32]. Among them, CoMn2O4 has been extensively studied due to its structural stability, low cost, environmental friendliness, and excellent catalytic activity. Nguyen et al. [33] reported a three-dimensional flower-like CoMn2O4 catalyst with high-performance in Fenton-like process. With the synergistic effect of Co and Mn, abundant oxygen vacancies were generated, accelerating electron transfer and greatly improving the catalytic performance for tetracycline degradation. Xu et al. [34] revealed a novel carbon-doped CoMn2O4/Mn3O4 composite in activating PMS for ciprofloxacin degradation, exhibiting excellent catalytic performance and good recyclability. Yang et al. [35] reported an efficient catalyst with CoMn2O4 uniformly dispersing on the surface of halloysite to expose more active sites. Nevertheless, it has rarely been reported that a heterogeneous activation of NaClO by the nano-CoMn2O4 spinel for the decomposition of organic pollutants though CoMn2O4 has been employed in wastewater treatment.
Synthetic dyes are widely used in various industries, such as textiles, cosmetics, printing, pharmaceuticals, and food processing, with approximately 7 × 105 tons of different types of dyes and pigments being produced globally each year [36]. Among these, azo dyes, which have an -N=N- unit in their molecular structure, account for 60–70% of all synthetic dyes produced [37]. It is reported that the textile industry alone consumes over 10 billion liters of water and 800,000 tons of dyes daily, with 120,000 tons ultimately entering wastewater [38]. Azo dyes and their decomposition products are highly toxic due to their carcinogenic, teratogenic, and mutagenic properties, which can not only degrade water quality but also cause adverse reactions, such as genetic mutations, skin inflammation, allergies, and skin irritation, when humans are exposed to synthetic azo dyes [39]. Methylene blue (MB), a heterocyclic aromatic azo dye, is a highly carcinogenic thiazine contaminant and one of the most widely used dyes in various industries [40]. MB has a complex structure and is resistant to biodegradation, posing potential harm to ecosystems and being difficult to decolorize [41,42]. Once ingested by humans, it can cause various injuries, including to the nervous system and eyes [36].
Herein, in this work, we systematically investigated the decolorization of methylene blue (MB) by the CoMn2O4-activated NaClO process. The performance of MB decolorization and impacted parameters in the CoMn2O4/NaClO system were investigated. The primary active species for MB oxidation was elucidated. Finally, possible intermediates during MB degradation process were identified, and a potential degradation pathway for MB was proposed.

2. Results and Discussion

2.1. Characterization of CoMn2O4

Thermogravimetric analysis was carried out from room temperature to 800 °C in a static air atmosphere to evaluate the thermal properties of the catalyst (Figure 1). The first mass loss was about 6.0% below 400 °C, which resulted from the evaporation of anhydrous ethanol, crystal water in the metallic precursor, and a little water [43]. The second mass loss was 8.0% (400–500 °C), which might be related to the conversion of inorganic cobalt to organic cobalt [44] and inorganic manganese to organic manganese [45]. The third mass loss was 2.5% (685–800 °C), which was associated with the phase distortion of organic cobalt and organic manganese due to the gradual loss of excess oxygen [45,46]. Therefore, an annealing temperature of 600 °C was employed.
To verify against the pure CoMn2O4 spinel, the prepared catalyst was characterized using XRD. As shown in Figure 2a, the diffraction peaks at 18.44°, 29.50°, 33.14°, 36.64°, 50.90°, 56.90°, 60.94°, and 75.18° of the prepared materials agreed well with the peaks of the pure CoMn2O4 according to the standard card CoMn2O4 (PDF#44-0141), indicating that the CoMn2O4 spinel was successfully synthesized [47]. The diffraction peaks at 2θ = 18.44°, 36.64°, 45.02°, 54.64°, 59.26°, and 65.42° fitted well with Co3O4. At the same time, the diffraction peaks of MnO2 at 2θ were 36.64°, 39.04°, 56.90°, 59.26°, and 65.42°, respectively [33]. At the degrees of 29.86°, 33.02°, 36.20°, and 62.26°, the materials exhibited high diffraction intensity and sharp shape, implying the high crystallinity of the catalysts and excellent arrangement of internal particles [48].
Figure 2b presents the FT-IR spectrum of the CoMn2O4 spinel. The peak at 3423 cm−1 can be attributed to the O-H stretching vibration of associated OH groups on the surface of the catalyst [49]. The peaks at 991, 633, and 533 cm−1 were attributed to the lattice vibration of cobalt-based spinel and the stretching vibrations of Co-O and Mn-O, respectively [50,51]. The peak at 430 cm−1 can be referred to the characteristic vibration of the Co-O-Mn bond [52]. These peaks confirmed the formation of spinel CoMn2O4 after calcination.
Figure 3a–d depict the SEM images of the CoMn2O4 spinel at different magnifications and the particle size distribution histogram of SEM. Figure 3a revealed that the morphology of the CoMn2O4 spinel consists of layered particles with a rough surface, which potentially increase its specific surface areas and the active sites between pollutant molecules and the catalysts, facilitating the catalytic reactions [48]. Figure 3b–d demonstrate that the CoMn2O4 catalyst particles were uniformly distributed with a loose and rough structure, whose average particle size was found to be 125 nm (approx.) from the particle size distribution histogram, exhibiting a porous structure that can provide more extensive active reaction sites [53,54].
The TEM images and the particle size distribution histogram of the TEM of CoMn2O4 are presented in Figure 3e–h. Figure 3e–g illustrated that CoMn2O4 was composed of layered nanosheets, whose average particle size was found to be 105 nm (approx.) from the particle size distribution histogram, whose result was similar to SEM. In Figure 3h, abundant lattice fringes were observed, with lattice spacings of d = 0.251 nm corresponding to the (311) plane of CoMn2O4, and d = 0.478 nm related to the (111) plane of CoMn2O4.
The specific surface area and pore size distribution were obtained through N2 adsorption–desorption isotherms by Bruner–Emmett–Teller (Figure 4). The nitrogen physisorption analysis revealed that the specific surface area of CoMn2O4 was about 6.7865 m2/g, with a pore volume of 0.01939 cm3/g. CoMn2O4 showed typical type-IV isotherms with hysteresis loops. The hysteresis was suggestive of capillary condensation [55], indicating they were mesoporous materials (2–50 nm) and the size of these pores to be about 19.7 nm. The electrochemical property of spinel is stable in an aqueous environment and under illumination with a mechanism conduction by electron hopping [56]. Thus, based on the BET analysis, the quantity and capability of mesoporous CoMn2O4 were able to accommodate some molecular magnitude and broaden pathways for ion transport [55]. The mesoporous properties and surface area could provide CoMn2O4 with available active sites, which could provide opportunities for the catalyst to be in full contact with oxygen and electrolyte, contributing to its enhanced catalytic activities and the desorption of products to accelerate the reaction rate [57,58].
XPS is the primary analysis of the elemental composition and valence states on the surface of the materials. Figure 5 depicts the XPS spectra of the prepared CoMn2O4 spinel. Figure 5a displays the full XPS spectrum of the CoMn2O4 catalyst, revealing only Mn, Co, and O elements in the prepared materials with no observed impurities other than carbon. This result further confirmed that the product calcinated at 600 °C was pure CoMn2O4.
In Figure 5b, the O 1s spectrum can be deconvoluted into three peaks. The peaks centered at 530.12 eV, 531.53 eV, and 533.05 eV were identified as Olatt (lattice oxygen), Oads (adsorbed oxygen), and Osur (surface adsorbed OH groups or molecular water) [34,59]. Olatt was the main oxygen species in the CoMn2O4 sample, which was due to the formation of the Co-Mn solid solution, leading to lattice defects, exposing more active sites, and increasing the concentration of lattice oxygen. The increase in the lattice oxygen content contributed to the catalytic activity of the sample [60]. Furthermore, the atomic ratio of Olatt/Oads affected the catalytic oxidation of organic pollutants. The Olatt/Oads atomic ratio of the CoMn2O4 nano-spinel was as high as 2.69, indicating high catalytic oxidation performance [59].
In the high-resolution Co 2p spectrum in Figure 5c, the Co 2p spectrum was fitted with a pair of spin-orbit peaks (Co 2p3/2 and Co 2p1/2). Peaks corresponding to Co3+ at the Co 2p3/2 and Co 2p1/2 energy levels were observed at 780.34 eV and 794.67 eV, respectively. Additionally, peaks associated with Co2+ at the Co 2p3/2 and Co 2p1/2 energy levels were identified at 781.42 eV and 797.21 eV, respectively. The findings indicated that Co exists in mixed valence states of Co3+ and Co2+ [60]. The presence of low-valence Co2+ may indicate the generation of oxygen vacancies in the nanomaterials, which was beneficial for promoting catalytic activity [48]. The peaks located at 784.39 eV, 786.93 eV, and 802.94 eV were attributed to satellite peaks [61].
As displayed in Figure 5d, the high-resolution Mn 2p spectrum was also fitted with a pair of spin-orbit peaks (Mn 2p3/2 and Mn 2p1/2). The peaks at 640.74 eV, 641.63 eV, and 643.22 eV corresponding to Mn 2p3/2 were assigned to Mn2+, Mn3+, and Mn4+, respectively. Furthermore, the peak of Mn 2p3/2 could be deconvoluted into three sub-peaks at 652.64 eV, 653.17 eV, and 654.44 eV, which were also attributed to Mn2+, Mn3+, and Mn4+, respectively [48,62]. It has been reported that redox reactions produced oxygen vacancies, and a higher concentration of Mn3+ was considered to indicate the presence of more oxygen vacancies [48,63].

