Next Article in Journal
Feature Review Papers in Biocatalysis and Enzyme Engineering
Previous Article in Journal
Single-Atom Catalysts for Electrochemical Nitrate Reduction to Ammonia: Rational Design, Mechanistic Insights, and System Perspectives
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ru-Modified α-MnO2 as an Efficient PMS Activator for Carbamazepine Degradation: Performance and Mechanism

1
Sichuan Changning Natural Gas Development Co., Ltd., Chengdu 610051, China
2
Sichuan Jiacheng Petroleum and Natural Gas Pipeline Quality Inspection Testing Co., Ltd., Chengdu 402260, China
3
School of Civil and Hydraulic Engineering, Chongqing University of Science and Technology, Chongqing 401331, China
4
Chongqing Institute of Geology and Mineral Resources, Chongqing 401120, China
5
School of Environment and Resources, Chongqing Technology and Business University, Chongqing 400067, China
*
Authors to whom correspondence should be addressed.
Catalysts 2025, 15(11), 1085; https://doi.org/10.3390/catal15111085
Submission received: 22 October 2025 / Revised: 8 November 2025 / Accepted: 10 November 2025 / Published: 17 November 2025
(This article belongs to the Special Issue Advanced Catalysts for Energy Conversion and Environmental Protection)

Abstract

Although Ru-based catalysts have been investigated in various oxidation systems, their application in sulfate radical-based AOPs, particularly as heterogeneous activators for acidic wastewater treatment, remains limited. Herein, Ru was incorporated into α-MnO2 via lattice doping and surface loading to construct Rulatt/α-MnO2 and Rusurf/α-MnO2, and their PMS activation performance toward carbamazepine (CBZ) degradation was evaluated. Rulatt/α-MnO2 exhibited superior activity, achieving near-complete CBZ removal within minutes under acidic conditions. PMS dosage, catalyst loading, and pH affected the degradation efficiency, with acidic environments significantly enhancing PMS activation. Cl slightly promoted CBZ degradation, whereas HCO3 and natural organic matter inhibited it. Mechanistic analysis revealed that Ru activated PMS through a nonradical pathway, continuously generating 1O2 via a reversible Ru (II)/Ru (III)/Ru (IV) cycle, while the Mn (III)/Mn (IV) redox couple acted as an electron buffer to sustain Ru cycling and improve durability. The catalyst maintained high activity in complex water matrices, demonstrating strong potential for practical remediation of CBZ-contaminated acidic wastewater.

1. Introduction

With the continuous advancement of society and the increasing public awareness of human health, the production and consumption of organic pharmaceuticals have markedly expanded. Nevertheless, their extensive application, together with the insufficient removal efficiency of conventional wastewater treatment technologies, has resulted in the widespread occurrence of these pharmaceuticals in wastewater [1]. Among various organic pharmaceuticals, carbamazepine (CBZ) is recognized as a particularly recalcitrant compound due to its chemically stable molecular structure and extremely low degradation rate in natural attenuation processes such as photolysis, volatilization, and biodegradation [2]. Owing to its high consumption and persistent resistance to degradation, CBZ has been identified as a globally monitored persistent organic pollutant (POP) of great environmental concern [3]. Pharmaceutical production and some hospital/laboratory activities often generate acid-containing process waters (e.g., synthesis, solvent washes, and equipment cleaning); when not fully neutralized or when mixed with other acidic discharges, these streams can produce acidic wastewater that still contains detectable carbamazepine [4]. Furthermore, carbamazepine, excreted by humans, enters domestic and hospital wastewater; when these drug-containing streams mix with inadequately neutralized industrial acidic effluents in the sewer network, they can produce composite wastewater that is both acidic and contaminated with carbamazepine [5,6]. In addition, numerous industrial activities, including pharmaceutical manufacturing, metal processing, and chemical synthesis, often generate acidic wastewater. Conventional treatment technologies for organic contaminants generally exhibit poor efficiency under such acidic conditions. This inefficiency arises from both the suppression of degradation pathways in acidic environments and the inherent chemical stability of CBZ. Without effective treatment strategies, these wastewaters may pose long-term ecological risks, including endocrine disruption, genotoxicity, and potential threats to human health. Therefore, developing efficient and cost-effective approaches for CBZ removal under acidic conditions remains a significant challenge.
Advanced oxidation processes (AOPs) based on peroxymonosulfate (PMS) activation have emerged as one of the most attractive technologies for the removal of organic contaminants, including carbamazepine (CBZ). This is mainly because they feature a short treatment time and high degradation efficiency [7]. Compared with conventional Fenton-based processes, sulfate radical ( SO 4 ) driven advanced oxidation processes (AOPs) exhibit stronger oxidative potential, a longer half-life (30–40 μs), and a broader operational pH range (2–9). Although peroxymonosulfate (PMS) itself can oxidize organic compounds with relatively low reactivity, the reactive oxygen species (ROS) generated during PMS activation—such as singlet oxygen (1O2), sulfate radicals ( SO 4 ), hydroxyl radicals (·OH), and superoxide radicals ( O 2 )—play a more decisive role in pollutant removal [8,9,10]. Various approaches have been developed to activate PMS, including ultraviolet irradiation [11], ultrasound [12], alkali activation [13], and catalyst activation [14,15,16,17]. Among these, activation by metal-based catalysts is particularly attractive due to its high efficiency, low energy consumption, and minimal secondary pollution, showing great potential for water treatment. Among noble metal catalysts, ruthenium (Ru)-based activators have attracted increasing attention owing to their high catalytic activity and excellent redox performance. Compared with other platinum-group metal catalysts, Ru exhibits a relatively lower cost, while offering higher activation efficiency and selectivity than traditional iron-based activators [18,19]. Consequently, Ru catalysts have been widely employed in fields such as hydrogenation catalysis [20] and the degradation of organic pollutants [21]. Moreover, homogeneous Ru species are difficult to recover from aqueous systems, leading to high operational costs and the risk of secondary water contamination, thereby restricting their large-scale application in PMS-based processes. Although a few studies have investigated the performance of Ru-based catalysts for PMS activation, their large-scale application is still limited by issues such as catalyst stability and cost. Therefore, it is necessary to develop Ru-based heterogeneous catalysts that are more stable and cost-effective.
To overcome this challenge, various materials have been employed as supports for ruthenium (Ru), including carbon-based materials [22,23], graphitic carbon nitride (g-C3N4) [24], and metal oxides [25]. Among these, manganese dioxide (MnO2)—the most common manganese oxide compound—occurs widely in natural environments and has attracted significant attention due to its low toxicity and high chemical stability. Its large specific surface area and diverse pore structures enable efficient dispersion and anchoring of active components [26]. Moreover, MnO2, owing to its outstanding electron-shuttling capability, is frequently utilized as a support or co-catalyst for transition metals to facilitate effective electron transfer; such interactions can act synergistically with other supports, thereby enhancing PMS activation and pollutant degradation efficiency [27,28]. MnO2 exists in various crystalline polymorphs, including α-, β-, γ-, and δ-MnO2, which are constructed from different microstructural units [29]. Among them, α-MnO2 has been reported to exhibit superior PMS activation performance compared with β-, γ-, and δ-MnO2, owing to its unique tunnel structure, abundant mixed-valence Mn sites, and high density of oxygen vacancies [30]. Consequently, α-MnO2 has demonstrated excellent performance as a doped or supported material in PMS-based systems. For instance, α-MnO2 nanorods supported on palygorskite (α-MnO2/Pal) achieved nearly 100% degradation of Rhodamine B within 180 min in the PMS system [31]; similarly, a novel MnO2/UiO-66 composite synthesized via an oil-bath heating method efficiently activated PMS under UV irradiation for the degradation of oxytetracycline (OTC) [26]. However, the development of heterogeneous Ru/α-MnO2 composite catalysts and the systematic investigation of their catalytic performance in PMS activation remain scarce in the literature.
Considering the above aspects, this study designed and synthesized a heterogeneous catalyst (Ru/α-MnO2) by loading catalytically active Ru onto α-MnO2. The prepared activator was systematically characterized, and a series of experiments was conducted to evaluate its performance and applicability in the degradation of carbamazepine (CBZ) in aqueous systems. The objectives of this work are as follows: (1) to synthesize and characterize the heterogeneous activator Ru/α-MnO2; (2) to investigate the CBZ degradation performance of PMS activation by Ru/α-MnO2 and evaluate the effects of various operating parameters; (3) to elucidate the activation mechanism of PMS by Ru/α-MnO2 and identify the possible degradation pathways of CBZ; and (4) to assess the reusability and practical applicability of Ru/α-MnO2 in real wastewater treatment.