2.2. Catalytic Performance of CoMn2O4

The calcination temperature not only affected the structural properties of the nanomaterials, but also influenced the stability of the catalysts. Figure 6a illustrates the impact of calcination temperature on MB decolorization and the leaching of metal ions from the nano-spinel. As the calcinating temperature increased from 300 °C to 600 °C, the MB decolorization rate increased from 94.3% to 98.7%. When the temperature further increased to 800 °C, the degradation rate was 96.9%. This is because at lower temperatures, Co and Mn salts cannot fully decompose into their respective oxides, resulting in lower catalyst purity and poorer degradation performance. At higher temperatures, the bonding state of metal oxides in the catalyst may change, leading to loss of activity.
The leaching of Mn and Co decreased from 4.25 mg/L and 2.95 mg/L to 0.25 mg/L and 0.18 mg/L, respectively, as the calcination temperature increased from 300 °C to 800 °C. Higher calcination temperatures better remove water and crystallization reactions in the catalyst, producing more stable metal oxides and reducing heavy metal leaching. Considering these factors, 600 °C was chosen as the optimal calcination temperature.
Figure 6b depicts the MB decolorization efficiency in different systems. When only the CoMn2O4 nano-catalyst was added, the MB decolorization rate was merely 3.7%, indicating weak adsorption of MB by CoMn2O4, consistent with the findings of Guo, Li, and Fan [53,62,64]. When NaClO was used alone, the MB decolorization efficiency was 22.1%, which can be ascribed to the strong oxidative properties of NaClO [65]. With the involvement of Co3O4 and MnO2, the decolorization of MB was both promoted with efficiencies of 37.7% and 27.6%, respectively, which was due to the activation of NaClO by metal oxides [13]. The CoMn2O4/NaClO system exhibited greatest catalytic activity, with an MB decolorization rate of 98.7%, demonstrating that the addition of the CoMn2O4 enhanced NaClO oxidation. This is attributed to the synergistic effect between metallic compositions in the nano-spinel and ClO species [13]. In Table 1, the previous studies on the degradation of methylene blue by the corresponding catalyst or sodium hypochlorite were provided. It is also proved that the spinel catalyst CoMn2O4 has the ability to rapidly degrade methylene blue in a low concentration of NaClO environment at a lower dose.

2.3. Effect of Parameters

Figure 7a illustrates the effect of CoMn2O4 dosage (0–2 g/L) on MB decolorization efficiency. As the CoMn2O4 dosage increased from 0 g/L to 1 g/L, the MB degradation rate increased from 43.8% to 98.67%. This significant improvement may be due to the increased catalytic metal compositions and growing number of active sites with the increasement of nano-spinel [53]. Further increasing the nano-spinel dosage to 2 g/L did not provide additional benefits. Thus, 1 g/L was selected as the optimal catalyst dosage, which is significantly lower than those reported in previous studies [13,53], potentially reducing large-scale application costs [15].
Figure 7b shows the effect of the initial effective chlorine concentration on MB degradation efficiency. As the effective chlorine concentration increased from 0.05% to 0.2%, the MB decolorization efficiency improved significantly due to enhanced active oxygen and hypochlorite content [53,67]. In addition, with higher NaClO, the amount favors NaClO molecules into the active sites, thus producing more active oxidative species [68]. However, when the concentration increased to 0.5%, the decolorization efficiency decreased dramatically. This may be due to (1) a scavenging effect similar to H2O2 scavenging hydroxyl radicals [13,69] and (2) slower NaClO decomposition caused by high pH environment in the presence of NaOH in commercial NaClO.
Figure 7c demonstrates the MB decolorization efficiency at different initial MB concentrations (5–50 mg/L). Although the degradation rate decreased slightly with increasing MB concentration, the degradation efficiency remained above 98.5% after 40 min for all concentrations. This indicates that the CoMn2O4/NaClO system can achieve deep decolorization of MB even at high substrate concentrations. Therefore, MB of 50 mg/L was used in the subsequent experiments.
Figure 7d depicts the impact of initial pH (4.3–9.3) on MB degradation. As pH increased from 4.3 to 9.3, the MB decolorization rate ranged narrowly around 95.0%, revealing the good adaptability to pH of the CoMn2O4/NaClO process. The leaching of Mn and Co decreased from 0.26 mg/L and 0.21 mg/L to 0.11 mg/L and 0.09 mg/L, respectively, as the initial pH increased from 4.3 to 9.3. Although the metal precipitation rate was positively correlated with the degradation efficiency, the initial pH value had little effect on the degradation efficiency of MB. This probably can be attributed to the findings that the pH value of the solution maintained at around 10.0 after addition of NaClO, and the pH values varied narrowly during the reactions, although the initial pH value ranged in the range of 4.3–9.3. Guo reported that MB oxidation was accelerated at a lower pH circumstance in the Ni-Fe bimetallic catalyst/NaClO system, which was explained by the stronger oxidative capacity of HOCl derived from NaClO under acidic conditions [53]. This finding in the current study implied that the dominant active species and mechanism for MB oxidation probably differed from Guo’s study, the species corresponded to HOCl (e.g., 1O2, ClO), and acidic-induced species probably were not responsible for MB oxidation in this case.
Finally, complete oxidative decolorization of MB within 40 min was obtained at an initial MB concentration of 50 mg/L, CoMn2O4 dosage of 1 g/L, effective chlorine dose of 0.1%, and initial pH of 4.3. Under these circumstances, the spectral evolution of MB during the reaction was recorded in Figure 8a. The significant decrease in the characteristic absorption band intensity (λmax = 662 nm) and a shift over time indicated that MB underwent deep decolorization and the molecular structure of MB was destroyed.
Oxidative performance of various ubiquitous pollutants by the CoMn2O4/NaClO system under the optimal conditions was further explored, and the results were shown in Figure 8b. It can be seen that removals of 99.2%, 92.8%, 91.5%, and 90.7% were achieved within 40 min for Rhodamine B (RB), tetracycline hydrochloride (TC), phenol (PH), and norfloxacin (FPA), respectively, revealing the remarkable oxidation capacity of the CoMn2O4/NaClO process.