2. Results and Discussion

2.1. Catalyst Characterization

The surface morphology and structural characteristics of the catalysts play a critical role in their PMS activation efficiency. Field-emission scanning electron microscopy (FE-SEM) was employed to examine the morphology and microstructure of pristine α-MnO2 and the Ru-doped α-MnO2 composites prepared by two different methods. As shown in Figure 1a, α-MnO2 exhibits a uniform nanorod structure; the incorporation of Ru did not alter the overall morphology of α-MnO2. Energy-dispersive X-ray spectroscopy (EDS) mapping was further performed to analyze the elemental distribution of O, Mn, and Ru within selected regions of the composites (Figure S1a,b). The results confirm a homogeneous dispersion of Ru on the surface of both composites, with O, Mn, and Ru elements clearly detected in the mapped areas. The Ru contents in the as-prepared Rulatt/α-MnO2 and Rusurf/α-MnO2 were quantified as 1.43 wt% and 1.02 wt%, respectively. However, due to the difficulty of directly visualizing Ru species on the α-MnO2 support by SEM–EDS, transmission electron microscopy (TEM) was employed for further investigation. As shown in Figure 1b, Rulatt/α-MnO2 retained the distinct nanorod morphology of α-MnO2 without observable surface nanoparticles, consistent with the SEM results. In contrast, the TEM images of Rusurf/α-MnO2 surface revealed numerous nanoparticles on the MnO2 surface. This observation is likely attributed to the hydrolysis of RuCl3 during the impregnation process, followed by decomposition to form Ru-based oxides or hydroxides. Therefore, it is reasonable to infer that the nanoparticles observed on the surface of α-MnO2 in Rusurf/α-MnO2 are composed of Ru oxides and hydroxides [32].
The crystal phase structures of the three catalysts were characterized by X-ray diffraction (XRD). As shown in Figure 1c, the diffraction patterns of all samples exhibit prominent peaks at 12.8°, 18.1°, 28.8°, 37.5°, 41.9°, 49.9°, 60.3°, and 69.7°, corresponding to the (110), (200), (310), (211), (301), (411), (521), and (541) planes of α-MnO2, respectively. These reflections are in good agreement with the standard pattern of α-MnO2 (PDF# 44-0141) [26]. In addition, Rusurf/α-MnO2 displays an additional diffraction peak at 32.9°, which matches the characteristic peak of Mn2O3 (PDF# 41-1442) [31]. This result indicates that α-MnO2 and Rulatt/α-MnO2 retain the pure α-MnO2 phase, whereas the calcination process during the preparation of Rusurf/α-MnO2 partially transformed α-MnO2 into Mn2O3. All three samples show sharp and intense diffraction peaks, suggesting good crystallinity of the synthesized materials. Notably, the diffraction pattern of Rulatt/α-MnO2 is almost indistinguishable from that of pristine α-MnO2, and no characteristic peaks of Ru or other Ru-based phases are detected. The absence of Ru-related reflections may be attributed to the fact that Ru was successfully incorporated into the α-MnO2 lattice [33]. These findings are consistent with the TEM and EDS observations.
The specific surface area (SSA) of catalysts plays a crucial role in the heterogeneous activation of PMS. To evaluate the catalytic activity of α-MnO2 with different exposed crystal facets, it is essential to first measure the SSA of each sample and normalize the catalytic performance accordingly. The Brunauer–Emmett–Teller (BET) surface areas of the three materials were analyzed (Figure 1d), and the SSA values of Rulatt/α-MnO2, Rusurf/α-MnO2, and pristine α-MnO2 were determined to be 68.74, 31.86, and 53.48 m2/g, respectively. The results indicate that Rulatt/α-MnO2 possesses the largest BET surface area, which can expose more active sites during the reaction, thereby increasing the probability of effective collisions between active sites and target molecules and facilitating the catalytic process. Fourier transform infrared (FT-IR) spectroscopy was further employed to investigate the surface functional groups of Rulatt/α-MnO2 and pristine α-MnO2 (Figure 1e). The characteristic peaks at 472 and 522 cm−1 are assigned to the stretching vibrations of Mn–O bonds [34], while the peak at 717 cm−1 corresponds to the vibration of bridging Mn–O–Mn bonds. The band observed at 1625 cm−1 is attributed to the bending vibration of adsorbed H2O molecules (H–O–H), and the broad absorption at 3420 cm−1 is associated with the stretching vibration of surface hydroxyl groups on Rulatt/α-MnO2 [34,35]. Notably, no distinct new functional groups were detected on the surface of Rulatt/α-MnO2, suggesting that Ru species were successfully incorporated into the α-MnO2 lattice rather than forming separate surface-bound species.

2.2. Evaluation of CBZ Degradation Efficiency Across Different Systems

The CBZ removal performance of PMS, α-MnO2/PMS, Rulatt/α-MnO2/PMS, and Rusurf/α-MnO2/PMS systems was compared under identical conditions, with both PMS and activators dosages fixed at 200 mg/L. As shown in Figure 2, the PMS-only system achieved less than 3% CBZ removal, indicating that although PMS possesses strong intrinsic oxidizing capability, it generates few reactive species at ambient temperature and thus has negligible direct oxidation capacity toward CBZ under the tested conditions. In contrast, the introduction of α-MnO2, Rusurf/α-MnO2/PMS, and Rulatt/α-MnO2/PMS markedly enhanced PMS activation and CBZ degradation. Within 5 min, CBZ removal reached 16.74% in the α-MnO2/PMS system, 50% in the Rusurf/α-MnO2/PMS system, and up to 70% in the Rulatt/α-MnO2/PMS system. These results clearly demonstrate that Rulatt/α-MnO2 exhibits superior catalytic performance for PMS activation and CBZ degradation compared with the other systems. This enhanced activity can be attributed to the excellent structural stability of Rulatt/α-MnO2 and the synergistic interaction between Ru species and the α-MnO2 framework, where Ru is incorporated in different forms, thus facilitating more efficient PMS activation. The degradation efficiency, underlying mechanism, and practical applicability of the Rulatt/α-MnO2/PMS system for CBZ removal will be further discussed in the following sections.

2.3. The Influence of Several Reaction Parameters on the Degradation of Carbamazepine

The initial pH of the solution exhibited a pronounced influence on CBZ degradation. As shown in Figure 3, the highest catalytic efficiency was achieved under strongly acidic conditions (pH = 3), where complete CBZ removal was accomplished within 8 min, and the observed rate constant (Kobs) reached 0.33502 min−1. As the pH increased from 5 to 11, the 10 min removal efficiencies decreased to 64.3%, 65.8%, 62%, and 20.4%, respectively. These results indicate that the system maintains high degradation efficiency under acidic to near-neutral conditions, exhibiting a broad operational pH range (pH 3–9). However, when the pH increased from 9 to 11, the CBZ removal efficiency dropped sharply from 62% to 20.4%. Kobs progressively declined with increasing pH, reaching a minimum of 0.0223 min−1 at pH 11—approximately 15-fold lower than that observed at pH 3. This decline can be explained by the acid–base speciation of PMS, a diprotic weak acid whose protonation state varies with pH [36]. Under acidic conditions, PMS predominantly exists as the monovalent anion HSO5, which is more readily activated to generate sulfate radicals ( SO 4 ). SO 4 possesses a longer half-life and a higher redox potential than hydroxyl radicals (·OH) [37], resulting in enhanced CBZ degradation in acidic media. In alkaline conditions, however, SO 4 tends to react with H2O or OH to yield ·OH and SO42−, while deprotonation of HSO5 leads to SO52−, whose oxidative potential is lower than that of SO 4 [38]. Moreover, an excessively high pH promotes ·OH self-quenching reactions, further reducing the concentration of reactive radicals and thus the overall degradation efficiency.
The dosage of the activator directly influences the total number of active sites available in the system. In this study, the effect of Rulatt/MnO2 dosage on CBZ removal was investigated (Figure 4). When the catalyst loading was 100 mg/L, CBZ removal reached 99.91% after 10 min, with an observed rate constant Kobs of 0.2342 min−1. Increasing the dosage to 200 mg/L dramatically accelerated the reaction, achieving complete CBZ removal within 8 min and increasing Kobs to 0.33502 min−1. Further increasing the catalyst loading to 300, 400, and 500 mg/L enabled nearly complete CBZ degradation within only 3 min, with Kobs rising to 0.69548, 0.67915, and 0.74015 min−1, respectively. These results reveal a positive correlation between Rulatt/MnO2 dosage and CBZ degradation efficiency (and reaction kinetics) within a certain range, demonstrating the excellent PMS activation capability of this catalyst. With PMS concentration fixed, increasing the amount of Rulatt/MnO2 introduces more α-MnO2 and Ru active species, thereby providing additional reactive sites that can interact with PMS to generate more reactive oxygen species (ROS), which in turn accelerate CBZ degradation. For consistency and comparability in subsequent experiments, a catalyst dosage of 200 mg/L was selected.
The effect of PMS dosage on CBZ removal was investigated by varying the PMS concentration from 100 to 500 mg/L (Figure 5). As shown in the figure, the reaction rate gradually increased with increasing PMS concentration, achieving complete CBZ removal within 10, 8, 4, and 3 min, respectively. Meanwhile, the apparent rate constant (Kobs) also progressively increased with PMS concentration. Specifically, Kobs rose from 0.20686 min−1 at 100 mg/L PMS to 0.80676 min−1 at 500 mg/L PMS—approximately 3.9 times higher than that observed at 100 mg/L. This trend indicates that a higher PMS dosage enhances the contact between PMS molecules and the active sites of Rulatt/α-MnO2, thereby accelerating the degradation of CBZ. When PMS is supplied in sufficient amounts, it can rapidly occupy the available active sites on Rulatt/α-MnO2, generating more reactive species to effectively oxidize CBZ. Considering both cost efficiency and operational practicality, a PMS concentration of 200 mg/L was selected for subsequent experiments.