2.4. Mechanism Analysis

It has been reported that in NaClO-involved heterogeneous AOPs, active species such as O2•−, HO, and 1O2 were inducible and predominant for decomposition of organic contaminants. To identify the main active species, the commonly recognized radical quenchers tert-butanol (TBA), benzoquinone (BQ), and furfuryl alcohol (FFA) were employed for capturing HO, O2•−, and 1O2, respectively [53,70,71]. As shown in Figure 9a, the addition of either 1 mM or 5 mM TBA was only slightly inhibited by the decolorization of MB with the efficiencies reduced from 98.1% to 96.6% and 94.5%, respectively. However, the addition of benzoquinone played a prominent inhibitory effect on MB decolorization, reducing the removal of MB from 98.1% to 57.8% with 1 mM BQ and 44.8% with 5 mM BQ. The results indicated that HO did not contribute to MB decolorization; however, O2•− or O2•− derived 1O2 (Equation (2)) [72] probably was responsible for MB decolorization.
O2•− + HO1O2 + OH,
To identify the contribution of 1O2, 5 mM FFA was introduced to the CoMn2O4/NaClO system. Interestingly, the decolorization of MB barely diminished (95.6%), implying the negligible contribution of 1O2 to MB removal. The confirmation of the neglected contribution of 1O2 verified the deduction that acidic-induced species (e.g., HOCl, 1O2, ClO) were not contributive to MB removal, as discussed in the section on the impact of the initial pH. To further confirm the presence of O2•− in the CoMn2O4/NaClO system, an electron paramagnetic resonance (EPR) analysis was conducted, and the results were shown in Figure 9b. No obvious signal was found by the addition of NaClO alone, and after the addition of both NaClO and CoMn2O4, the typical sextuple EPR spectrum corresponding to O2•− was observed, indicating that O2•− was produced in the CoMn2O4/NaClO system [65,73].
Mn4+/Mn3+ + ClO → Mn3+/Mn2+ + O2•− + Cl,
Co3+ + ClO → Co2+ + O2•− + Cl,
O2 + e → O2•−,
Based on the above findings, the possible decolorization mechanism of MB in the CoMn2O4/NaClO system is proposed. Mn4+, Mn3+, and Co3+ from the nano-spinel CoMn2O4 may activate NaClO to directly produce O2•− for catalytic decomposition of MB (Equations (3) and (4) [74,75,76]. Additionally, the small amount of generated O2 via Equation (1) could capture electrons donated by oxygen vacancies to form O2•− (Equation (5)). Finally, MB molecules were attacked, followed by the breakage of structural bonds to form intermediate products, and then were mineralized into molecules of H2O and CO2 [13].

2.5. Proposed MB Decolorization Pathway

High-Performance Liquid Chromatography-Mass Spectrometry (HPLC-MS) analyses were used to identify intermediates formed during MB decolorization. As illustrated in Figure 10a, products of A (C16H21N3OS), B (C8H8N2O2), C (C6H5NaO3S), and D (C8H6N2O4) were presented in the full-scan MS spectra. According to these detected intermediates, Figure 10b presents possible degradation pathways for MB. The initial oxidation occurred at the sulfur site, producing intermediate A [77]. Product A was further oxidized to generate intermediates B, C, and D in the sequential reaction. The dimethylamino group in MB undergone progressive transformation to methyl, aldehyde, and nitro groups, before further mineralization to ammonium (NH4+), sulfate (SO42−), H2O, and CO2. Some studies reported some small organic acid products during the degradation process [78], which was not detected in this study, probably due to the fact that the primary products had not been mineralized to this extent within limited reaction time (10 min) of the sample. The cracking mode and the product situation of the benzene ring in the late degradation stage of MB should be further explored.

2.6. The Reusability of CoMn2O4

The reusability of nano-CoMn2O4 was further investigated as illustrated in Figure 11a. The nano-spinel maintained high catalytic activity even after five cycles of use, achieving over 90% MB decolorization rate. This result was consistent with the findings of Guo and Zhang [53,79]. The slight decrease in catalytic activity of CoMn2O4 may be due to partial coverage of active sites on the surface of nanomaterials by adsorbed species, leading to pore blockage [80,81] and the leaching of Mn and Co atoms, resulting in the loss of active sites [82,83,84].
The specific surface area and pore size distribution of the catalyst after five rounds of AOPs were obtained using N2 adsorption–desorption isotherms (Figure 11b). Nitrogen physical adsorption analysis revealed that the specific surface area of the catalyst was reduced to about 6.41327 m2/g, and the pore volume was 0.01843 cm3/g. The hysteresis phenomenon was still presented, and the size of these holes was reduced to about 18.6 nm. This result confirmed the conjecture of repeated experiments.

3. Materials and Methods

3.1. Chemicals

Cobalt nitrate hexahydrate (Co(NO3)2·6H2O), manganese acetate tetrahydrate (Mn(CH3COO)2·4H2O), phenol (PH), benzoquinone (BQ), tert-butanol (TBA), and norfloxacin (FPA) were purchased from Shanghai MacLean Biochemical Technology Co., Ltd. (Shanghai, China). Oxalic acid (H2C2O4) and 10% sodium hypochlorite (NaClO) were obtained from Tianjin Tianli Chemical Reagent Co., Ltd. (Tianjin, China). Absolute ethanol, methylene blue (MB), and rhodamine B (RB) were purchased from Sinophaceutical Chemical Reagent Co., Ltd. (Beijing, China). Tetracycline hydrochloride (TC) was purchased from Shanghai Yuanyuan Biotechnology Co., Ltd. (Shanghai, China). Deionized water was used for all experiments except for studying different water matrices.

3.2. Synthesis of CoMn2O4

The CoMn2O4 spinel was prepared using a coprecipitation pyrolysis method. Firstly, 20 mL anhydrous ethanol solutions of 0.5 M cobalt nitrate hexahydrate and 1 M manganese acetate tetrahydrate were mixed in a conical flask with vigorous stirring at 60 °C in a water bath until completely dissolved. After addition of 150 mL of 0.24 M oxalic acid, the mixture was stirred for an additional 40 min at 60 °C. The mixture was then separated by centrifugalizing, and the obtained precipitates were washed alternately with anhydrous ethanol and deionized water three times, followed by drying the gained solid at 60 °C for 24 h. The dried solid was further calcined in a muffle furnace at 600 °C with a heating rate of 1 °C min−1 and maintained for 3 h. Finally, after cooling to room temperature, the obtained materials were ground and sieved through a 40–60 mesh screen for future use.

3.3. Experimental Procedures

Batch experiments of MB decolorization were conducted in a 1 L beaker filled with MB with a desired concentration. A default amount of NaClO and the prepared CoMn2O4 were added to the beaker while stirring on a magnetic stirrer at 500 rpm and room temperature. At predetermined time intervals, samples were withdrawn and filtered through a 0.45 μm membrane for further analysis. The initial pH of MB solution was regulated with diluted H2SO4 and NaOH if needed.

3.4. Characterization

The thermal analysis was studied using a SDT Q600 V20.9 Build 20 module DSC-TGA standard equipment (TA Instruments, New Castle, DE, USA) at a heating-temperature-ramping rate of 10 °C min−1 in the temperature range of 20–800 °C under air atmosphere. The specific surface area and porosity were determined using Micromeritics ASAP 2460 automatic (Micromeritics Instrument Corporation, Norcross, GA, USA) specific surface area and porosity analyzer (USA). The X-ray Diffraction (XRD) analysis was performed using a Rigaku Smart Lab SE diffractometer (Rigaku Corporation, Tokyo, Japan). The Fourier transform infrared spectroscopy (FT-IR) analysis was conducted using a Thermo Scientific Nicolet 6700 spectrometer (Thermo Fisher Scientific, Waltham, MA, USA). The scanning electron microscopy (SEM) was performed using a TESCAN MIRA LMS microscope (TESCAN, Brno, Czech Republic). The transmission electron microscopy (TEM) was conducted using an FEI Talos F200× microscope (Thermo Fisher Scientific, Waltham, MA, USA). The X-ray photoelectron spectroscopy (XPS) was carried out using a Thermo Scientific K-Alpha XPS instrument (Thermo Fisher Scientific, Waltham, MA, USA).

3.5. Analytic Methods

MB were measured using a UV–vis spectrophotometer (UV 2600) at 662 nm. The leached Mn and Co were determined using an atomic absorption spectrometer (AA-6880). The mass spectrometry analysis was conducted using an Agilent QTOF 6550 (Agilent Technologies, Inc., Santa Clara, CA, USA) with a flow rate of 0.3 mL min−1. The mobile phase consisted of deionized water (A) and methanol (B) in a 95:5 ratio. The mass spectrometer operated in positive ion mode with a 5 μL injection volume. Solution pH was measured using a LEICI PHS-3C pH meter (Shanghai INESA Scientific Instrument Co., Ltd., Shanghai, China).

3.6. Repeat Discoloration Experiment Procedures

After AOPs, the suspension was filtered to obtain catalyst powders. Then, the powders were washed alternately with anhydrous ethanol and deionized water by centrifugation three times, followed by drying the gained solid at 60 °C for 24 h. The dried catalysts were put into the reaction again, whose progress was the same as the first experimental procedure until it was circulated five times. The specific surface area and porosity of the dried powder after the five rounds of cyclic reaction were determined using Micromeritics ASAP 2460 automatic specific surface area and porosity analyzer (Norcross, GA, USA).

4. Conclusions

Nano-spinel CoMn2O4 was successfully synthesized using a coprecipitation pyrolysis method and confirmed via a series of characterizations. The CoMn2O4 exhibited superior catalytic activity for NaClO activation than metal oxides with an MB decolorization efficiency of 98.7% within 40 min. Under optimal conditions at a CoMn2O4 dosage of 1 g/L, an effective chlorine dose of 0.1%, and an initial pH of 4.3, the CoMn2O4/NaClO system was effective in the oxidation of various pollutants with removals higher than 90.7%. O2•− was identified and confirmed to be the main active species for MB degradation. The nano-spinel can be reused for five cycles with slight loss of catalytic activity (less than 10%). This work provided a theoretical reference for the oxidative degradation of organic pollutants using heterogeneous activation of commercial NaClO by nano-CoMn2O4 spinel.