2.4. Catalytic Mechanism of Rulatt/α-MnO2

Identification of Reactive Species

It has been reported that both radical species and singlet oxygen (1O2) are commonly generated in PMS activation systems, contributing to the degradation of organic pollutants in Ru/MnOx/PMS systems [15,27]. Therefore, quenching experiments were conducted to identify the dominant reactive species involved in the Rulatt/MnO2/PMS system (Figure 6a). tert-Butyl alcohol (TBA) was employed as a scavenger for hydroxyl radicals (·OH), while methanol (MeOH) was used to quench both sulfate radicals ( SO 4 ) and ·OH. Furfuryl alcohol (FFA) was applied as a selective scavenger for singlet oxygen (1O2) to evaluate its contribution [39], and p-benzoquinone (p-BQ) was used to capture superoxide radicals ( O 2 ) [40]. As shown in Figure 6a, the addition of p-BQ almost completely suppressed CBZ degradation since the apparent rate constant (Kobs) sharply decreased, dropping to a value 26-fold lower than that of the blank group. This result indicates a critical role of O 2 in the reaction system. In addition to radical pathways, nonradical processes were also investigated. The introduction of FFA exhibited a nearly complete inhibition effect, with CBZ removal dropping to only 0.12% within 3 min; the apparent rate constant (Kobs)decreased to 46.5 times lower than that of the blank group. Suggesting that CBZ degradation was almost entirely suppressed. These findings indicate that 1O2 is the predominant reactive oxygen species responsible for CBZ degradation in the Rulatt-MnO2/PMS system, while O 2 also contributes significantly to the overall oxidation process.
Interestingly, the addition of MeOH or TBA did not lead to the expected decrease in CBZ removal; the apparent rate constant (Kobs) increased by 2.6-fold with MeOH and 1.2-fold with TBA, indicating a significant enhancement in the reaction kinetics. In other words, their presence promoted CBZ degradation. According to previous studies, this phenomenon may be attributed to the inherently short lifetime and low reactivity of superoxide radicals ( O 2 ) in pure water. In the presence of less polar protic solvents such as methanol, the solvation environment of O 2 is altered, resulting in significantly enhanced stability and reactivity [41]. In addition, TBA can react with O 2 to generate peroxy radicals (ROO·), which are capable of directly oxidizing Mn2+ in aqueous solution to form manganese oxides (e.g., MnO2). In the presence of TBA, the average oxidation state of manganese species shifts from trivalent to more tetravalent forms, potentially enhancing PMS activation. Taken together, the quenching experiments suggest that the Rulatt/MnO2/PMS system predominantly follows a nonradical pathway for CBZ degradation. Based on the above discussion, it can be inferred that 1O2 and O 2 are the primary reactive species responsible for PMS activation in the Rulatt/MnO2/PMS system. Previous reports have suggested that O 2 often serves as a precursor or auxiliary oxidant for 1O2 formation [42,43]; specifically, O 2 and SO5 can be consumed by water molecules to generate 1O2 [44]. The introduction of a low concentration of methanol improves the stability and reactivity of O 2 , thereby promoting 1O2 generation and accelerating CBZ degradation.
Electron paramagnetic resonance (EPR) spectroscopy was further employed to investigate the generation of reactive oxygen species (ROS) during PMS activation by Rulatt/MnO2/PMS (Figure 7). As shown in Figure 7a, the EPR spectra exhibited a characteristic 1:1:1 triplet signal of TEMP–1O2 adducts, confirming the production of singlet oxygen (1O2) [10], with the signal intensity increasing progressively over time. In addition, distinct signals assigned to the DMPO– O 2 adduct were observed [45], and their intensity also increased with reaction time, further verifying the presence of O 2 in the system. In contrast, no characteristic peaks corresponding to DMPO– SO 4 or DMPO–·OH were detected, indicating that sulfate radicals ( SO 4 ) and hydroxyl radicals (·OH) were not generated in significant amounts. These findings demonstrate that CBZ degradation in the Rulatt/MnO2/PMS system is primarily governed by a nonradical pathway, dominated by 1O2, with O 2 either serving as a precursor for 1O2 formation or participating as an auxiliary oxidant. The EPR results, combined with the quenching experiments, provide strong evidence that the PMS activation and CBZ degradation process is mainly driven by 1O2 and O 2 .
Manganese oxides are generally considered capable of reacting with PMS to generate sulfate radicals ( SO 4 ) and hydroxyl radicals (·OH), as described in previous studies [46,47]. However, the quenching experiments and EPR results in this work revealed that neither SO 4 nor ·OH was detected in the Rulatt/α-MnO2/PMS system. Therefore, it is necessary to further elucidate the actual role of manganese oxides in this catalytic process. To achieve this, phosphate (H2PO4) was introduced as a blocking agent for the Mn active sites. Previous studies have shown that H2PO4 can form stable complexes with Mn (III) [48], allowing us to distinguish whether Mn (III) participates in an alternative degradation pathway rather than directly activating PMS to generate radicals. Using this strategy, we aimed to clarify whether manganese oxides contribute through a nonradical pathway and to provide stronger evidence regarding the mechanistic role of Mn in PMS activation. As shown in the results (Figure 8), the addition of H2PO4 significantly suppressed the CBZ removal efficiency, with only 4% degradation observed after 10 min. This strong inhibition suggests that Mn (III) plays a crucial role in the degradation process of the Rulatt/α-MnO2/PMS system, rather than simply catalyzing PMS to produce free radicals. To further clarify the involvement of Mn (III) and its valence-state changes during the reaction, X-ray photoelectron spectroscopy (XPS) analysis was performed.
To further elucidate the activation mechanism of the Rulatt/α-MnO2/PMS system, X-ray photoelectron spectroscopy (XPS) was employed to analyze the valence states of the metal elements. The oxygen species of pristine α-MnO2 and Rulatt/α-MnO2/PMS were examined, as shown in Figure S2. The O 1s spectra of both materials could be deconvoluted into three peaks located at 533.3, 531.8, and 530.4 eV, corresponding to chemisorbed water or hydroxyl oxygen (Olatt), surface-adsorbed oxygen species near oxygen vacancies (Oads), and lattice oxygen (Osurf), respectively [49]. Generally, the content of Oads can reflect the abundance of oxygen vacancies (Ov), as Oads is typically formed by oxygen molecules adsorbing at Ov sites [41]. It can be observed that the relative proportion of Oads is similar between α-MnO2 and Rulatt/α-MnO2/PMS, suggesting that the introduction of Ru did not significantly alter the structural framework of α-MnO2 and that both materials possess abundant surface-active sites capable of participating in the reaction. However, despite their comparable Oads content, the two catalysts show markedly different PMS activation efficiencies, implying that the oxygen vacancies exposed on the crystal facets are not the dominant factor governing catalytic activity in PMS activation. Previous studies have reported that Mn (II) and Mn (III) species serve as key active sites, with Mn (III) playing a particularly critical role in PMS activation [50]. As shown in the Mn 2p3/2 spectra (Figure S2), the peaks at 640.7, 642.2, and 643.2 eV correspond to the binding energies of Mn (II), Mn (III), and Mn (IV), respectively [50]. Quantitative analysis revealed that the proportion of Mn (III) in Rulatt/α-MnO2/PMS is slightly higher than that in pristine α-MnO2, indicating that the incorporation of Ru increased the Mn (III) content in the α-MnO2 framework, thereby enhancing its intrinsic PMS activation capability.
X-ray photoelectron spectroscopy (XPS) was further conducted to examine the valence-state changes in O, Mn, and Ru in Rulatt/α-MnO2/PMS before and after reaction in order to elucidate the fundamental activation mechanism. As shown in Figure 9a, the proportion of surface-adsorbed oxygen species (Oads) markedly decreased from 34.3% to 19.0%, accompanied by a relative increase in chemisorbed water/hydroxyl oxygen (Olatt) from 55.0% to 72.8%. This observation is consistent with previous reports, where Oads is typically consumed or transformed during PMS activation dominated by nonradical pathways (e.g., 1O2 generation), leading to a decrease in Oads content and a relative enrichment of Olatt [51]. Therefore, these results indirectly indicate that singlet oxygen (1O2) generation plays a predominant role in PMS activation within this system, which is in good agreement with the quenching and EPR analyses.
The Mn 2p XPS spectra revealed a pronounced decrease in the proportion of Mn (III) from 58.0% to 37.5% after reaction, accompanied by a corresponding increase in Mn (IV). Although previous studies have widely reported that Mn (III) plays a key role in PMS activation, the combined results of the present work indicate that the valence transition of Mn is not associated with radical generation from PMS. Together with the phosphate blocking experiments, it can be inferred that Mn participates in the reaction via an alternative pathway. Similar phenomena have been reported in earlier studies, in which Mn acted as an electron reservoir to facilitate the redox cycling of Ru, while Ru sites dominated PMS activation and drove contaminant degradation through a nonradical pathway. As shown in Figure 9c,d, the evolution of Ru oxidation states further supports this hypothesis. The Ru 3p spectra show a relative decrease in Ru (II) and a corresponding increase in Ru (III) after reaction. Meanwhile, in the Ru 3d spectra, the fraction of Ru (III) slightly decreased relative to Ru (IV), with a modest enrichment of Ru (IV) [52]. Overall, Ru tends to be oxidized to higher valence states under PMS exposure, while a reversible redox transition between Ru (III) and Ru (IV) occurs during the catalytic process. Based on these results and the previous evidence, a plausible degradation mechanism for the Rulatt/α-MnO2/PMS system can be proposed: (1) PMS is activated at the Ru sites, where low-valent Ru is oxidized to higher oxidation states, concurrently generating singlet oxygen (1O2) that initiates CBZ degradation [53]; (2) Mn(III) in α-MnO2 is converted to Mn(IV), releasing electrons to reduce the high-valent Ru species and thus sustain the reversible Ru redox cycle [53]; (3) during this process, Mn(IV) does not generate SO 4 or ·OH upon interaction with PMS but is instead reduced back to Mn(III), establishing a Mn(III)/Mn(IV) redox loop [54]. Collectively, these findings suggest that Ru primarily governs PMS activation through a nonradical pathway, while α-MnO2 functions as an electron reservoir that supports the reversible Ru valence cycling and enhances the overall catalytic stability.
In summary, the degradation of carbamazepine (CBZ) in this catalytic system is predominantly governed by a singlet oxygen (1O2)-driven pathway rather than the widely reported sulfate radical ( SO 4 ) or hydroxyl radical (·OH) routes. This conclusion is strongly supported by radical quenching experiments and EPR analysis. In particular, EPR spectra clearly revealed the presence of O 2 and 1O2 signals but showed no detectable SO 4 or ·OH, confirming that nonradical pathways dominate the PMS activation process [8]. It is noteworthy that O 2 is commonly considered a precursor or auxiliary oxidant for 1O2 formation; it can be generated through surface electron-transfer reactions or via SO 5 2 - decomposition, subsequently promoting the continuous production of 1O2 [55], thereby synergistically contributing to CBZ oxidation. Furthermore, combined evidence from quenching tests, EPR detection of reactive oxygen species, Mn (III) blocking experiments, and XPS analysis elucidated the evolution of surface valence states and oxygen species during PMS activation. Mn (III) content decreased significantly with a concomitant enrichment of Mn (IV), indicating that Mn acts as an electron reservoir undergoing progressive oxidation. Meanwhile, Ru (II) was partially consumed, and the fraction of Ru (III)/Ru (IV) slightly increased, reflecting a reversible Ru (II) → Ru (III) → Ru (IV) redox cycle. In this cycle, Ru sites serve as redox mediators to activate PMS [25], while interfacial electron transfer from Mn to Ru sustains the Ru turnover and enhances its catalytic stability [56]. Simultaneously, the marked decrease in surface-adsorbed oxygen species (Oads) and the relative enrichment of chemisorbed water/hydroxyl oxygen (Olatt) indicate that Oads is consumed or transformed during the generation of O 2 and 1O2, indirectly validating the nonradical pathway. Collectively, the synergistic coupling of the Mn (IV)/Mn (III) redox cycle with the reversible Ru (II)/Ru (III)/Ru (IV) transition facilitates efficient PMS activation and CBZ degradation. The overall process is dominated by 1O2, with O 2 acting as a precursor and auxiliary oxidant, while Ru–Mn interfacial electron transfer and the dynamic reconstruction of surface oxygen species further enhance catalytic performance (Equations (1)–(13)) [25,57,58,59,60,61].
HSO5 + SO52− → SO42− + HSO4 + 1O2
Ruα+ + HSO5 → Ru(α + 1)+ + SO52− + H+
HSO5 + H2O → HSO4 + H2O2
Ru(α + 1)+ + H2O2 → Ruα+ + HO2 + H+
HO 2     O 2 + H +
O 2 + HO 2 HO 2 + O 2 1
2 O 2 + 2 H + H 2 O 2 + O 2 1
Mn (III) → Mn (IV) + e
Mn (IV) + HSO5 → Mn (III) + SO5 + H+
Ru(α + 1)+ + e → Ruα+
2 O 2 + 2 H 2 O H 2 O 2 + 2 OH + O 2 1
4SO5 + 2H2O → 4HSO4 + 31O2
O 2 1 / O 2 + C B Z I n t e r m e d i a t e s + C O 2 + H 2 O