Author Contributions

Conceptualization, X.W.; methodology, G.H.; software, T.Z.; validation, J.B.; formal analysis, T.Z.; investigation, T.Z.; resources, T.Z.; data curation, J.B.; writing—original draft preparation, T.Z. and G.H.; writing—review and editing, X.W.; visualization, T.Z.; supervision, X.W.; project administration, X.W.; funding acquisition, X.W. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by Environmental Protection Research Project of Hubei Provincial De-partment of Environmental Protection (2023ZHB-06), National Yangtze River Ecological Envi-ronment Protection and Restoration Joint Research Center (2022-LHYJ-02-0506-06) and Northwest Water Resources and Environmental Ecology Ministry of Education Key Laboratory Open Fund Project (No. 2022SZY03).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The datasets used and analyzed in this study are available from the corresponding author upon reasonable request.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Reoyo-Prats, B.; Sellier, A.; Khaska, S.; Le Gal, L.C.; Weiss, K.; Goetz, V.; Plantard, G. Continuous degradation of micropollutants in real world treated wastewaters by photooxidation in dynamic conditions. Water Res. 2022, 221, 118777. [Google Scholar]
  2. Jiaqi, B.; Zhiwei, D.; Hui, L.; Tianhao, L.; Yanjing, Y.; Shian, Z. Bimetallic modified halloysite particle electrode enhanced electrocatalytic oxidation for the degradation of sulfanilamide. J. Environ. Manag. 2022, 312, 114975. [Google Scholar]
  3. Bingkun, H.; Zelin, W.; Hongyu, Z.; Jiayi, L.; Chenying, Z.; Zhaokun, X.; Zhicheng, P.; Gang, Y.; Bo, L. Recent advances in single-atom catalysts for advanced oxidation processes in water purification. J. Hazard. Mater. 2021, 412, 125253. [Google Scholar]
  4. Li, W.; Guo, H.; Wang, C.; Zhang, Y.; Cheng, X.; Wang, J.; Yang, B.; Du, E. ROS reevaluation for degradation of 4-chloro-3, 5-dimethylphenol (PCMX) by UV and UV/persulfate processes in the water: Kinetics, mechanism, DFT studies and toxicity evolution. Chem. Eng. J. 2020, 390, 124610. [Google Scholar] [CrossRef]
  5. Lin, H.; Gao, W.; Meng, F.; Liao, B.-Q.; Leung, K.-T.; Zhao, L.; Chen, J.; Hong, H. Membrane bioreactors for industrial wastewater treatment: A critical review. Crit. Rev. Environ. Sci. Technol. 2012, 42, 677–740. [Google Scholar] [CrossRef]
  6. Miao, W.; Liu, Y.; Wang, D.; Du, N.; Ye, Z.; Hou, Y.; Mao, S.; Ostrikov, K.K. The role of Fe-Nx single-atom catalytic sites in peroxymonosulfate activation: Formation of surface-activated complex and non-radical pathways. Chem. Eng. J. 2021, 423, 130250. [Google Scholar] [CrossRef]
  7. Priyadarshini, M.; Das, I.; Ghangrekar, M.M.; Blaney, L. Advanced oxidation processes: Performance, advantages, and scale-up of emerging technologies. J. Environ. Manag. 2022, 316, 115295. [Google Scholar] [CrossRef]
  8. Abd Manan, T.S.B.; Khan, T.; Sivapalan, S.; Jusoh, H.; Sapari, N.; Sarwono, A.; Ramli, R.M.; Harimurti, S.; Beddu, S.; Sadon, S.N. Application of response surface methodology for the optimization of polycyclic aromatic hydrocarbons degradation from potable water using photo-Fenton oxidation process. Sci. Total Environ. 2019, 665, 196–212. [Google Scholar] [CrossRef]
  9. Zhou, H.; Zhang, H.; He, Y.; Huang, B.; Zhou, C.; Yao, G.; Lai, B. Critical review of reductant-enhanced peroxide activation processes: Trade-off between accelerated Fe3+/Fe2+ cycle and quenching reactions. Appl. Catal. B 2021, 286, 119900. [Google Scholar] [CrossRef]
  10. Wang, F.; Xu, J.; Wang, Z.; Lou, Y.; Pan, C.; Zhu, Y. Unprecedentedly efficient mineralization performance of photocatalysis-self-Fenton system towards organic pollutants over oxygen-doped porous g-C3N4 nanosheets. Appl. Catal. B 2022, 312, 121438. [Google Scholar] [CrossRef]
  11. Yuan, Y.; Liu, J.; Gao, B.; Hao, J. Ozone direct oxidation pretreatment and catalytic oxidation post-treatment coupled with ABMBR for landfill leachate treatment. Sci. Total Environ. 2021, 794, 148557. [Google Scholar] [CrossRef] [PubMed]
  12. Wang, H.; Guo, W.; Liu, B.; Wu, Q.; Luo, H.; Zhao, Q.; Si, Q.; Sseguya, F.; Ren, N. Edge-nitrogenated biochar for efficient peroxydisulfate activation: An electron transfer mechanism. Water Res. 2019, 160, 405–414. [Google Scholar] [CrossRef] [PubMed]
  13. Zhi, S.; Tian, L.; Li, N.; Zhang, K. A novel system of MnO2-mullite-cordierite composite particle with NaClO for Methylene blue decolorization. J. Environ. Manag. 2018, 213, 392–399. [Google Scholar] [CrossRef] [PubMed]
  14. Ding, H.; Hu, J. Degradation of carbamazepine by UVA/WO3/hypochlorite process: Kinetic modelling, water matrix effects, and density functional theory calculations. Environ. Res. 2021, 201, 111569. [Google Scholar] [CrossRef]
  15. Subhan, S.; Rahman, A.U.; Yaseen, M.; Rashid, H.U.; Ishaq, M.; Sahibzada, M.; Tong, Z. Ultra-fast and highly efficient catalytic oxidative desulfurization of dibenzothiophene at ambient temperature over low Mn loaded Co-Mo/Al2O3 and Ni-Mo/Al2O3 catalysts using NaClO as oxidant. Fuel 2019, 237, 793–805. [Google Scholar] [CrossRef]
  16. Kirihara, M.; Okada, T.; Sugiyama, Y.; Akiyoshi, M.; Matsunaga, T.; Kimura, Y. Sodium hypochlorite pentahydrate crystals (NaOCl 5H2O): A convenient and environmentally benign oxidant for organic synthesis. Org. Process Res. Dev. 2017, 21, 1925–1937. [Google Scholar] [CrossRef]
  17. Lau, S.S.; Abraham, S.M.; Roberts, A.L. Chlorination revisited: Does Cl–serve as a catalyst in the chlorination of phenols? Environ. Sci. Technol. 2016, 50, 13291–13298. [Google Scholar] [CrossRef]
  18. Pan, X.; Wei, J.; Zou, M.; Chen, J.; Qu, R.; Wang, Z. Products distribution and contribution of (de) chlorination, hydroxylation and coupling reactions to 2, 4-dichlorophenol removal in seven oxidation systems. Water Res. 2021, 194, 116916. [Google Scholar] [CrossRef]
  19. Zhang, J.; Chen, Y.; Liao, Y.; Wang, Q.; Yu, J. Studies on the degradation of trace phenol and indole odorants by chlorine and permanganate in drinking water treatment. Chemosphere 2022, 286, 131551. [Google Scholar] [CrossRef]
  20. Sedlak, D.L.; von Gunten, U. The chlorine dilemma. Science 2011, 331, 42–43. [Google Scholar] [CrossRef]
  21. He, T.; Deng, L.; Lai, B.; Xu, S.; Wang, L.; Zhang, Y.; Zheng, D.; Hu, C. The performance of Fe2+/ClO system in advanced removal of fulvic acid under mild conditions. J. Environ. Chem. Eng. 2022, 10, 107515. [Google Scholar] [CrossRef]
  22. Li, H.; Lin, M.; Xiao, T.; Long, J.; Liu, F.; Li, Y.; Liu, Y.; Liao, D.; Chen, Z.; Zhang, P. Highly efficient removal of thallium (I) from wastewater via hypochlorite catalytic oxidation coupled with adsorption by hydrochar coated nickel ferrite composite. J. Hazard. Mater. 2020, 388, 122016. [Google Scholar] [CrossRef] [PubMed]
  23. Kim, K.-W.; Lee, E.-H.; Chung, D.-Y.; Moon, J.-K.; Shin, H.-S.; Kim, J.-S.; Shin, D.-W.J.C.E.J. Manufacture characteristics of metal oxide–hydroxides for the catalytic decomposition of a sodium hypochlorite solution. Chem. Eng. J. 2012, 200, 52–58. [Google Scholar] [CrossRef]