2.5. The Influence of Coexisting Anions and Natural Organic Matter

In real aquatic environments, the presence of coexisting anions and natural organic matter (NOM) can interfere with the degradation of target pollutants. Therefore, the effects of common coexisting anions (NO3, Cl, and HCO3) as well as humic acid (HA) on CBZ removal were investigated. As shown in Figure 10, the presence of Cl remarkably accelerated CBZ degradation, achieving nearly 100% removal within 5 min. Similar promoting effects of chloride have also been observed in other studies [62,63], which can be attributed to the fact that Cl, as an electron-rich anion, can act as an electron donor during PMS activation, thereby enhancing CBZ oxidation. The NO3 containing system also showed a slight improvement in CBZ removal compared with the control, reaching 99.95% degradation within 5 min. Compared with the control, the presence of HCO3 exhibited a modest promoting effect on CBZ degradation. This could be associated with its ability to modulate deprotonation of functional groups and facilitate PMS decomposition toward nonradical pathways; under certain conditions, HCO3 may also generate carbonate radicals ( CO 3 ) that contribute to pollutant oxidation [64]. Considering that wastewater typically contains a substantial amount of organic matter, humic acid (HA) was employed as a representative model to evaluate the effect of NOM on CBZ degradation in the Rulatt/α-MnO2/PMS system. The addition of HA significantly inhibited CBZ degradation, reducing the removal efficiency to only 63% after 10 min, likely due to the competitive consumption of reactive oxygen species (ROS) by both HA and CBZ [47].

3. Materials and Methods

3.1. Chemicals

All reagents were of analytical grade and used without further purification. Potassium permanganate (KMnO4), manganese(II) sulfate monohydrate (MnSO4·H2O), ruthenium(III) chloride hydrate (RuCl3·xH2O), and potassium peroxymonosulfate (2KHSO5·KHSO4·K2SO4, 47%) were purchased from Macklin Biochemical Co., Ltd. (Shanghai, China). Carbamazepine (CBZ) was obtained from Macklin and stored at 4 °C before use. tert-Butyl alcohol (TBA), p-benzoquinone (p-BQ), methanol (MeOH), and furfuryl alcohol (FFA) were also supplied by Macklin. All solutions were prepared using high-purity deionized (DI) water.

3.2. Preparation of Materials

A total of 1.26427 g potassium permanganate (KMnO4), 1.35208 g manganese(II) sulfate monohydrate (MnSO4·H2O), and 0.136 g ruthenium(III) chloride hydrate (RuCl3·xH2O) were dissolved in 50 mL deionized (DI) water and stirred for 30 min to obtain a deep purple–black precursor solution. The mixture was then transferred to a Teflon-lined stainless steel autoclave and subjected to hydrothermal treatment at 140 °C for 6 h. After cooling to room temperature, the resulting solid was washed three times with ultrapure water, dried at 100 °C, and ground into a fine powder to obtain the catalyst denoted as Rulatt/α-MnO2. For comparison, α-MnO2 was prepared following the same procedure but without the addition of RuCl3·xH2O; specifically, 1.26427 g KMnO4 and 1.35208 g MnSO4·H2O were dissolved in 50 mL DI water, stirred for 30 min, and then treated under identical hydrothermal, washing, drying, and grinding conditions. The surface-loaded Rusurf/α-MnO2 catalyst was synthesized by an impregnation method: a RuCl3·xH2O aqueous solution was mixed with the as-prepared α-MnO2, followed by solvent evaporation at 100 °C, washing three times with DI water, drying at 100 °C, and calcination at 500 °C for 2 h with a heating rate of 2 °C min−1. Finally, the product was naturally cooled to room temperature to obtain the desired Rusurf/α-MnO2 catalyst.

3.3. Sample Characterizations

The surface morphologies of Rulatt/α-MnO2 and Rusurf/α-MnO2 were examined using scanning electron microscopy (SEM, ZEISS GeminiSEM 300, Freiburg, Germany) operated at an accelerating voltage of 30 kV. The nanostructures were further observed by transmission electron microscopy (TEM, Talos F200 S, Thermo Fisher Scientific, Waltham, MA, USA) at 200 kV. The internal structure, phase composition, and crystallinity of the materials were analyzed by X-ray diffraction (XRD, PANalytical X’Pert Powder, Spectris, Almelo, The Netherlands) with Cu Kα radiation over a 2θ range of 5–90° at a scanning rate of 5° min−1. The Brunauer–Emmett–Teller (BET) specific surface areas were determined using a surface area analyzer (ASAP 2460, Micromeritics, Norcross, GA, USA). The surface elemental compositions and oxidation states of Mn and Ru were characterized by X-ray photoelectron spectroscopy (XPS, K-Alpha, Thermo Scientific, Waltham, MA, USA), and the spectra were deconvoluted using XPSPEAK41 software. Surface functional groups were identified by Fourier transform infrared spectroscopy (FT-IR, Nicolet iS20, Thermo Fisher Scientific, Waltham, MA, USA). Electron paramagnetic resonance (EPR, EMXplus-6/1, Bruker, Karlsruhe, Germany) was employed to detect reactive oxygen species (ROS), with 5,5-dimethyl-1-pyrroline-N-oxide (DMPO) and 2,2,6,6-tetramethyl-4-piperidinol (TEMP) used as spin-trapping agents.

3.4. Test Apparatus and Analytical Methods

The catalytic degradation performance of Rulatt/α-MnO2 and Rusurf/α-MnO2 toward CBZ was evaluated in a 100 mL beaker under continuous stirring at 800 rpm at room temperature. Briefly, a predetermined amount of Rulatt/α-MnO2 was ultrasonically dispersed in 50 mL deionized (DI) water for 30 s, followed by the addition of 50 mL CBZ stock solution (10 mg L−1). The pH of the mixed solution was adjusted to 7.0 using 0.1M H2SO4 and 0.1M NaOH as needed. Subsequently, 50 mg of peroxymonosulfate (PMS) was added to initiate the degradation reaction. At predetermined time intervals, 2.4 mL of the reaction solution was withdrawn and immediately quenched with 0.6 mL anhydrous methanol to stop further reaction. The quenched samples were filtered through a 0.22 μm PTFE membrane, and 1 mL of the filtrate was transferred to an HPLC vial for organic contaminant analysis using high-performance liquid chromatography (HPLC, Agilent 1260, Waltham, MA, USA). Degradation experiments using Rusurf/α-MnO2 were performed under identical conditions. The apparent reaction rate constants (k) were determined using a pseudo-first-order kinetic model.
ln(C0/C) = kt
The reaction rate constant is determined by a first-order kinetic model, where C0, C, k, and t denote the initial concentration (mg·L−1), concentration at time t (mg·L−1), reaction rate constant (min−1), and reaction time (min), respectively.

4. Conclusions

In this study, a simple strategy was employed to prepare heterogeneous Ru-based catalysts (Rulatt/α-MnO2 and Rulatt/α-MnO2) by introducing Ru into the α-MnO2 through two different incorporation methods, and their ability to activate PMS for the degradation of CBZ was systematically investigated. Structural characterization revealed that both Rulatt/α-MnO2 and Rusurf/α-MnO2 possessed larger specific surface areas and superior structural stability compared with pristine α-MnO2. Catalytic degradation experiments demonstrated that Rulatt/α-MnO2 exhibited significantly enhanced PMS activation and CBZ removal, particularly under acidic conditions, outperforming both Rusurf/α-MnO2 and α-MnO2. The resulting Ru-doped α-MnO2 overcomes the recovery and leaching problems of homogeneous Ru catalysts, enhances the catalytic performance compared with pristine α-MnO2, and fills a research gap by providing a routinely synthesizable Ru/α-MnO2 heterogeneous material and a systematic evaluation of its performance for PMS activation. Operational parameters, including catalyst dosage, PMS concentration, and solution pH, were found to influence CBZ degradation efficiency. Mechanistic investigations revealed that Ru sites dominate the nonradical PMS activation pathway. As shown in Figure 11, where Ru undergoes reversible valence transitions and generates singlet oxygen (1O2) as the primary reactive species responsible for CBZ oxidation. Meanwhile, the Mn (III)/Mn (IV) redox cycle within α-MnO2 functions as an electron reservoir, sustaining Ru valence cycling and enhancing the catalyst’s stability and long-term reactivity. Multiple characterizations confirmed the structural robustness of Rulatt/α-MnO2 during PMS activation. The Rulatt/α-MnO2 maintained excellent performance in batch experiments under acidic conditions and in the presence of interfering anions, indicating promising potential for practical application in the remediation of CBZ-contaminated acidic water environments.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal15111085/s1, Figure S1: EDS mapping of (a) Rulatt/α-MnO2, (b) Rusurf/α-MnO2, and (c) α-MnO2; Figure S2: XPS of (a) O 1s and (b) Mn 2p.

Author Contributions

P.H.: Formal analysis, Data curation, Methodology, Writing—review and editing. L.Q.: Investigation, Formal analysis, Data curation. M.F.: Supervision, Investigation, Writing—review and editing. Y.C.: Investigation, Formal analysis, Data curation. P.T.: Methodology, Formal analysis, Investigation. B.X.: Investigation, Data curation. W.S.: Methodology, Investigation. Q.W.: Writing—original draft, Resources, Investigation, Funding acquisition, Project administration. J.Z.: Methodology, Investigation. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by National Natural Science Foundation of China [grant number 52300032], Natural Science Foundation of Chongqing [grant number CSTB2023NSCQ-MSX0789], Chongqing Municipal Education Commission [grant number KJQN202201548, KJQN202401554].

Data Availability Statement

The original contributions presented in this study are included in the article/Supplementary Material. Further inquiries can be directed to the corresponding author(s).