  24. Li, G.; Wang, N.; Liu, B.; Zhang, X. Decolorization of azo dye Orange II by ferrate (VI)–hypochlorite liquid mixture, potassium ferrate (VI) and potassium permanganate. Desalination 2009, 249, 936–941. [Google Scholar] [CrossRef]
  25. Fu, J.; Xu, Z.; Li, Q.-S.; Chen, S.; An, S.-Q.; Zeng, Q.-F.; Zhu, H.-L. Treatment of simulated wastewater containing Reactive Red 195 by zero-valent iron/activated carbon combined with microwave discharge electrodeless lamp/sodium hypochlorite. J. Environ. Sci. 2010, 22, 512–518. [Google Scholar] [CrossRef]
  26. Xiong, S.; Huang, N.; Peng, Y.; Chen, J.; Li, J. Balance of activation and ring-breaking for toluene oxidation over CuO-MnOx bimetallic oxides. J. Hazard. Mater. 2021, 415, 125637. [Google Scholar] [CrossRef]
  27. Zhang, C.; Wang, J.; Yang, S.; Liang, H.; Men, Y. Boosting total oxidation of acetone over spinel MCo2O4 (M = Co, Ni, Cu) hollow mesoporous spheres by cation-substituting effect. J. Colloid Interface Sci. 2019, 539, 65–75. [Google Scholar] [CrossRef]
  28. Shi, P.; Su, R.; Zhu, S.; Zhu, M.; Li, D.; Xu, S. Supported cobalt oxide on graphene oxide: Highly efficient catalysts for the removal of Orange II from water. J. Hazard. Mater. 2012, 229, 331–339. [Google Scholar] [CrossRef]
  29. Li, K.; Chen, C.; Zhang, H.; Hu, X.; Sun, T.; Jia, J. Effects of phase structure of MnO2 and morphology of δ-MnO2 on toluene catalytic oxidation. Appl. Surf. Sci. 2019, 496, 143662. [Google Scholar] [CrossRef]
  30. Wang, F.; Dai, H.; Deng, J.; Bai, G.; Ji, K.; Liu, Y. Manganese oxides with rod-, wire-, tube-, and flower-like morphologies: Highly effective catalysts for the removal of toluene. Environ. Sci. Technol. 2012, 46, 4034–4041. [Google Scholar] [CrossRef]
  31. Deng, J.; Ya, C.; Ge, Y.; Cheng, Y.; Chen, Y.; Xu, M.; Wang, H. Activation of peroxymonosulfate by metal (Fe, Mn, Cu and Ni) doping ordered mesoporous Co3O4 for the degradation of enrofloxacin. RSC Adv. 2018, 8, 2338–2349. [Google Scholar] [CrossRef] [PubMed]
  32. Qiu, S.; Gou, L.; Cheng, F.; Zhang, M.; Guo, M. An efficient and low-cost magnetic heterogenous Fenton-like catalyst for degrading antibiotics in wastewater: Mechanism, pathway and stability. J. Environ. Manag. 2022, 302, 114119. [Google Scholar] [CrossRef] [PubMed]
  33. Nguyen, T.-B.; Huang, C.; Doong, R.-a.; Wang, M.-H.; Chen, C.-W.; Dong, C.-D. Manipulating the morphology of 3D flower-like CoMn2O4 bimetallic catalyst for enhancing the activation of peroxymonosulfate toward the degradation of selected persistent pharmaceuticals in water. Chem. Eng. J. 2022, 436, 135244. [Google Scholar] [CrossRef]
  34. Xu, J.; Wang, Y.; Wan, J.; Wang, L. Facile synthesis of carbon-doped CoMn2O4/Mn3O4 composite catalyst to activate peroxymonosulfate for ciprofloxacin degradation. Sep. Purif. Technol. 2022, 287, 120576. [Google Scholar] [CrossRef]
  35. Yang, X.; Wei, G.; Wu, P.; Liu, P.; Liang, X.; Chu, W. Novel halloysite nanotube-based ultrafine CoMn2O4 catalyst for efficient degradation of pharmaceuticals through peroxymonosulfate activation. Appl. Surf. Sci. 2022, 588, 152899. [Google Scholar] [CrossRef]
  36. Din, M.I.; Khalid, R.; Najeeb, J.; Hussain, Z. Fundamentals and photocatalysis of methylene blue dye using various nanocatalytic assemblies-a critical review. J. Cleaner Prod. 2021, 298, 126567. [Google Scholar] [CrossRef]
  37. Aragaw, T.A. Potential and prospects of reductases in azo dye degradation: A minireview. Microbe 2024, 4, 100162. [Google Scholar] [CrossRef]
  38. Hien, S.A.; Trellu, C.; Oturan, N.; Assémian, A.S.; Briton, B.G.H.; Drogui, P.; Adouby, K.; Oturan, M.A. Comparison of homogeneous and heterogeneous electrochemical advanced oxidation processes for treatment of textile industry wastewater. J. Hazard. Mater. 2022, 437, 129326. [Google Scholar] [CrossRef]
  39. Goswami, D.; Mukherjee, J.; Mondal, C.; Bhunia, B. Bioremediation of azo dye: A review on strategies, toxicity assessment, mechanisms, bottlenecks and prospects. Sci. Total Environ. 2024, 954, 176426. [Google Scholar] [CrossRef]
  40. Misran, E.; Supardan, M.D.; Iryani, D.A.; Pramananda, V.; Sihombing, A.F.; Sitorus, D.V. Ultrasonic assisted adsorption of methylene blue using blood clam shell as a low-cost adsorbent. Results Eng. 2024, 23, 102715. [Google Scholar] [CrossRef]
  41. El-Sharkawy, E.; Soliman, A.Y.; Al-Amer, K.M. Comparative study for the removal of methylene blue via adsorption and photocatalytic degradation. J. Colloid Interface Sci. 2007, 310, 498–508. [Google Scholar] [CrossRef] [PubMed]
  42. Ye, Y.; Bruning, H.; Yntema, D.; Mayer, M.; Rijnaarts, H. Homogeneous photosensitized degradation of pharmaceuticals by using red light LED as light source and methylene blue as photosensitizer. Chem. Eng. J. 2017, 316, 872–881. [Google Scholar] [CrossRef]
  43. Lang, L.; Xu, Z. In situ synthesis of porous Fe3O4/C microbelts and their enhanced electrochemical performance for lithium-ion batteries. ACS Appl. Mater. Interfaces 2013, 5, 1698–1703. [Google Scholar] [CrossRef]
  44. Taghavimoghaddam, J.; Knowles, G.P.; Chaffee, A.L. Preparation and characterization of mesoporous silica supported cobalt oxide as a catalyst for the oxidation of cyclohexanol. J. Mol. Catal. A Chem. 2012, 358, 79–88. [Google Scholar] [CrossRef]
  45. Khan, A.; Zhang, K.; Taraqqi-A-Kamal, A.; Wang, X.; Chen, Y.; Zhang, Y. Degradation of antibiotics in aqueous media using manganese nanocatalyst-activated peroxymonosulfate. J. Colloid Interface Sci. 2021, 599, 805–818. [Google Scholar] [CrossRef]
  46. Yun, W.-C.; Lin, K.-Y.A.; Tong, W.-C.; Lin, Y.-F.; Du, Y. Enhanced degradation of paracetamol in water using sulfate radical-based advanced oxidation processes catalyzed by 3-dimensional Co3O4 nanoflower. Chem. Eng. J. 2019, 373, 1329–1337. [Google Scholar] [CrossRef]
  47. Cai, F.; Sun, C.; Sun, Z.; Lai, Y.; Ding, H. Sulfur-functionalized CoMn2O4 as a Fenton-like catalyst for the efficient rhodamine B degradation. Appl. Surf. Sci. 2023, 623, 157044. [Google Scholar] [CrossRef]
  48. Li, H.; Li, X.; Li, Y.; Guo, M. Novel CoMn2O4 as a highly efficient catalyst for the oxidation of o-, m-, p-xylene: Preparation and kinetic study. Mol. Catal. 2022, 528, 112482. [Google Scholar] [CrossRef]
  49. Deborde, M.; Von Gunten, U. Reactions of chlorine with inorganic and organic compounds during water treatment—Kinetics and mechanisms: A critical review. Water Res. 2008, 42, 13–51. [Google Scholar] [CrossRef]
  50. Li, C.-X.; Chen, C.-B.; Lu, J.-Y.; Cui, S.; Li, J.; Liu, H.-Q.; Li, W.-W.; Zhang, F. Metal organic framework-derived CoMn2O4 catalyst for heterogeneous activation of peroxymonosulfate and sulfanilamide degradation. Chem. Eng. J. 2018, 337, 101–109. [Google Scholar] [CrossRef]
  51. Li, L.; Lai, C.; Huang, F.; Cheng, M.; Zeng, G.; Huang, D.; Li, B.; Liu, S.; Zhang, M.; Qin, L. Degradation of naphthalene with magnetic bio-char activate hydrogen peroxide: Synergism of bio-char and Fe–Mn binary oxides. Water Res. 2019, 160, 238–248. [Google Scholar] [CrossRef] [PubMed]