Conflicts of Interest

Authors Panfeng Hu, Long Qin, Yuanling Cheng, Pan Tang, Beibei Xin, and Wei Song were employed by Sichuan Changning Natural Gas Development Co., Ltd., Author Manman Feng was employed by Sichuan Jiacheng Petroleum and Natural Gas Pipeline Quality Inspection Testing Co., Ltd., Quanfeng Wang and Jujiao Zhao declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Kujawska, A.; Kielkowska, U.; Atisha, A.; Yanful, E.; Kujawski, W. Comparative analysis of separation methods used for the elimination of pharmaceuticals and personal care products (PPCPs) from water-A critical review. Sep. Purif. Technol. 2022, 290, 36. [Google Scholar] [CrossRef]
  2. Wu, M.H.; Xiang, J.J.; Que, C.J.; Chen, F.F.; Xu, G. Occurrence and fate of psychiatric pharmaceuticals in the urban water system of Shanghai, China. Chemosphere 2015, 138, 486–493. [Google Scholar] [CrossRef]
  3. Gabet, A.; Guy, C.; Fazli, A.; Métivier, H.; de Brauer, C.; Brigante, M.; Mailhot, G. The ability of recycled magnetite nanoparticles to degrade carbamazepine in water through photo-Fenton oxidation at neutral pH. Sep. Purif. Technol. 2023, 317, 123877. [Google Scholar] [CrossRef]
  4. Alkahtani, M.Q.; Morabet, R.E.; Khan, R.A.; Khan, A.R. Pharmaceuticals removal from hospital wastewater by fluid-ized aerobic bioreactor in combination with tubesettler. Sci. Rep. 2024, 14, 24052. [Google Scholar] [CrossRef] [PubMed]
  5. Trognon, J.; Albasi, C.; Choubert, J.M. A critical review on the pathways of carbamazepine transformation products in oxidative wastewater treatment processes. J. Sci. Total Environ. 2024, 912, 169040. [Google Scholar] [CrossRef] [PubMed]
  6. Feijoo, S.; Kamali, M.; Dewil, R. A review of wastewater treatment technologies for the degradation of pharma-ceutically active compounds: Carbamazepine as a case study. J. Chem. Eng. J. 2023, 455, 140589. [Google Scholar] [CrossRef]
  7. Peng, Y.T.; Tang, H.M.; Yao, B.; Gao, X.; Yang, X.; Zhou, Y.Y. Activation of peroxymonosulfate (PMS) by spinel ferrite and their composites in degradation of organic pollutants: A Review. Chem. Eng. J. 2021, 414, 128800. [Google Scholar] [CrossRef]
  8. Dai, C.; You, X.; Liu, Q.; Han, Y.; Duan, Y.; Hu, J.; Li, J.; Li, Z.; Zhou, L.; Zhang, Y.; et al. Peroxymonosulfate activation by Ru/CeO2 for degradation of Triclosan: Efficacy, mechanisms and applicability in groundwater. Chem. Eng. J. 2023, 463, 142479. [Google Scholar] [CrossRef]
  9. Pan, Y.J.; Meng, F.Y.; Bai, J.X.; Song, B.; Cao, Q. Highly efficient peroxymonosulfate activation by CoFe2O4@attapulgite-biochar composites: Degradation properties and mechanism insights. J. Environ. Chem. Eng. 2024, 12, 112579. [Google Scholar] [CrossRef]
  10. Gao, P.; Fan, X.; Su, Y.; Wei, H.; Yang, C.; Wang, Q. Efficient peroxymonosulfate activation for sulfamethoxazole degradation by CoFe2O4 decorated mesoporous silica through a simple one-step grinding-calcination method. J. Environ. Chem. Eng. 2025, 13, 116890. [Google Scholar] [CrossRef]
  11. Yang, J.L.; Zhu, M.S.; Dionysiou, D.D. What is the role of light in persulfate-based advanced oxidation for water treatment? Water Res. 2021, 189, 116627. [Google Scholar] [CrossRef]
  12. Zeng, L.; Wang, C.H.; Hu, J. Ultrasonic degradation of tetracycline combining peroxymonosulfate and BiVO4 microspheres. J. Water Process. Eng. 2023, 56, 104428. [Google Scholar] [CrossRef]
  13. Hu, J.; Zeng, X.K.; Yin, Y.C.; Liu, Y.; Li, Y.; Hu, X.Y.; Zhang, L.; Zhang, X.W. Accelerated alkaline activation of peroxydisulfate by reduced rubidium tungstate nanorods for enhanced degradation of bisphenol A. Environ. Sci.-Nano 2020, 7, 3547–3556. [Google Scholar] [CrossRef]
  14. Zhao, Z.F.; Lin, L.; Liu, S.S.; Chen, Y.Q.; Daniels, S.V.; Xu, Z.J.; Chen, Z.H.; Li, H.T.; Wu, Y.Q.; Guo, L.L.; et al. Unveiling the versatile performance of transition metal sulfides in peroxymonosulfate activation. Chem. Eng. J. 2024, 497, 154682. [Google Scholar] [CrossRef]
  15. Tian, H.R.; Cui, K.P.; Chen, X.; Liu, J.; Zhang, Q. Size-matched hierarchical porous carbon materials anchoring single-atom Fe-N4 sites for PMS activation: An in-depth study of key active species and catalytic mechanisms. J. Hazard. Mater. 2024, 461, 132647. [Google Scholar] [CrossRef]
  16. Dai, C.M.; Huang, X.Y.; Liu, Q.; You, X.J.; Duan, Y.P.; Li, J.X.; Hu, J.J.; Zhang, Y.L.; Liu, S.G.; Fu, R.B. Peroxymonosulfate activation by ruthenium in homogeneous systems for degradation of triclosan: Comparison between Ru(II) and Ru(III). Sep. Purif. Technol. 2024, 332, 125820. [Google Scholar] [CrossRef]
  17. Chen, M.T.; Zhu, L.H.; Liu, S.G.; Li, R.; Wang, N.; Tang, H.Q. Efficient degradation of organic pollutants by low-level Co2+ catalyzed homogeneous activation of peroxymonosulfate. J. Hazard. Mater. 2019, 371, 456–462. [Google Scholar] [CrossRef]
  18. Li, Q.; Zheng, S.S.; Xu, Y.X.; Xue, H.G.; Pang, H. Ruthenium based materials as electrode materials for supercapacitors. Chem. Eng. J. 2018, 333, 505–518. [Google Scholar] [CrossRef]
  19. Zell, T.; Langer, R. From Ruthenium to Iron and ManganeseA Mechanistic View on Challenges and Design Principles of Base-Metal Hydrogenation Catalysts. ChemCatChem 2018, 10, 1930–1940. [Google Scholar] [CrossRef]
  20. Bae, S.Y.; Mahmood, J.; Jeon, I.Y.; Baek, J.B. Recent advances in ruthenium-based electrocatalysts for the hydrogen evolution reaction. Nanoscale Horiz. 2020, 5, 43–56. [Google Scholar] [CrossRef]
  21. Sushma; Kumari, M.; Saroha, A.K. Performance of various catalysts on treatment of refractory pollutants in industrial wastewater by catalytic wet air oxidation: A review. J. Environ. Manag. 2018, 228, 169–188. [Google Scholar] [CrossRef]
  22. Yan, Y.; Yang, Q.; Shang, Q.; Ai, J.; Yang, X.; Wang, D.; Liao, G. Ru doped graphitic carbon nitride mediated peroxymonosulfate activation for diclofenac degradation via singlet oxygen. Chem. Eng. J. 2022, 430, 133174. [Google Scholar] [CrossRef]
  23. Pereira Lopes, R.; Astruc, D. Biochar as a support for nanocatalysts and other reagents: Recent advances and applications. Coord. Chem. Rev. 2021, 426, 213585. [Google Scholar] [CrossRef]
  24. Yin, Y.; Liu, M.; Shi, L.; Zhang, S.; Hirani, R.A.K.; Zhu, C.; Chen, C.; Yuan, A.; Duan, X.; Wang, S.; et al. Highly dispersive Ru confined in porous ultrathin g-C3N4 nanosheets as an efficient peroxymonosulfate activator for removal of organic pollutants. J. Hazard. Mater. 2022, 435, 128939. [Google Scholar] [CrossRef]
  25. Lin, B.Y.; Wu, Y.Y.; Fang, B.Y.; Li, C.Y.; Ni, J.; Wang, X.Y.; Lin, J.X.; Jiang, L.L. Ru surface density effect on ammonia synthesis activity and hydrogen poisoning of ceria-supported Ru catalysts. Chin. J. Catal. 2021, 42, 1712–1723. [Google Scholar] [CrossRef]
  26. Luo, X.; Liang, H.; Qu, F.; Ding, A.; Cheng, X.; Tang, C.Y.; Li, G. Free-standing hierarchical alpha-MnO2@CuO membrane for catalytic filtration degradation of organic pollutants. Chemosphere 2018, 200, 237–247. [Google Scholar] [CrossRef] [PubMed]
  27. Wei, H.X.; Zhao, J.J.; Rahaman, M.H.; Zhu, M.; Zhai, J. Enhanced peroxymonosulfate activation for carbamazepine degradation under strongly alkaline conditions using Cu-doped Mn3O4 catalyst: Characterization, catalytic performance, and mechanism insights. J. Clean. Prod. 2023, 429, 139600. [Google Scholar] [CrossRef]
  28. He, C.; Liao, Y.H.; Chen, C.; Xia, D.H.; Wang, Y.Y.; Tian, S.H.; Yang, J.L.; Shu, D. Realizing a redox-robust Ag/MnO2 catalyst for efficient wet catalytic ozonation of S-VOCs: Promotional role of Ag(0)/Ag(I)-Mn based redox shuttle. Appl. Catal. B-Environ. Energy 2022, 303, 120881. [Google Scholar] [CrossRef]
  29. Boyom-Tatchemo, F.W.; Devred, F.; Ndiffo-Yemeli, G.; Laminsi, S.; Gaigneaux, E.M. Plasma-induced redox reactions synthesis of nanosized α-, γ- and δ-MnO2 catalysts for dye degradation. Appl. Catal. B-Environ. 2020, 260, 118159. [Google Scholar] [CrossRef]
  30. Shen, S.; Zhou, X.; Zhao, Q.; Jiang, W.; Wang, J.; He, L.; Ma, Y.; Yang, L.; Chen, Z. Understanding the nonradical activation of peroxymonosulfate by different crystallographic MnO2: The pivotal role of Mn(III) content on the surface. J. Hazard. Mater. 2022, 439, 129613. [Google Scholar] [CrossRef]
  31. Huang, C.; Wang, Y.; Gong, M.; Wang, W.; Mu, Y.; Hu, Z.-H. α-MnO2/Palygorskite composite as an effective catalyst for heterogeneous activation of peroxymonosulfate (PMS) for the degradation of Rhodamine B. Sep. Purif. Technol. 2020, 230, 115877. [Google Scholar] [CrossRef]
  32. Zhang, J.; Sun, B.; Guan, X.H.; Wang, H.; Bao, H.L.; Huang, Y.Y.; Qiao, J.L.; Zhou, G.M. Ruthenium Nanoparticles Supported on CeO2 for Catalytic Permanganate Oxidation of Butylparaben. Environ. Sci. Technol. 2013, 47, 13011–13019. [Google Scholar] [CrossRef]
  33. Li, J.H.; Liu, Z.Q.; Cullen, D.A.; Hu, W.H.; Huang, J.E.; Yao, L.B.; Peng, Z.M.; Liao, P.L.; Wang, R.G. Distribution and Valence State of Ru Species on CeO2 Supports: Support Shape Effect and Its Influence on CO Oxidation. ACS Catal. 2019, 9, 11088–11103. [Google Scholar] [CrossRef]
  34. Maiti, S.; Pramanik, A.; Mahanty, S. Interconnected Network of MnO2 Nanowires with a “Cocoonlike” Morphology: Redox Couple-Mediated Performance Enhancement in Symmetric Aqueous Supercapacitor. ACS Appl. Mater. Interfaces 2014, 6, 10754–10762. [Google Scholar] [CrossRef]
  35. Ma, M.D.; Zhu, Q.; Jiang, Z.Y.; Jian, Y.F.; Chen, C.W.; Liu, Q.Y.; He, C. Achieving toluene efficient mineralization over K/α-MnO2 via oxygen vacancy modulation. J. Colloid. Interface Sci. 2021, 598, 238–249. [Google Scholar] [CrossRef] [PubMed]
  36. Ghanbari, F.; Moradi, M. Application of peroxymonosulfate and its activation methods for degradation of environmental organic pollutants: Review. Chem. Eng. J. 2017, 310, 41–62. [Google Scholar] [CrossRef]
  37. Ushani, U.; Lu, X.Q.; Wang, J.H.; Zhang, Z.Y.; Dai, J.J.; Tan, Y.J.; Wang, S.S.; Li, W.J.; Niu, C.X.; Cai, T.; et al. Sulfate radicals-based advanced oxidation technology in various environmental remediation: A state-of-the-art review. Chem. Eng. J. 2020, 402, 126232. [Google Scholar] [CrossRef]
  38. Olmez-Hanci, T.; Arslan-Alaton, I. Comparison of sulfate and hydroxyl radical based advanced oxidation of phenol. Chem. Eng. J. 2013, 224, 10–16. [Google Scholar] [CrossRef]
  39. Cai, H.Y.; Zou, J.; Lin, J.N.; Li, J.W.; Huang, Y.X.; Zhang, S.Y.; Yuan, B.L.; Ma, J. Sodium hydroxide-enhanced acetaminophen elimination in heat/peroxymonosulfate system: Production of singlet oxygen and hydroxyl radical. Chem. Eng. J. 2022, 429, 132438. [Google Scholar] [CrossRef]
  40. Ding, C.L.; Liu, Z.; Pan, S.Y.; Zhao, C.; Wang, Z.W.; Gao, B.Y.; Li, Q. Activation of peroxydisulfate via Fe@sulfur-doped carbon-supported nanocomposite for degradation of norfloxacin: Efficiency and mechanism. Chem. Eng. J. 2023, 460, 141729. [Google Scholar] [CrossRef]
  41. Dominguez, C.M.; Rodriguez, V.; Montero, E.; Romero, A.; Santos, A. Methanol-enhanced degradation of carbon tetrachloride by alkaline activation of persulfate: Kinetic model. Sci. Total Environ. 2019, 666, 631–640. [Google Scholar] [CrossRef]
  42. Wu, W.T.; Huang, Z.; Liu, Y.Y.; Hong, J.M.; Zhang, Q. Novel Co-MOF/GO triggered peroxymonosulfate activation: Insight into singlet oxygen generation mechanism from peroxymonosulfate self-decomposition and secondary radical conversion. Chem. Eng. J. 2023, 476, 146365. [Google Scholar] [CrossRef]
  43. Wang, G.L.; Liu, Y.C.; Dong, X.L.; Zhang, X.F. Transforming radical to non-radical pathway in peroxymonosulfate activation on nitrogen doped carbon sphere for enhanced removal of organic pollutants: Combined effect of nitrogen species and carbon structure. J. Hazard. Mater. 2022, 437, 129357. [Google Scholar] [CrossRef] [PubMed]
  44. Gao, Z.W.; Zhang, D.D.; Jun, Y.S. Does Tert-Butyl Alcohol Really Terminate the Oxidative Activity of •OH in Inorganic Redox Chemistry? J. Environ. Sci. Technol. 2021, 55, 10442–10450. [Google Scholar] [CrossRef] [PubMed]
  45. Chen, Z.Y.; Guo, J.Y.; Li, S.R.; Pu, L.; Huang, L. Insight in sulfadiazine degradation by peroxymonosulfate activated by polydopamine-derived nitrogen-doped carbon supported CoFe2O4: Co leaching inhibition and degradation enhancement. Ecotox. Environ. Saf. 2024, 285, 117126. [Google Scholar] [CrossRef] [PubMed]
  46. Ouyang, H.; Wu, C.; Qiu, X.H.; Tanaka, K.; Ohnuki, T.; Yu, Q.Q. New insight of Mn(III) in δ-MnO2 for peroxymonosulfate activation reaction: Via direct electron transfer or via free radical reactions. Environ. Res. 2023, 217, 114874. [Google Scholar] [CrossRef]
  47. Wei, H.; Zhao, J.; Shang, B.; Zhai, J. Carbamazepine degradation through peroxymonosulfate activation by α-MnO2 with the different exposed facets: Enhanced radical oxidation by facet-dependent surface effect. J. Environ. Chem. Eng. 2022, 10, 108532. [Google Scholar] [CrossRef]
  48. Fan, J.H.; Wang, Q.Q.; Yan, W.; Chen, J.B.; Zhou, X.F.; Xie, H.J. Mn3O4-g-C3N4 composite to activate peroxymonosulfate for organic pollutants degradation: Electron transfer and structure-dependence. J. Hazard. Mater. 2022, 434, 128818. [Google Scholar] [CrossRef]
  49. Cai, T.; Huang, H.; Deng, W.; Dai, Q.G.; Liu, W.; Wang, X.Y. Catalytic combustion of 1,2-dichlorobenzene at low temperature over Mn-modified Co3O4 catalysts. Appl. Catal. B-Environ. Energy 2015, 166, 393–405. [Google Scholar] [CrossRef]
  50. He, C.; Wang, Y.C.; Li, Z.Y.; Huang, Y.J.; Liao, Y.H.; Xia, D.H.; Lee, S.C. Facet Engineered α-MnO2 for Efficient Catalytic Ozonation of Odor CH3SH: Oxygen Vacancy-Induced Active Centers and Catalytic Mechanism. Environ. Sci. Technol. 2020, 54, 12771–12783. [Google Scholar] [CrossRef]
  51. Lu, Z.C.; Zhang, P.; Hu, C.; Li, F. Insights into singlet oxygen generation and electron-transfer process induced by a single-atom Cu catalyst with saturated Cu-N4 sites. iScience 2022, 25, 104930. [Google Scholar] [CrossRef]
  52. Yin, Y.; Li, W.N.; Hu, B.; Wang, Q.X.; Asif, A.H.; Shi, L.; Zhang, S.; Sun, H.Q.; Cui, S. Strengthened carbamazepine degradation via producing 1O2 with Ru-doped hollow tubular MoS2 under peroxymonosulfate activation. J. Environ. Chem. Eng. 2025, 13, 115328. [Google Scholar] [CrossRef]
  53. Lin, X.N.; Ge, Q.; Zhou, X.B.; Wang, Y.; Zhu, C.Y.; Liu, K.Y.; Wan, J.Q. Enhancement of Electron Transfer Between Fe/Mn Promotes Efficient Activation of Peroxomonosulfate by FeMn-NBC. Water 2025, 17, 1700. [Google Scholar] [CrossRef]
  54. Wei, H.X.; Zhao, J.J.; Rahaman, M.H.; Wang, Q.F.; Zhai, J. The catalytic activity of different Mn(III) species towards peroxymonosulfate activation for carbamazepine degradation. Catal. Commun. 2023, 173, 106563. [Google Scholar] [CrossRef]
  55. Yun, E.T.; Lee, J.H.; Kim, J.; Park, H.D.; Lee, J. Identifying the Nonradical Mechanism in the Peroxymonosulfate Activation Process: Singlet Oxygenation Versus Mediated Electron Transfer. Environ. Sci. Technol. 2018, 52, 7032–7042. [Google Scholar] [CrossRef] [PubMed]
  56. Deng, J.; Ya, C.; Ge, Y.J.; Cheng, Y.Q.; Chen, Y.J.; Xu, M.Y.; Wang, H.Y. 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]
  57. Li, X.G.; Guo, Y.X.; Yan, L.G.; Yan, T.; Song, W.; Feng, R.; Zhao, Y.W. Enhanced activation of peroxymonosulfate by ball-milled MoS2 for degradation of tetracycline: Boosting molybdenum activity by sulfur vacancies. Chem. Eng. J. 2022, 429, 132234. [Google Scholar] [CrossRef]
  58. Chen, Y.L.; Bai, X.; Ji, Y.T.; Shen, T. Reduced graphene oxide-supported hollow Co3O4@N-doped porous carbon as peroxymonosulfate activator for sulfamethoxazole degradation. Chem. Eng. J. 2022, 430, 132951. [Google Scholar] [CrossRef]
  59. Wang, S.Z.; Liu, Y.; Wang, J.L. Peroxymonosulfate Activation by Fe-Co-O-Codoped Graphite Carbon Nitride for Degradation of Sulfamethoxazole. Environ. Sci. Technol. 2020, 54, 10361–10369. [Google Scholar] [CrossRef]
  60. Shao, P.H.; Tian, J.Y.; Yang, F.; Duan, X.G.; Gao, S.S.; Shi, W.X.; Luo, X.B.; Cui, F.Y.; Luo, S.L.; Wang, S.B. Identification and Regulation of Active Sites on Nanodiamonds: Establishing a Highly Efficient Catalytic System for Oxidation of Organic Contaminants. Adv. Funct. Mater. 2018, 28, 1705295. [Google Scholar] [CrossRef]
  61. Ren, W.; Cheng, C.; Shao, P.H.; Luo, X.B.; Zhang, H.; Wang, S.B.; Duan, X.G. Origins of Electron-Transfer Regime in Persulfate-Based Nonradical Oxidation Processes. Environ. Sci. Technol. 2022, 56, 78–97. [Google Scholar] [CrossRef]
  62. Zhang, L.; Zhao, X.F.; Niu, C.G.; Tang, N.; Guo, H.; Wen, X.J.; Liang, C.; Zeng, G.M. Enhanced activation of peroxymonosulfate by magnetic Co3MnFeO6 nanoparticles for removal of carbamazepine: Efficiency, synergetic mechanism and stability. Chem. Eng. J. 2019, 362, 851–864. [Google Scholar] [CrossRef]
  63. Wu, J.X.; Cagnetta, G.; Wang, B.; Cui, Y.Z.; Deng, S.B.; Wang, Y.J.; Huang, J.; Yu, G. Efficient degradation of carbamazepine by organo-montmorillonite supported nCoFe2O4-activated peroxymonosulfate process. Chem. Eng. J. 2019, 368, 824–836. [Google Scholar] [CrossRef]
  64. Yang, J.C.E.; Zhu, M.P.; Dionysiou, D.D.; Yuan, B.L.; Fu, M.L. Interplay of bicarbonate and the oxygen-containing groups of carbon nanotubes dominated the metal-free activation of peroxymonosulfate. Chem. Eng. J. 2022, 430, 133102. [Google Scholar] [CrossRef]
Figure 1. Material Results Characterisation Chart. (a) SEM, (b) EDS mapping and TEM images of Ru/α-MnO2, (c) XRD, (d) BET, and (e) FT-IR analysis of α-MnO2 and Ru/α-MnO2.
Figure 1. Material Results Characterisation Chart. (a) SEM, (b) EDS mapping and TEM images of Ru/α-MnO2, (c) XRD, (d) BET, and (e) FT-IR analysis of α-MnO2 and Ru/α-MnO2.
Catalysts 15 01085 g001
Figure 2. The removal efficiency of CBZ by different AOPS systems. Conditions: [CBZ]0 = 10 mg/L, [PMS]0 = 200 mg/L = [α-MnO2]0 = [Rulatt/α-MnO2]0 = [Rusurf/α-MnO2]0 = 200 mg/L, time = 5 min, pH = 3.
Figure 2. The removal efficiency of CBZ by different AOPS systems. Conditions: [CBZ]0 = 10 mg/L, [PMS]0 = 200 mg/L = [α-MnO2]0 = [Rulatt/α-MnO2]0 = [Rusurf/α-MnO2]0 = 200 mg/L, time = 5 min, pH = 3.
Catalysts 15 01085 g002
Figure 3. (a) Effect of pH on CBZ degradation of Rulatt/α-MnO2/PMS system; (b) The corresponding fitted kinetics curves. Conditions: [CBZ]0 = 10 mg/L, [PMS]0 = 200 mg/L, [Rulatt/α-MnO2]0 = 200 mg/L.
Figure 3. (a) Effect of pH on CBZ degradation of Rulatt/α-MnO2/PMS system; (b) The corresponding fitted kinetics curves. Conditions: [CBZ]0 = 10 mg/L, [PMS]0 = 200 mg/L, [Rulatt/α-MnO2]0 = 200 mg/L.
Catalysts 15 01085 g003
Figure 4. (a) Effect of Rulatt/α-MnO2 dosage on CBZ degradation of Rulatt/α-MnO2/PMS system; (b) The corresponding fitted kinetics curves. Conditions: [CBZ]0 = 10 mg/L, [PMS]0 = 200 mg/L, pH = 3.
Figure 4. (a) Effect of Rulatt/α-MnO2 dosage on CBZ degradation of Rulatt/α-MnO2/PMS system; (b) The corresponding fitted kinetics curves. Conditions: [CBZ]0 = 10 mg/L, [PMS]0 = 200 mg/L, pH = 3.
Catalysts 15 01085 g004
Figure 5. (a) Effect of PMS dosage on CBZ degradation of Rulatt/α-MnO2/PMS system; (b) The corresponding fitted kinetics curves. Conditions: [CBZ]0 = 10 mg/L, [Rulatt/α-MnO2]0 = 200 mg/L, pH = 3.
Figure 5. (a) Effect of PMS dosage on CBZ degradation of Rulatt/α-MnO2/PMS system; (b) The corresponding fitted kinetics curves. Conditions: [CBZ]0 = 10 mg/L, [Rulatt/α-MnO2]0 = 200 mg/L, pH = 3.
Catalysts 15 01085 g005
Figure 6. (a) Effects of different scavengers on CBZ degradation via Rulatt-MnO2/PMS system; (b) The corresponding fitted kinetics curves. Conditions: [CBZ]0 = 10 mg/L, [Rulatt/α-MnO2]0 = 200 mg/L, [PMS]0 = 200 mg/L pH = 3, [extinguishing agent]0 = 50 mmol/L.
Figure 6. (a) Effects of different scavengers on CBZ degradation via Rulatt-MnO2/PMS system; (b) The corresponding fitted kinetics curves. Conditions: [CBZ]0 = 10 mg/L, [Rulatt/α-MnO2]0 = 200 mg/L, [PMS]0 = 200 mg/L pH = 3, [extinguishing agent]0 = 50 mmol/L.
Catalysts 15 01085 g006
Figure 7. EPR Characterization Results. (a) TEMP–1O2, (b) DMPO– O 2 , (c) D M P O SO 4 and DMPO–·OH.
Figure 7. EPR Characterization Results. (a) TEMP–1O2, (b) DMPO– O 2 , (c) D M P O SO 4 and DMPO–·OH.
Catalysts 15 01085 g007
Figure 8. Effects of H2PO4 on CBZ degradation via Rulatt-MnO2/PMS system. Conditions: [CBZ]0 = 10 mg/L, [Rulatt/α-MnO2]0 = 200 mg/L, [PMS]0 = 200 mg/L pH = 3, [H2PO4]0 = 10 mmol/L.
Figure 8. Effects of H2PO4 on CBZ degradation via Rulatt-MnO2/PMS system. Conditions: [CBZ]0 = 10 mg/L, [Rulatt/α-MnO2]0 = 200 mg/L, [PMS]0 = 200 mg/L pH = 3, [H2PO4]0 = 10 mmol/L.
Catalysts 15 01085 g008
Figure 9. XPS spectra of (a) of O 1s, (b) Mn 2p, (c) Ru 3p, and (d) Ru 3d of the fresh and used Rulatt/α-MnO2/PMS.
Figure 9. XPS spectra of (a) of O 1s, (b) Mn 2p, (c) Ru 3p, and (d) Ru 3d of the fresh and used Rulatt/α-MnO2/PMS.
Catalysts 15 01085 g009
Figure 10. (a) Effects of interfering ions on CBZ degradation via Rulatt/α-MnO2/PMS system; (b) The corresponding fitted kinetics curves. Conditions: [CBZ]0 = 10 mg/L, [Rulatt/α-MnO2]0 = 200 mg/L, [PMS]0 = 200 mg/L pH = 3, [interfering ions]0 = 10 mmol/L.
Figure 10. (a) Effects of interfering ions on CBZ degradation via Rulatt/α-MnO2/PMS system; (b) The corresponding fitted kinetics curves. Conditions: [CBZ]0 = 10 mg/L, [Rulatt/α-MnO2]0 = 200 mg/L, [PMS]0 = 200 mg/L pH = 3, [interfering ions]0 = 10 mmol/L.
Catalysts 15 01085 g010
Figure 11. Schematic diagram of the removal CBZ by Rulatt/α-MnO2/PMS.
Figure 11. Schematic diagram of the removal CBZ by Rulatt/α-MnO2/PMS.
Catalysts 15 01085 g011
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