  52. Luo, Y.; Zheng, Y.; Zuo, J.; Feng, X.; Wang, X.; Zhang, T.; Zhang, K.; Jiang, L. Insights into the high performance of Mn-Co oxides derived from metal-organic frameworks for total toluene oxidation. J. Hazard. Mater. 2018, 349, 119–127. [Google Scholar] [CrossRef] [PubMed]
  53. Guo, X.; Chen, S.; Liu, Z.; Yang, C.; Chen, W. Catalytic oxidation of methylene blue by using Ni-Fe bimetallic catalyst/NaClO system: Performance, kinetics, mechanism, and DFT calculations. Sep. Purif. Technol. 2023, 306, 122612. [Google Scholar] [CrossRef]
  54. Wu, Q.; Siddique, M.S.; Yu, W. Iron-nickel bimetallic metal-organic frameworks as bifunctional Fenton-like catalysts for enhanced adsorption and degradation of organic contaminants under visible light: Kinetics and mechanistic studies. J. Hazard. Mater. 2021, 401, 123261. [Google Scholar] [CrossRef]
  55. Woodward, R.T.; Fam, D.W.; Anthony, D.B.; Hong, J.; McDonald, T.O.; Petit, C.; Shaffer, M.S.; Bismarck, A. Hierarchically porous carbon foams from pickering high internal phase emulsions. Carbon 2016, 101, 253–260. [Google Scholar] [CrossRef]
  56. Kihal, I.; Douafer, S.; Lahmar, H.; Rekhila, G.; Hcini, S.; Trari, M.; Benamira, M. CoMn2O4: A new spinel photocatalyst for hydrogen photo-generation under visible light irradiation. J. Photochem. Photobiol. A 2024, 446, 115170. [Google Scholar] [CrossRef]
  57. Li, Z.; He, H.; Cao, H.; Sun, S.; Diao, W.; Gao, D.; Lu, P.; Zhang, S.; Guo, Z.; Li, M. Atomic Co/Ni dual sites and Co/Ni alloy nanoparticles in N-doped porous Janus-like carbon frameworks for bifunctional oxygen electrocatalysis. Appl. Catal. B 2019, 240, 112–121. [Google Scholar] [CrossRef]
  58. Zhang, C.-L.; Liu, J.-T.; Li, H.; Qin, L.; Cao, F.-H.; Zhang, W. The controlled synthesis of Fe3C/Co/N-doped hierarchically structured carbon nanotubes for enhanced electrocatalysis. Appl. Catal. B 2020, 261, 118224. [Google Scholar] [CrossRef]
  59. Wang, Y.; Yang, D.; Li, S.; Zhang, L.; Zheng, G.; Guo, L. Layered copper manganese oxide for the efficient catalytic CO and VOCs oxidation. Chem. Eng. J. 2019, 357, 258–268. [Google Scholar] [CrossRef]
  60. Zhou, A.; Guo, X.; Zhong, S.; Kang, Q.; Chen, M.; Jin, D.; Fan, M.; Zhou, R.; Ma, T. Three birds, one-stone strategy for synthesis of hierarchically arrayed defective MnCo2O4@NF catalyst for photothermal preferential oxidation of CO in H2-rich streams. Chem. Eng. J. 2023, 471, 144835. [Google Scholar] [CrossRef]
  61. Zhao, Q.; Zheng, Y.; Song, C.; Liu, Q.; Ji, N.; Ma, D.; Lu, X. Novel monolithic catalysts derived from in-situ decoration of Co3O4 and hierarchical Co3O4@MnOx on Ni foam for VOC oxidation. Appl. Catal. B 2020, 265, 118552. [Google Scholar] [CrossRef]
  62. Fan, G.; Yao, L.; Wang, Y.; Peng, X.; Xu, J.; Pang, S.; Xu, K.-q.; Du, B.; Chen, J.; Hong, Z. The dual pathway mechanisms of peroxyacetic acid activation by CoMn2O4 spinel for efficient levofloxacin degradation. J. Environ. Chem. Eng. 2023, 11, 109774. [Google Scholar] [CrossRef]
  63. Wu, Y.; Yuan, S.; Feng, R.; Ma, Z.; Gao, Y.; Xing, S. Comparative study for low-temperature catalytic oxidation of o-xylene over doped OMS-2 catalysts: Role of Ag and Cu. Mol. Catal. 2017, 442, 164–172. [Google Scholar] [CrossRef]
  64. Li, J.; Shi, Q.; Zhao, R.; Liu, Y.; Liu, P.; Liu, L. Facile synthesis of CoMn2O4 spinel catalyst as peroxymonosulfate activator for efficient rhodamine B degradation. Mater. Lett. 2023, 351, 135108. [Google Scholar] [CrossRef]
  65. Li, A.; Song, H.; Meng, H.; Lu, Y.; Li, C. Ultrafast desulfurization of diesel oil with ionic liquid based PMoO catalysts and recyclable NaClO oxidant. Chem. Eng. J. 2020, 380, 122453. [Google Scholar] [CrossRef]
  66. Jaafar, A.; Boussaoud, A. Effects of some Advanced Oxidation Processes and Chlorine on disappearance of Methylene Blue. Int. J. Innov. Appl. Stud. 2015, 10, 509–515. [Google Scholar]
  67. Zhang, Y.; Rong, C.; Song, Y.; Wang, Y.; Pei, J.; Tang, X.; Zhang, R.; Yu, K. Oxidation of the antibacterial agent norfloxacin during sodium hypochlorite disinfection of marine culture water. Chemosphere 2017, 182, 245–254. [Google Scholar] [CrossRef]
  68. Sun, Y.; Yang, Z.; Tian, P.; Sheng, Y.; Xu, J.; Han, Y.-F. Oxidative degradation of nitrobenzene by a Fenton-like reaction with Fe-Cu bimetallic catalysts. Appl. Catal. B 2019, 244, 1–10. [Google Scholar] [CrossRef]
  69. Xu, L.; Wang, J. A heterogeneous Fenton-like system with nanoparticulate zero-valent iron for removal of 4-chloro-3-methyl phenol. J. Hazard. Mater. 2011, 186, 256–264. [Google Scholar] [CrossRef]
  70. Guo, Y.; Long, J.; Huang, J.; Yu, G.; Wang, Y. Can the commonly used quenching method really evaluate the role of reactive oxygen species in pollutant abatement during catalytic ozonation? Water Res. 2022, 215, 118275. [Google Scholar] [CrossRef]
  71. Hua, Z.; Guo, K.; Kong, X.; Lin, S.; Wu, Z.; Wang, L.; Huang, H.; Fang, J. PPCP degradation and DBP formation in the solar/free chlorine system: Effects of pH and dissolved oxygen. Water Res. 2019, 150, 77–85. [Google Scholar] [CrossRef] [PubMed]
  72. Yang, X.; Wei, G.; Wu, P.; Liu, P.; Liang, X.; Chu, W. Controlling oxygen vacancies of CoMn2O4 by loading on planar and tubular clay minerals and its application for boosted PMS activation. J. Hazard. Mater. 2022, 436, 129060. [Google Scholar] [CrossRef]
  73. Huang, Q.; Meng, G.; Zhang, X.; Fang, Z.; Yan, Y.; Liao, B.; Zhang, L.; Chen, P. Natural manganese sand activates sodium hypochlorite to enhance ionic organic contaminants removal: Optimization, modeling, and mechanism. Sci. Total Environ. 2023, 866, 161310. [Google Scholar] [CrossRef] [PubMed]
  74. Zhang, L.; Nie, Y.; Hu, C.; Hu, X. Decolorization of methylene blue in layered manganese oxide suspension with H2O2. J. Hazard. Mater. 2011, 190, 780–785. [Google Scholar] [CrossRef]
  75. Zhou, G.; Lan, H.; Wang, H.; Xie, H.; Zhang, G.; Zheng, X. Catalytic combustion of PVOCs on MnOx catalysts. J. Mol. Catal. A Chem. 2014, 393, 279–288. [Google Scholar] [CrossRef]
  76. Kumar, L.; Verma, N.; Tomar, R.; Sehrawat, H.; Kumar, R.; Chandra, R. Development of bioactive 2-substituted benzimidazole derivatives using an MnO x/HT nanocomposite catalyst. Dalton Trans. 2023, 52, 3006–3015. [Google Scholar] [CrossRef]
  77. Lin, J.; Zong, R.; Zhou, M.; Zhu, Y. Photoelectric catalytic degradation of methylene blue by C60-modified TiO2 nanotube array. Appl. Catal. B 2009, 89, 425–431. [Google Scholar] [CrossRef]
  78. Chai, Y.; Cai, Y.; Guan, Y.; Ai, H.; Yang, Z. Enhanced degradation of organic dye in aqueous solutions by bicarbonate-activated hydrogen peroxide with a rosin-based copper catalyst. J. Water Process Eng. 2024, 66, 106035. [Google Scholar] [CrossRef]
  79. Zhang, W.; Tay, H.L.; Lim, S.S.; Wang, Y.; Zhong, Z.; Xu, R. Supported cobalt oxide on MgO: Highly efficient catalysts for degradation of organic dyes in dilute solutions. Appl. Catal. B 2010, 95, 93–99. [Google Scholar] [CrossRef]