Hu, P.; Qin, L.; Feng, M.; Cheng, Y.; Tang, P.; Xin, B.; Song, W.; Wang, Q.; Zhao, J. Ru-Modified α-MnO2 as an Efficient PMS Activator for Carbamazepine Degradation: Performance and Mechanism. Catalysts 2025, 15, 1085. https://doi.org/10.3390/catal15111085

AMA Style

Hu P, Qin L, Feng M, Cheng Y, Tang P, Xin B, Song W, Wang Q, Zhao J. Ru-Modified α-MnO2 as an Efficient PMS Activator for Carbamazepine Degradation: Performance and Mechanism. Catalysts. 2025; 15(11):1085. https://doi.org/10.3390/catal15111085

Chicago/Turabian Style

Hu, Panfeng, Long Qin, Manman Feng, Yuanling Cheng, Pan Tang, Beibei Xin, Wei Song, Quanfeng Wang, and Jujiao Zhao. 2025. "Ru-Modified α-MnO2 as an Efficient PMS Activator for Carbamazepine Degradation: Performance and Mechanism" Catalysts 15, no. 11: 1085. https://doi.org/10.3390/catal15111085

APA Style

Hu, P., Qin, L., Feng, M., Cheng, Y., Tang, P., Xin, B., Song, W., Wang, Q., & Zhao, J. (2025). Ru-Modified α-MnO2 as an Efficient PMS Activator for Carbamazepine Degradation: Performance and Mechanism. Catalysts, 15(11), 1085. https://doi.org/10.3390/catal15111085

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