  80. Dong, Q.; Chen, Y.; Wang, L.; Ai, S.; Ding, H. Cu-modified alkalinized g-C3N4 as photocatalytically assisted heterogeneous Fenton-like catalyst. Appl. Surf. Sci. 2017, 426, 1133–1140. [Google Scholar] [CrossRef]
  81. Queirós, S.; Morais, V.; Rodrigues, C.S.; Maldonado-Hódar, F.; Madeira, L.M. Heterogeneous Fenton’s oxidation using Fe/ZSM-5 as catalyst in a continuous stirred tank reactor. Sep. Purif. Technol. 2015, 141, 235–245. [Google Scholar] [CrossRef]
  82. Chen, Y.; Cui, K.; Cui, M.; Liu, T.; Chen, X.; Chen, Y.; Nie, X.; Xu, Z.; Li, C.-X. Insight into the degradation of tetracycline hydrochloride by non-radical-dominated peroxymonosulfate activation with hollow shell-core Co@NC: Role of cobalt species. Sep. Purif. Technol. 2022, 289, 120662. [Google Scholar] [CrossRef]
  83. Lyu, Z.; Xu, M.; Wang, J.; Li, A.; Corvini, P.F.-X. Hierarchical nano-vesicles with bimetal-encapsulated for peroxymonosulfate activation: Singlet oxygen-dominated oxidation process. Chem. Eng. J. 2022, 433, 133581. [Google Scholar] [CrossRef]
  84. Wang, H.; Wang, H.; Yan, Q. Peroxymonosulfate activation by algal carbocatalyst for organic dye oxidation: Insights into experimental and theoretical. Sci. Total Environ. 2022, 816, 151611. [Google Scholar] [CrossRef] [PubMed]
Figure 1. TGA (black plot) and DTG (red plot) analysis of as-prepared CoMn2O4.
Figure 1. TGA (black plot) and DTG (red plot) analysis of as-prepared CoMn2O4.
Ijms 26 00940 g001
Figure 2. (a) The XRD pattern of CoMn2O4; (b) FT-IR spectrum of CoMn2O4.
Figure 2. (a) The XRD pattern of CoMn2O4; (b) FT-IR spectrum of CoMn2O4.
Ijms 26 00940 g002
Figure 3. The SEM images of CoMn2O4 samples (ac), the particle size distribution histogram of SEM (d), the TEM images of CoMn2O4 samples (e,g,h), the particle size distribution histogram of TEM (f).
Figure 3. The SEM images of CoMn2O4 samples (ac), the particle size distribution histogram of SEM (d), the TEM images of CoMn2O4 samples (e,g,h), the particle size distribution histogram of TEM (f).
Ijms 26 00940 g003
Figure 4. N2 adsorption–desorption isotherms of CoMn2O4. (insert: pore size distribution).
Figure 4. N2 adsorption–desorption isotherms of CoMn2O4. (insert: pore size distribution).
Ijms 26 00940 g004
Figure 5. XPS spectra of CoMn2O4 samples in different reaction stages: (a) survey scan; (b) O 1s; and (c) Co 2p and (d) Mn 2p.
Figure 5. XPS spectra of CoMn2O4 samples in different reaction stages: (a) survey scan; (b) O 1s; and (c) Co 2p and (d) Mn 2p.
Ijms 26 00940 g005
Figure 6. (a) Effects of CoMn2O4 calcination temperature on the decolorization efficiency of MB and the leaching of metal ions; (b) MB degradation under different systems. Conditions: [MB] = 10 mg/L, [Catalysts] = 1 g/L, effective chlorine 0.05%, pH = 4.3 ± 0.1.
Figure 6. (a) Effects of CoMn2O4 calcination temperature on the decolorization efficiency of MB and the leaching of metal ions; (b) MB degradation under different systems. Conditions: [MB] = 10 mg/L, [Catalysts] = 1 g/L, effective chlorine 0.05%, pH = 4.3 ± 0.1.
Ijms 26 00940 g006
Figure 7. Effects of (a) CoMn2O4 dosage, conditions: [MB] = 10 mg/L, effective chlorine 0.05%, pH = 4.3 ± 0.1; (b) effective chlorine concentration, conditions: [MB] = 10 mg/L, [CoMn2O4] = 1 g/L, pH = 4.3 ± 0.1; (c) MB concentration, conditions: [CoMn2O4] = 1 g/L, pH = 4.3 ± 0.1; (d) initial pH and the leaching of metal ions, conditions: [MB] = 50 mg/L, effective chlorine 0.1%, [CoMn2O4] = 1 g/L.
Figure 7. Effects of (a) CoMn2O4 dosage, conditions: [MB] = 10 mg/L, effective chlorine 0.05%, pH = 4.3 ± 0.1; (b) effective chlorine concentration, conditions: [MB] = 10 mg/L, [CoMn2O4] = 1 g/L, pH = 4.3 ± 0.1; (c) MB concentration, conditions: [CoMn2O4] = 1 g/L, pH = 4.3 ± 0.1; (d) initial pH and the leaching of metal ions, conditions: [MB] = 50 mg/L, effective chlorine 0.1%, [CoMn2O4] = 1 g/L.
Ijms 26 00940 g007
Figure 8. (a) The spectral evolution of MB and (b) oxidative performance of various ubiquitous pollutants by CoMn2O4/NaClO system under the optimal conditions.
Figure 8. (a) The spectral evolution of MB and (b) oxidative performance of various ubiquitous pollutants by CoMn2O4/NaClO system under the optimal conditions.
Ijms 26 00940 g008
Figure 9. (a) The spectral evolution of MB under the system of CoMn2O4/NaClO; (b) degradation of various pollutants by CoMn2O4/NaClO. Conditions: [pollutant] = 50 mg/L, [Catalysts] = 1 g/L, effective chlorine 0.1%, pH = 4.3 ± 0.1.
Figure 9. (a) The spectral evolution of MB under the system of CoMn2O4/NaClO; (b) degradation of various pollutants by CoMn2O4/NaClO. Conditions: [pollutant] = 50 mg/L, [Catalysts] = 1 g/L, effective chlorine 0.1%, pH = 4.3 ± 0.1.
Ijms 26 00940 g009
Figure 10. (a) The mass spectra of MB solution after reaction with CoMn2O4/NaClO; (b) the proposed pathway of MB decolorization.
Figure 10. (a) The mass spectra of MB solution after reaction with CoMn2O4/NaClO; (b) the proposed pathway of MB decolorization.
Ijms 26 00940 g010
Figure 11. (a) Removal of MB using the recycled CoMn2O4. Conditions: [MB] = 50 mg/L, [CoMn2O4] = 1 g/L, effective chlorine 0.1%, pH = 4.3 ± 0.1; (b) N2 adsorption–desorption isotherms of CoMn2O4 after repeat discoloration experiment (insert: pore size distribution).
Figure 11. (a) Removal of MB using the recycled CoMn2O4. Conditions: [MB] = 50 mg/L, [CoMn2O4] = 1 g/L, effective chlorine 0.1%, pH = 4.3 ± 0.1; (b) N2 adsorption–desorption isotherms of CoMn2O4 after repeat discoloration experiment (insert: pore size distribution).
Ijms 26 00940 g011
Table 1. Degradation of MB by different catalysts and NaClO.
Table 1. Degradation of MB by different catalysts and NaClO.
CatalystMBCatalyst DosageEffective ChlorineTimeDegradation RateReference
-1.6 mg/L-0.003%80 min85.2%[66]
MnO2-MCP100.0 mg/L50 g/L0.014%300 min98.5%[13]
CoMn2O410.0 mg/L1 g/L0.050%40 min98.7%This article
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhao, T.; Han, G.; Bai, J.; Wu, X. Heterogeneous Activation of NaClO by Nano-CoMn2O4 Spinel for Methylene Blue Decolorization. Int. J. Mol. Sci. 2025, 26, 940. https://doi.org/10.3390/ijms26030940

AMA Style

Zhao T, Han G, Bai J, Wu X. Heterogeneous Activation of NaClO by Nano-CoMn2O4 Spinel for Methylene Blue Decolorization. International Journal of Molecular Sciences. 2025; 26(3):940. https://doi.org/10.3390/ijms26030940

Chicago/Turabian Style

Zhao, Tongwen, Gang Han, Juan Bai, and Xiaogang Wu. 2025. "Heterogeneous Activation of NaClO by Nano-CoMn2O4 Spinel for Methylene Blue Decolorization" International Journal of Molecular Sciences 26, no. 3: 940. https://doi.org/10.3390/ijms26030940

APA Style

Zhao, T., Han, G., Bai, J., & Wu, X. (2025). Heterogeneous Activation of NaClO by Nano-CoMn2O4 Spinel for Methylene Blue Decolorization. International Journal of Molecular Sciences, 26(3), 940. https://doi.org/10.3390/ijms26030940

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop