Next Article in Journal
Ligand-Based Virtual Screening, Molecular Docking, and Molecular Dynamic Simulations of New β-Estrogen Receptor Activators with Potential for Pharmacological Obesity Treatment
Next Article in Special Issue
Role of Peripheral Coordination Boron in Electrocatalytic Nitrogen Reduction over N-Doped Graphene-Supported Single-Atom Catalysts
Previous Article in Journal
Two New Aluminoborates with 3D Porous-Layered Frameworks
Previous Article in Special Issue
Liquid-Phase Dehydration of Glycerol to Acrolein with ZSM-5-Based Catalysts in the Presence of a Dispersing Agent
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Peroxymonosulfate Activation by Facile Fabrication of α-MnO2 for Rhodamine B Degradation: Reaction Kinetics and Mechanism

1
School of Energy & Environment, Zhongyuan University of Technology, Zhengzhou 450007, China
2
College of Chemistry, Zhengzhou University, Zhengzhou 450001, China
3
International Education College, Henan Agricultural University, Zhengzhou 450002, China
4
School of Ecology & Environment, Zhengzhou University, Zhengzhou 450001, China
5
Henan Key Laboratory of Functional Salt Materials, Zhongyuan University of Technology, Zhengzhou 450007, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2023, 28(11), 4388; https://doi.org/10.3390/molecules28114388
Submission received: 6 April 2023 / Revised: 1 May 2023 / Accepted: 22 May 2023 / Published: 27 May 2023
(This article belongs to the Special Issue Research on Heterogeneous Catalysis)

Abstract

:
The persulfate-based advanced oxidation process has been an effective method for refractory organic pollutants’ degradation in aqueous phase. Herein, α-MnO2 with nanowire morphology was facially fabricated via a one-step hydrothermal method and successfully activated peroxymonosulfate (PMS) for Rhodamine B (RhB) degradation. Influencing factors, including the hydrothermal parameter, PMS concentration, α-MnO2 dosage, RhB concentration, initial pH, and anions, were systematically investigated. The corresponding reaction kinetics were further fitted by the pseudo-first-order kinetic. The RhB degradation mechanism via α-MnO2 activating PMS was proposed according to a series of quenching experiments and the UV-vis scanning spectrum. Results showed that α-MnO2 could effectively activate PMS to degrade RhB and has good repeatability. The catalytic RhB degradation reaction was accelerated by increasing the catalyst dosage and the PMS concentration. The effective RhB degradation performance can be attributed to the high content of surface hydroxyl groups and the greater reducibility of α-MnO2, and the contribution of different ROS (reactive oxygen species) was 1O2 > O 2 · > SO 4 · > ·OH.

1. Introduction

Dye wastewater, as a typical refractory organic wastewater, mainly emitted from the printing and textile industry, is characterized by toxicity, high chroma, and chemical oxygen demand (COD) [1]. The complex organic dyes with an aromatic structure may pose a serious threat to the water ecosystem and human health, with negative effects including aesthetic quality, photosynthetic activity, light penetration, and gas solubility [2,3]. Therefore, advancing dye wastewater treatment has drawn much attention. Conventional treatment technology for the elimination of organic pollutants can be achieved by filtration [4], adsorption [5], precipitation [6], and biological treatment [7]. However, these methods usually have the disadvantages of low efficiency and throughput requirements, which may also require further treatment of sorbent waste and membrane concentrate streams, and were inferior to meet deep organic components’ degradation and mineralization demand [8,9,10].
Advanced oxidation processes (AOPs) have been continuously developed since they were initially proposed by Hoigné and Bader (1976) [11] and Glaze (1987) [12]. A variety of reactive oxygen species (ROS) can be generated by oxidants under light irradiation [13], electricity [14], microwave [15,16], ultrasound [17], high temperatures [18], high pressure [19], etc. The targeted organic molecules can be effectively degraded into non-toxic inorganic salts, CO2 and H2O [14,20,21,22]. Among the emerging AOPs, persulfate-based oxidation processes utilize peroxydisulfate (PDS) and peroxymonosulfate (PMS) as major oxidants to induce ROS, featuring a higher oxidation potential (2.5–3.1 V), longer half-lifespan (3–4 × 10−5 s), wider pH operation range (2–8), cheap cost, and easy storage [23,24,25,26]. PDS and PMS exhibit differences in the activation pathway and nonradical reactions due to their structural differences. In comparison with PDS (O3SO-OSO3), the PMS molecule is asymmetrical (O3SO-OH), with the attached sulfite (SO3) on one side and hydrogen on the other side. Therefore, PMS is more active than PDS owing to the nucleophilic attacks by organic pollutants with electron-rich moieties [27]. The ROS produced by activating persulfate mainly include powerful oxidants such as sulfate radical ( SO 4 · ), hydroxyl radical (OH), superoxide radical ( O 2 · ), and singlet oxygen (1O2). Among them, non-selective ·OH can effectively degrade and mineralize various organic pollutants in a wide range of pH values [28]. Additionally, various persulfate activation methods are developing, such as physical (ultraviolet, ultrasound, heating), chemical (electrical, heterogenous catalysis), and combinations of multiple methods [29,30]. Of these, persulfate activation by introducing homogeneous and heterogeneous catalysts are more promising due to their high activation efficiency, low operation costs, and mild operating conditions, without external energy [31].
Numerous studies have been carried on different persulfate activation materials, including transitional metal ions (Mn2+, Co2+, Cu2+, Fe2+) [32,33], transitional metal oxides [34], perovskite [35], and spinel catalysts [36]. Among them, transition metal oxides are excellent candidates in environmental applications in catalytic ozonation, photocatalysis, and VOCs’ oxidation [37,38], due to their wide sources and variable valence states. Manganese oxides (MnO, MnO2, Mn2O3, Mn3O4) are non-precious and have a high natural abundance, and they have been selected as promising candidates for the catalytic persulfate activation [39]. Especially, MnO2 has various crystal phases (α, β, γ, and δ), versatile morphologies (nanosheets, nanorods, nanotubes, and nanowires), different structures (tetragonal, hexagonal pore channel, and symmetrical cubic structures), and diverse exposed crystal planes, which can influence the physicochemical properties and further affect the catalytic performance of PS activation [21]. For example, Yang et al. [40] fabricated α-MnO2 with {110} and {100} as exposed facets, and successfully activated PMS for the degradation of Orange I (OI). Wang et al. [41] investigated different crystal forms of MnO2 (α-MnO2, β-MnO2, γ-MnO2, and δ-MnO2) in activating PMS for the removal of Reactive Yellow X-RG.
In this study, α-MnO2 with nanowire morphology were prepared using the facile hydrothermal method with different manganese precursor ratios and heating times. The as-prepared α-MnO2 were further characterized by X-ray diffraction (XRD), Brunauer–Emmett–Teller (BET), Fourier transform infrared (FTIR), field-emission scanning electron microscopy (SEM), and electrochemical analysis methods. Rhodamine B (RhB), a typical azo dye, was taking as target pollutant and thus simulating the refractory oraganic wastewater. The catalytic PMS activation performance was investigated under ambient temperature at 25 ± 2 °C. Influencing factors, including the manganese precursors ratio, hydrothermal time, PMS and α-MnO2 dosage, RhB initial concentration, pH value, and anions, were systematically investigated and discussed. The pseudo-first-order kinetic was further fitted to evaluate the corresponding reaction kinetics. The RhB degradation mechanism was also discussed based on scavenger tests and UV-vis scanning results.

2. Results and Discussion

2.1. Catalysts’ Characterization Result

The crystal phase and crystallinity of the as-prepared α-MnO2 were confirmed by XRD analysis, as shown in Figure 1a. It can be clearly seen that all the α-MnO2 by different hydrothermal times and precursor ratios exhibited a high diffraction peak intensity, indicating the good crystallization of α-MnO2 by using hydrothermal synthesis. Compared with the standard PDF, the diffraction peaks on the XRD pattern of all samples were well-matched to α-MnO2 (JCPDS: 44-0141). Specifically, the diffraction peaks at 2θ = 12.67°, 17.94°, 28.66°, 37.50°, 41.94°, and 49.79° were observed in the curves for all the catalysts, and these peaks were corresponding to the (110), (200), (310), (211), (301), and (411) crystal faces of α-MnO2. The results suggest that the crystal structure of α-MnO2 will not be affected by either the Mn7+/Mn2+ ratios in the precursor or the hydrothermal time.
To evaluate the surface area and the pore size distribution, α-MnO2 was examined by N2 adsorption/desorption measurements, as shown in Figure 1b. The longer hydrothermal reaction led to a decreased trend of the surface area, with α-MnO2-6 (1:1) of 50.6 m2/g, α-MnO2-6 (2:3) of 44.8 m2/g, α-MnO2-12 (1:1) of 35.4 m2/g, and α-MnO2-12 (2:3) of 43.1 m2/g. To obtain the surface functional group, α-MnO2 was analyzed by FTIR, as shown in Figure 1c. The main absorption peaks at 3440 and 1626 cm−1 were attributed to α-MnO2 surface hydroxyl groups’ stretching and bending vibrations. The peaks at approximately 1121, 716, and around 500 cm−1 were attributed to the stretching vibration of the Mn–O and Mn–O–Mn bonds. The redox performance of α-MnO2 was also preliminarily analyzed by electrochemical tests. Specifically, the powder α-MnO2 catalyst was inherent to the titanium plate and was used as a working electrode, and RuO2-IrO2/Ti was used as an auxiliary electrode. As shown in Figure 1d, a higher current density and a greater reductive ability were observed in α-MnO2-12 (2:3) and α-MnO2-12 (1:1), compared with that of α-MnO2-6.
The surface morphology of α-MnO2 at different hydrothermal times is shown in Figure 2. Apparently, α-MnO2 fabricated under 6 and 12 h hydrothermal reactions showed similar nanowires. The enhancement of the hydrothermal reaction time resulted in little influence on the α-MnO2 morphology. Moreover, the diameter of 6 h-1:1 was smaller than that of 12 h-1:1, indicating a larger surface area, corresponding to the BET results in Figure 1b. Specifically, the diameter of 6 h-1:1 α-MnO2 was 30~50 nm, with an average length of 1.5~2.0 μm. After the 12 h hydrothermal reaction, the α-MnO2 average diameter and length increased to 40~80 nm and 2.0~4.0 μm, respectively. Through the element mapping analysis, it was observed that oxygen, potassium, and manganese were uniformly dispersed in α-MnO2 particles, thus providing abundant active sites for PMS activation.

2.2. α-MnO2 Catalytic PMS Activation Performance

2.2.1. Effect of Hydrothermal Preparation Parameters

The influence of different catalyst preparation parameters was studied and depicted in Figure 3a. Under the PMS alone condition, the RhB degradation process was not obvious within 30 min, which may be attributed to the weak PMS oxidation capacity towards RhB without further activation. When only α-MnO2 was added in RhB solution, the RhB degradation efficiency varied from 2.7% to 40.6% within 30 min for different α-MnO2. It can be inferred that the as-prepared nanowire α-MnO2 with a surface area of 35.4~50.6 m2/g exhibited adsorption and a slight catalytic degradation ability of RhB [42]. For α-MnO2-6 (1:1) and α-MnO2-6 (2:3), the RhB degradation efficiency reached 95.92% and 91.12% after 30 min. When the preparation time of the catalyst increased to 12 h, the RhB degradation efficiency was 88.87% for α-MnO2-12 (1:1) and 90.43% for α-MnO2-12 (2:3). The degradation kinetics was fitted based on the preliminary 20 min data of the degradation curve, and the results are shown in Figure 3b, which were coherent with the pseudo-first-order reaction. When the molar ratio of KMnO4 and MnSO4·H2O was 1:1 and 2:3 for the 12 h hydrothermal reaction, the first-order rate constant (Kobs) of the RhB removal rate was 0.0718 and 0.0779. For the 6 h hydrothermal reaction, when the molar ratio of KMnO4 and MnSO4·H2O was 1:1 and 2:3, the reaction rate constant values of the RhB removal rate were 0.0918 and 0.0808, respectively. The above experiments showed that both the hydrothermal time and the precursor ratio had a certain effect on the degradation rate of MnO2-activated PMS. We selected α-MnO2-12 (1:1) for the following catalytic PMS activation experiment.

2.2.2. Effect of PMS Dosage

The oxidant concentration determines the free radical quantity in the reaction system. The effect of different PMS amounts on catalytic RhB degradation performance was investigated. The PMS dosage was set at 0.01, 0.02, 0.05, 0.1, 0.2, and 0.5 g/L. As can be seen from Figure 4a, with the increase of the PMS concentration, the degradation efficiency of RhB significantly increased from 62.2% at 0.01 g/L to 94.4% at 0.5 g/L. As can be seen from Figure 4b, the removal rate of RhB fit well with the pseudo-first-order kinetic, with R2 values greater than 0.95. With the increase of the PMS concentration from 0.01 g/L to 0.5 g/L, the reaction rate constant value drastically increased from 0.037 min−1 to 0.123 min−1, indicating that the RhB degradation was strongly dependent on the PMS concentration. More reactive oxygen species was produced via the direct increase of the PMS concentration [43].

2.2.3. Effect of α-MnO2 Dosage

The α-MnO2 catalyst dosage was also investigated, as shown in Figure 5. The number of reaction sites was determined by the catalyst concentration in the heterogeneous catalytic reaction [44]. As can be seen from Figure 5a, the degradation rate of RhB increased from 60.15% to 94.97% after 30 min, as the dosage of α-MnO2 increased from 0.05 g/L to 1 g/L. As shown in Figure 5b, the α-MnO2 dosage increased from 0.05 g/L to 0.1 g/L, 0.2 g/L, 0.5 g/L, and 1 g/L. The reaction rate constant value increased from 0.02 min−1 to 0.03 min−1, 0.05 min−1, 0.13 min−1, and 0.15 min−1, respectively. The improvement of the degradation rate can be attributed to the increasing number of PMS-activating sites provided by improving the α-MnO2 dosage, thus accelerating the production of reactive oxygen species.

2.2.4. Effect of RhB Initial Concentration

The organic pollutants’ concentration is often variable in the effluent; therefore, the influence of the initial concentration of RhB should be considered. As shown in Figure 6a, the degradation effect of RhB decreased with the increase of the RhB concentration. When the initial concentration was 20 mg/L, 50 mg/L, and 80 mg/L, the degradation rates of α-MnO2 to RhB were 94.97%, 84.67%, and 78.31% after the 30 min reaction. The corresponding reaction rate constant values were 0.0779 min−1, 0.0643 min−1, and 0.0538 min−1, respectively (Figure 6b). The higher RhB initial concentration led to an increase in the RhB absolute amount in the solution, which requires more reactive oxygen species to react with the target pollutants and degradation intermediates. Under the same α-MnO2 and PMS dosages, the number of free radicals generated by catalytic activation was constant. Therefore, the increase of the RhB concentration led to a decrease of the degradation efficiency. However, the decolorization rate could remain at 78.31% with a corresponding RhB decolorization amount of 6.26 mg in 30 min, which exhibited a fascinating degradation performance compared to that of photocatalytic and biological methods when targeting high-chroma dye wastewater.

2.2.5. Effect of pH Value and Cations

The pH value of a solution can change the surface charge distribution of Mn-based materials and affect the interaction and reactivity between the catalyst and PMS [45]. The effect of pH on the degradation of RhB in the α-MnO2 activating PMS system was also investigated. As can be seen from Figure 7a, when the solution pH varied from 3, 5, 7, and 9, the degradation rates of RhB after 30 min were 88.54%, 87%, 90.43%, and 84.2%, respectively. When the pH was 9, the RhB degradation efficiency was slightly inhibited. One reason may be that PMS was unstable under alkaline conditions, and the spontaneous decomposition of PMS might occur when high concentrations of hydroxide ions are present. In addition, the recombination of SO 4 · and ·OH to form HSO 4 and oxygen may have occurred at high pH values [42,46]. Adjusting the pH value from 3 to 9 resulted in an overall slight influence on the RhB degradation efficiency, indicating that the α-MnO2 activating PMS system was suitable under a wide range of pH conditions. It should be noted that the pH value slightly decreased when PMS was added, and at the end of the reaction, the pH value remained at 2.91 to 3.86 when the initial pH was from 3 to 9.
Inorganic anions and natural organic matter (NOM) coexist in natural water and can influence the organic pollutants’ oxidation process depending on the competitive reaction of anions with different ROS [47]. Therefore, it is also necessary to investigate the effect of anions and NOM on RhB degradation in the α-MnO2 activating PMS system. Four common anions ( H C O 3 , Cl, NO3, and H2 P O 4 ) and humic acid (HA) were added into the RhB solution before each test. As shown in Figure 7b, HA, NO3, H2 P O 4 , and HCO3 inhibited the RhB degradation efficiency to different extents, while adding 2 mM of Cl had little effect on RhB degradation: the degradation rate of RhB after the 30 min reaction reached 91.75%. The RhB degradation efficiency decreased to 47.35% at 30 min with 10 mg/L of HA. This can be explained by the fact that HA with abundant electron sites competed with RhB molecules for the electron-loving groups ( SO 4 · and ·OH) and the active sites (−OH) on the surface of MnO2 [48]. H C O 3 and H2PO4 greatly exhibited RhB degradation. When 2 mM of H C O 3 was added, the degradation rate of RhB decreased from 90.43% to 16.5%, and owing to this, HCO3 could react with SO 4 · and ·OH, followed by generation of carbonic acid with a weak oxidation ability, as shown in Equations (1)–(4) [48]. When 2 mM of H2 P O 4 was added, the degradation of RhB was significantly inhibited, and the removal rate of RhB decreased from 90.43% to 31.97% within 30 min. This is because large amounts of SO 4 · and ·OH react with H2PO4. N O 3 could interact with the free electrons when PMS was activated and convert to NO2, which could scavenge part of the ROS [49]. The RhB degradation rate slightly declined to 85.1% after inducement of 2 mM of NO2.
HCO 3 - + SO 4 · - SO 4 2 - + HCO 3 ·
HCO 3 - + · OH OH - + HCO 3 ·
CO 3 2 - + SO 4 · - SO 4 2 - + CO 3 ·
CO 3 2 - + · OH OH - + CO 3 ·

2.2.6. Radical Scavenger Tests

In order to unveil the degradation mechanism of RhB in the α-MnO2 activating PMS system, quenching experiments were conducted to identify the potential ROS contributions. Note that MeOH and TBA were employed as SO 4 · and ·OH scavengers, and p-BQ and FFA were chosen to quench O 2 · and 1O2, respectively. As shown in Figure 8, after the 30 min reaction, the RhB degradation efficiency was inhibited by 2.62% and 7.59% when adding 0.1 M of TBA and MeOH, respectively. While the RhB degradation efficiency further decreased to 72.80% and 65.06%, the RhB degradation efficiency was inhibited by 17.64% and 25.34% by introducing p-BQ and FFA as quenchers, respectively. Based on the radical results, we can infer that all the above free radicals are involved in the degradation reaction, and the contributions to RhB degradation followed the order of: 1O2 > O 2 · > SO 4 · > ·OH. O 2 · and 1O2 were the main ROS for RhB degradation [40], and 1O2 may have also partly originated from PMS self-decomposition [50].

2.2.7. Repeatability

The powder catalyst has certain limitations due to the separation and collection handling in industrial applications. Herein, we tested the repeatability of α-MnO2 for four cycling catalytic reactions. After each use, the catalyst was retained for direct next cycle reaction without further treatment. As can be seen from Figure 9, the degradation efficiency showed a declining trend with the increase of the cycle times overall, and the degradation efficiency of RhB in water did not significantly decrease after the initial two cycles. After three cycles, the RhB degradation efficiency decreased from 90.43% to 76.83%, which may be due to the loss of the catalyst in the cycling tests, the adsorption of RhB degradation intermediates, and the leaching of manganese ions in activating the PMS process. The cycling tests indicated that the as-prepared α-MnO2 has good reusability.

2.3. RhB Degradation Mechnism

The UV-visible spectra of the reactive solution in α-MnO2 activating PMS degradation of RhB at different reaction times is shown in Figure 10. It was apparent that all major absorbance peak intensities rapidly decreased with the increasing reaction time, indicating that the RhB molecule gradually degraded. The main absorption peak of the solution at 553 nm sharply weakened, and finally, there was no absorption peak intensity. When the catalytic reaction time was 25 min, the chroma of the RhB solution became colorless, indicating that the chromophoric group had been destroyed. The reduction of the peaks at 553 nm mainly resulted from the breakage of the conjugated structure of C=N and C=O groups, which was responsible for the fading color of RhB. Based on the TOC measurement, the TOC value decreased from 3.18 mg/L to 0.304 mg/L after the 40 min reaction. Then, 90% of TOC was removed during RhB degradation, indicating that the intermediate degradation products were mineralized.

2.4. Possible Activation Mechanism

Based on the results, α-MnO2 activating PMS of the RhB degradation mechanism was proposed and depicted in Figure 11. Firstly, HSO−5 in solution was attached on the surface of α-MnO2. Different valences (Mn2+, Mn3+, and Mn4+) and surface −OH on the α-MnO2 surface acted as active sites to react with   HSO 5 - , thus generating sulfur-based radicals ( SO 5 · and SO 4 · ) (Equations (5)–(8)) [25], of which, a portion of SO 4 · would convert into ·OH through Equations (9) and (10) [21]. Secondary radicals, O 2 · - , were generated via Equations (11)–(13), and the non-radical 1O2 was produced via Equations (14) and (15) [51]. All the above ROS contributed to the RhB degradation process (Equation (16)).
Mn 4 + +   HSO 5 - Mn 3 + + SO 5 · - + H +
Mn 3 + +   HSO 5   - Mn 4 + + SO 4 · - + - OH  
Mn 2 + + HSO 5 -   Mn 3 + + SO 5 · - + - OH
Mn 3 + + HSO 5 - Mn 2 + + SO 5 · - + H +
SO 4 · - + H 2 O   · OH   + SO 5 2 - + H +
SO 4 · - + - OH   · OH   +   SO 4 2 -
HSO 5 - + H 2 O   H 2 O 2 +   HSO 4 -
H 2 O 2 + · OH     HO 2 · +   H 2 O  
HO 2 · + H + O 2 · -
2 O 2 · - +   H + O 2 1 + H 2 O 2
· OH   O 2 1 + - OH
O 2 1 /   O 2 · - /   SO 4 · - / · OH + RhB     intermediates   CO 2   + H 2 O

3. Materials and Methods

3.1. Chemicals

All chemicals were of analytical grade and used directly without further purification, unless otherwise specified. α-MnO2 were fabricated using KMnO4 and MnSO4·H2O as precursors. Peroxymonosulfate (2KHSO5·KHSO4·K2SO4) and Rhodamine B (RhB) were bought from Macklin Biochemical Technology and used as received. Potassium chloride (KCI), sodium bicarbonate (NaHCO3), and monopotassium phosphate (KH2PO4) were used to investigate the cation influence of the RhB degradation efficiency. The radical scavengers used in this study included tert-Butanol (TBA), p-Benzoquinone (p-BQ), methanol (MeOH), and furfuryl alcohol (FFA), respectively. The pH value was adjusted using sulfur acid solution and sodium hydroxide.

3.2. Catalysts’ Preparation

The α-MnO2 catalysts with nanowire morphology were synthesized via the facile hydrothermal method. Specifically, stoichiometric amounts of KMnO4 and MnSO4·H2O with Mn7+:Mn2+ ratios of 1:1 and 2:3 were dissolved in 80 mL of deionized water under vigorous stirring for 30 min. The obtained solid-solution mixture was further transferred to a 100 mL hydrothermal kettle made of Teflon liner and heated at 150 °C for 6 and 12 h. Ac-cording to the precursors’ different molar ratios of KMnO4 and MnSO4·H2O, the prepared α-MnO2 was denoted as α-MnO2-6 (1:1), α-MnO2-6 (2:3), α-MnO2-12 (1:1), and α-MnO2-12 (2:3), respectively. After cooling to room temperature, a solid material was obtained after filtration of the solid-solution mixture, washed, and dried at 80 °C overnight.

3.3. Analytical Instruments

The crystal phase of the prepared α-MnO2 crystal structure was determined by powder X-ray diffraction (XRD-6100, Shimadzu, Kyoto, Japan), using Cu Ka as a radiation source and a scanning range of 10–90° at a scanning speed of 5°·min−1. The N2 adsorption and desorption curve of α-MnO2 was evaluated over an Autosorb-iQ instrument (Quantachrome Corporation, Boynton Beach, FL, USA). Before testing, the sample was degassed at 300 °C for 6 h, and then the N2 adsorption/desorption test was conducted under liquid N2 at 196 °C. The specific surface area of the catalyst was determined using the multi-point Brunauer–Emmett–Teller (BET) method. The surface morphology of the sample was evaluated by a field-emission scanning electron microscope (FESEM: JEOL JSM-7800F, Tokyo, Japan) and the surface elements’ composition (Mn, K, and O) was surveyed via energy-dispersive spectroscopy (EDS). Fourier transform infrared (FTIR) spectra were recorded on a Nicolet 6700 spectrometer (Thermo Fisher, Waltham, MA, USA). The electrochemical measurements were carried out on a CHI660E electrochemical workstation.

3.4. Experimental Procedure

RhB degradation experiments by α-MnO2 activating PMS were conducted in a beaker containing 100 mL of RhB solution under continuous stirring at ambient temperature (25 ± 2 °C). Catalyst was initially added to RhB solution for 20 min to achieve saturated adsorption before the inducement of PMS. Samples of the RhB degradation solution were further withdrawn at 5 min intervals using a syringe and analyzed using the DR6000 UV-vis Spectrophotometer (Hach, Ames, IA, USA). The decolorization of RhB was investigated by measuring the maximum absorbance at a wavelength of 553 nm and by computing the concentration from the calibration curve. In order to avoid the adsorption influence of the dye color via membrane filtration, the RhB and degradation samples of 1 mL were withdrawn using a 5 mm optical cell and further analyzed without filtration after quenching with an excess amount of Na2S2O8. RhB absorption spectra were recorded at different intervals at a scanning wavelength range of 200–800 nm to analyze the degradation process. The degradation rate (η) of RhB and the pseudo-first-order reaction kinetics were calculated as follows:
η = C/C0 × 100%
ln(C/C0) = −kt
where C and C0 are the RhB concentration at the sampling time and the initial concentration of RhB (mg/L), according to the absorbance at 553 nm, k is the pseudo-first-order reaction rate constant (min−1), and t is the reaction time (min).
In order to investigate the role of reactive oxygen species (ROS) in the catalytic system, methanol (MeOH), p-Benzoquinone (p-BQ), tert-butanol (TBA), and furfuryl alcohol (FFA) were used as SO 4 · , O 2 · , ·OH, and O 2 1 scavengers, respectively [8,40]. In the repeatable experiment, the catalyst remained for the direct next cycle reaction, without further treatment after each use.

4. Conclusions

In this study, α-MnO2 was successfully prepared via a hydrothermal method, which exhibited higher catalytic activity in PMS activation to degrade RhB. The influencing factors, including α-MnO2 preparation, PMS and α-MnO2 dosage, RhB initial concentration, pH value, and anions, were systematically investigated. The as-prepared α-MnO2 with nanowire morphology exhibited good crystallization, with little affection of the hydrothermal time and the precursor ratio. The effective RhB degradation performance can be attributed to the high content of surface hydroxyl groups and the greater reducibility of α-MnO2. More reactive oxygen species was produced via the direct increase of α-MnO2 and the PMS concentration. The α-MnO2 activating PMS process is also capable of high-chroma dye wastewater treatment and has a wide pH value range. After four cycling times, the RhB degradation retained good stability. The main ROS that participated in RhB degradation followed the contribution trend of: O 2 1 > O 2 · > SO 4 · > ·OH.

Author Contributions

Conceptualization, methodology, investigation, writing—original draft, funding acquisition, J.L. (Juexiu Li); investigation, data curation, Q.S.; visualization, data curation, M.S., J.L. (Jinming Liu) and R.Z.; resources, J.C. and X.W.; supervision, Y.L. and W.G.; writing—review and editing, P.L. and K.C. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the Key Research and Development Program of Henan Province (No. 212102310506), the Natural Science Foundation of Henan (No. 202300410518), the Natural Science Foundation of Zhongyuan University of Technology (No. K2022QN027), the Postgraduate Education Quality Improving Project of Zhongyuan University of Technology (No. JG202217, ALK202309), and the Discipline Strength Improving Project of Zhongyuan University of Technology (No. SD202242). The College Students Innovation Training Program of China (202210465037), the College Students Innovation Training Program of Henan Province (202210465052), and the College Students Innovation Training Program of Zhongyuan University of Technology (202210465121) are also acknowledged. The authors would also like to thank Suzhou Deyo Bot Advanced Materials Co., Ltd. (www.dy-test.com) for providing support on material characterization.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data will be made available upon request from the corresponding author. The data are not publicly available due to privacy.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of α-MnO2 are available from the authors.

References

  1. Bal, G.; Thakur, A. Distinct approaches of removal of dyes from wastewater: A review. Mater. Today Proc. 2021, 50, 1575–1579. [Google Scholar] [CrossRef]
  2. Akar, S.T.; Aslan, S.; Akar, T. Conversion of natural mineral to effective geosorbent by coating MnO2 and its application potential for dye contaminated wastewaters. J. Clean. Prod. 2018, 189, 887–897. [Google Scholar] [CrossRef]
  3. Cao, M.; Zhuang, Z.; Liu, Y.; Zhang, Z.; Xuan, J.; Zhang, Q.; Wang, W. Peptide-mediated green synthesis of the MnO2@ZIF-8 core–shell nanoparticles for efficient removal of pollutant dyes from wastewater via a synergistic process. J. Colloid Interface Sci. 2022, 608, 2779–2790. [Google Scholar] [CrossRef] [PubMed]
  4. Hube, S.; Eskafi, M.; Hrafnkelsdóttir, K.F.; Bjarnadóttir, B.; Bjarnadóttir, M.A.; Axelsdóttir, S.; Wu, B. Direct membrane filtration for wastewater treatment and resource recovery: A review. Sci. Total Environ. 2020, 710, 136375. [Google Scholar] [CrossRef] [PubMed]
  5. Liu, Y.; Zhang, X.; Wang, J. A critical review of various adsorbents for selective removal of nitrate from water: Structure, performance and mechanism. Chemosphere 2022, 291, 132728. [Google Scholar] [CrossRef]
  6. Lapointe, M.; Farner, J.M.; Hernandez, L.M.; Tufenkji, N. Understanding and Improving Microplastic Removal during Water Treatment: Impact of Coagulation and Flocculation. Environ. Sci. Technol. 2020, 54, 8719–8727. [Google Scholar] [CrossRef]
  7. Singh, A.; Pal, D.B.; Mohammad, A.; Alhazmi, A.; Haque, S.; Yoon, T.; Srivastava, N.; Gupta, V.K. Biological remediation technologies for dyes and heavy metals in wastewater treatment: New insight. Bioresour. Technol. 2022, 343, 126154. [Google Scholar] [CrossRef]
  8. 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]
  9. Li, X.; Zhang, S.; Yu, M.; Xu, H.; Lv, J.; Yang, S.; Zhu, X.; Li, L. One-pot pyrolysis method for synthesis of Fe/N co-doped biochar as an effective peroxymonosulfate activator for RhB degradation. J. Taiwan Inst. Chem. Eng. 2021, 128, 209–219. [Google Scholar] [CrossRef]
  10. Luo, H.; Guo, J.; Shen, T.; Zhou, H.; Liang, J.; Yuan, S. Study on the catalytic performance of LaMnO3 for the RhB degradation. J. Taiwan Inst. Chem. Eng. 2020, 109, 15–25. [Google Scholar] [CrossRef]
  11. Hoigné, J.; Bader, H. The role of hydroxyl radical reactions in ozonation processes in aqueous solutions. Water Res. 1976, 10, 377–386. [Google Scholar] [CrossRef]
  12. Glaze, W.H.; Kang, J.-W.; Chapin, D.H. The Chemistry of Water Treatment Processes Involving Ozone, Hydrogen Peroxide and Ultraviolet Radiation. Ozone Sci. Eng. 1987, 9, 335–352. [Google Scholar] [CrossRef]
  13. Wu, S.; Hu, H.; Lin, Y.; Zhang, J.; Hu, Y.H. Visible light photocatalytic degradation of tetracycline over TiO2. Chem. Eng. J. 2020, 382, 122842. [Google Scholar] [CrossRef]
  14. Long, L.; Bai, C.; Zhou, X.; Zhang, S.; Zhang, Y.; Chen, C.; He, J.; Song, C.; Yang, G. A novel strategy for promoting PMS activation: Enhanced utilization of side reactions. Sep. Purif. Technol. 2022, 297, 121432. [Google Scholar] [CrossRef]
  15. Qi, C.; Liu, X.; Lin, C.; Zhang, H.; Li, X.; Ma, J. Activation of peroxymonosulfate by microwave irradiation for degradation of organic contaminants. Chem. Eng. J. 2017, 315, 201–209. [Google Scholar] [CrossRef]
  16. Qi, C.; Liu, X.; Lin, C.; Zhang, X.; Ma, J.; Tan, H.; Ye, W. Degradation of sulfamethoxazole by microwave-activated persulfate: Kinetics, mechanism and acute toxicity. Chem. Eng. J. 2014, 249, 6–14. [Google Scholar] [CrossRef]
  17. Xu, D.; Ma, H. Degradation of rhodamine B in water by ultrasound-assisted TiO2 photocatalysis. J. Clean. Prod. 2021, 313, 127758. [Google Scholar] [CrossRef]
  18. Arvaniti, O.S.; Ioannidi, A.A.; Mantzavinos, D.; Frontistis, Z. Heat-activated persulfate for the degradation of micropollutants in water: A comprehensive review and future perspectives. J. Environ. Manag. 2022, 318, 115568. [Google Scholar] [CrossRef]
  19. Manz, K.E.; Adams, T.J.; Carter, K.E. Furfural degradation through heat-activated persulfate: Impacts of simulated brine and elevated pressures. Chem. Eng. J. 2018, 353, 727–735. [Google Scholar] [CrossRef]
  20. Chen, G.; Zhang, X.; Gao, Y.; Zhu, G.; Cheng, Q.; Cheng, X. Novel magnetic MnO2/MnFe2O4 nanocomposite as a heterogeneous catalyst for activation of peroxymonosulfate (PMS) toward oxidation of organic pollutants. Sep. Purif. Technol. 2019, 213, 456–464. [Google Scholar] [CrossRef]
  21. Wang, Z.; Wang, Z.; Li, W.; Lan, Y.; Chen, C. Performance comparison and mechanism investigation of Co3O4-modified different crystallographic MnO2 (α, β, γ, and δ) as an activator of peroxymonosulfate (PMS) for sulfisoxazole degradation. Chem. Eng. J. 2022, 427, 130888. [Google Scholar] [CrossRef]
  22. Zhou, Y.; Feng, S.; Duan, X.; Zheng, W.; Shao, C.; Wu, W.; Jiang, Z.; Lai, W. MnO2/UIO-66 improves the catalysed degradation of oxytetracycline under UV/H2O2/PMS system. J. Solid State Chem. 2021, 300, 122231. [Google Scholar] [CrossRef]
  23. Koo, P.-L.; Jaafar, N.F.; Yap, P.-S.; Oh, W.-D. A review on the application of perovskite as peroxymonosulfate activator for organic pollutants removal. J. Environ. Chem. Eng. 2022, 10, 107093. [Google Scholar] [CrossRef]
  24. Zhao, Y.; Huang, B.; An, H.; Dong, G.; Feng, J.; Wei, T.; Ren, Y.; Ma, J. Enhanced activation of peroxymonosulfate by Sr-doped LaFeO3 perovskite for Orange I degradation in the water. Sep. Purif. Technol. 2021, 256, 117838. [Google Scholar] [CrossRef]
  25. Zhu, L.; Shi, Z.; Deng, L.; Duan, Y. Efficient degradation of sulfadiazine using magnetically recoverable MnFe2O4/δ-MnO2 hybrid as a heterogeneous catalyst of peroxymonosulfate. Colloids Surf. A Physicochem. Eng. Asp. 2021, 609, 125637. [Google Scholar] [CrossRef]
  26. Zhu, K.; Bin, Q.; Shen, Y.; Huang, J.; He, D.; Chen, W. In-situ formed N-doped bamboo-like carbon nanotubes encapsulated with Fe nanoparticles supported by biochar as highly efficient catalyst for activation of persulfate (PS) toward degradation of organic pollutants. Chem. Eng. J. 2020, 402, 126090. [Google Scholar] [CrossRef]
  27. Ding, Y.; Wang, X.; Fu, L.; Peng, X.; Pan, C.; Mao, Q.; Wang, C.; Yan, J. Nonradicals induced degradation of organic pollutants by peroxydisulfate (PDS) and peroxymonosulfate (PMS): Recent advances and perspective. Sci. Total Environ. 2021, 765, 142794. [Google Scholar] [CrossRef] [PubMed]
  28. Xie, J.; Zhang, C.; Waite, T.D. Hydroxyl radicals in anodic oxidation systems: Generation, identification and quantification. Water Res. 2022, 217, 118425. [Google Scholar] [CrossRef] [PubMed]
  29. Fu, R.; Zhang, P.-S.; Jiang, Y.-X.; Sun, L.; Sun, X.-H. Wastewater treatment by anodic oxidation in electrochemical advanced oxidation process: Advance in mechanism, direct and indirect oxidation detection methods. Chemosphere 2023, 311, 136993. [Google Scholar] [CrossRef] [PubMed]
  30. Lin, C.; Li, J.-L.; Li, X.; Yang, S.; Luo, W.; Zhang, Y.; Kim, S.-H.; Kim, D.-H.; Shinde, S.S.; Li, Y.-F.; et al. In-situ reconstructed Ru atom array on α-MnO2 with enhanced performance for acidic water oxidation. Nat. Catal. 2021, 4, 1012–1023. [Google Scholar] [CrossRef]
  31. Jiang, Z.-R.; Wang, P.; Zhou, Y.-X.; Wang, C.; Jiang, J.; Lan, Y.; Chen, C. Fabrication of a 3D-blocky catalyst (CoMnOx@sponge) via mooring Co-Mn bimetallic oxide on sponge to activate peroxymonosulfate for convenient and efficient degradation of sulfonamide antibiotics. Chem. Eng. J. 2022, 446, 137306. [Google Scholar] [CrossRef]
  32. Peng, J.; Wang, Z.; Wang, S.; Liu, J.; Zhang, Y.; Wang, B.; Gong, Z.; Wang, M.; Dong, H.; Shi, J.; et al. Enhanced removal of methylparaben mediated by cobalt/carbon nanotubes (Co/CNTs) activated peroxymonosulfate in chloride-containing water: Reaction kinetics, mechanisms and pathways. Chem. Eng. J. 2021, 409, 128176. [Google Scholar] [CrossRef]
  33. Qi, C.; Wen, Y.; Zhao, Y.; Dai, Y.; Li, Y.; Xu, C.; Yang, S.; He, H. Enhanced degradation of organic contaminants by Fe(III)/peroxymonosulfate process with l-cysteine. Chin. Chem. Lett. 2022, 33, 2125–2128. [Google Scholar] [CrossRef]
  34. Pan, M.; Wang, N.; Weng, Z.; Zou, X.; Huang, X. The synergistic activation of peroxymonosulfate for the degradation of Acid Scarlet GR by palygorskite/MnO2/Fe3O4 nanocomposites. Dalton Trans. 2023, 52, 1009–1020. [Google Scholar] [CrossRef]
  35. Wang, T.; Qian, X.; Yue, D.; Yan, X.; Yamashita, H.; Zhao, Y. CaMnO3 perovskite nanocrystals for efficient peroxydisulfate activation. Chem. Eng. J. 2020, 398, 125638. [Google Scholar] [CrossRef]
  36. Manos, D.; Miserli, K.; Konstantinou, I. Perovskite and Spinel Catalysts for Sulfate Radical-Based Advanced Oxidation of Organic Pollutants in Water and Wastewater Systems. Catalysts 2020, 10, 1299. [Google Scholar] [CrossRef]
  37. Guo, M.; Fang, R.; Liu, X.; Yang, Z. Experimental study of volatile organic compounds catalytic combustion on Cu-Mn catalysts with different carriers. Int. J. Energy Res. 2021, 45, 8749–8762. [Google Scholar] [CrossRef]
  38. Qin, C.; Guo, M.; Jiang, C.; Yu, R.; Huang, J.; Yan, D.; Li, S.; Dang, X. Simultaneous oxidation of toluene and ethyl acetate by dielectric barrier discharge combined with Fe, Mn and Mo catalysts. Sci. Total Environ. 2021, 782, 146931. [Google Scholar] [CrossRef]
  39. Zheng, X.; Niu, X.; Zhang, D.; Lv, M.; Ye, X.; Ma, J.; Lin, Z.; Fu, M. Metal-based catalysts for persulfate and peroxymonosulfate activation in heterogeneous ways: A review. Chem. Eng. J. 2022, 429, 132323. [Google Scholar] [CrossRef]
  40. Yang, Y.; Zhao, Y.; Zong, Y.; Wu, R.; Zhang, M.; Feng, J.; Wei, T.; Ren, Y.; Ma, J. Activation of peroxymonosulfate by α-MnO2 for Orange Ⅰ removal in water. Environ. Res. 2022, 210, 112919. [Google Scholar] [CrossRef]
  41. Wang, J.; Zhang, D.; Nie, F.; Zhang, R.; Fang, X.; Wang, Y. The role of MnO2 crystal morphological scale and crystal structure in selective catalytic degradation of azo dye. Environ. Sci. Pollut. Res. Int. 2023, 30, 15377–15391. [Google Scholar] [CrossRef]
  42. Wu, Y.; Li, Y.; He, J.; Fang, X.; Hong, P.; Nie, M.; Yang, W.; Xie, C.; Wu, Z.; Zhang, K.; et al. Nano-hybrids of needle-like MnO2 on graphene oxide coupled with peroxymonosulfate for enhanced degradation of norfloxacin: A comparative study and probable degradation pathway. J. Colloid Interface Sci. 2020, 562, 1–11. [Google Scholar] [CrossRef]
  43. Yi, H.; Wang, Y.; Diao, L.; Xin, Y.; Chai, C.; Cui, D.; Ma, D. Ultrasonic treatment enhances the formation of oxygen vacancies and trivalent manganese on α-MnO2 surfaces: Mechanism and application. J. Colloid Interface Sci. 2022, 626, 629–638. [Google Scholar] [CrossRef] [PubMed]
  44. Wang, Y.; Liu, M.; Hu, C.; Xin, Y.; Ma, D.; Gao, M.; Xie, H. Enhanced MnO2/peroxymonosulfate activation for phthalic acid esters degradation: Regulation of oxygen vacancy. Chem. Eng. J. 2022, 433, 134048. [Google Scholar] [CrossRef]
  45. Xiao, J.; Li, R.; Dong, H.; Li, Y.; Li, L.; Xiao, S.; Jin, Z. Activation of sulfite via zero-valent iron-manganese bimetallic nanomaterials for enhanced sulfamethazine removal in aqueous solution: Key roles of Fe/Mn molar ratio and solution pH. Sep. Purif. Technol. 2022, 297, 121479. [Google Scholar] [CrossRef]
  46. Feng, M.; Qu, R.; Zhang, X.; Sun, P.; Sui, Y.; Wang, L.; Wang, Z. Degradation of flumequine in aqueous solution by persulfate activated with common methods and polyhydroquinone-coated magnetite/multi-walled carbon nanotubes catalysts. Water Res. 2015, 85, 1–10. [Google Scholar] [CrossRef] [PubMed]
  47. Wei, Y.; Miao, J.; Ge, J.; Lang, J.; Yu, C.; Zhang, L.; Alvarez, P.J.J.; Long, M. Ultrahigh Peroxymonosulfate Utilization Efficiency over CuO Nanosheets via Heterogeneous Cu(III) Formation and Preferential Electron Transfer during Degradation of Phenols. Environ. Sci. Technol. 2022, 56, 8984–8992. [Google Scholar] [CrossRef] [PubMed]
  48. Wang, Z.; Han, Y.; Fan, W.; Wang, Y.; Huang, L. Shell-core MnO2/Carbon@Carbon nanotubes synthesized by a facile one-pot method for peroxymonosulfate oxidation of tetracycline. Sep. Purif. Technol. 2021, 278, 119558. [Google Scholar] [CrossRef]
  49. Zhao, Y.; Wang, H.; Ji, J.; Li, X.; Yuan, X.; Duan, A.; Guan, X.; Jiang, L.; Li, Y. Recycling of waste power lithium-ion batteries to prepare nickel/cobalt/manganese-containing catalysts with inter-valence cobalt/manganese synergistic effect for peroxymonosulfate activation. J. Colloid Interface Sci. 2022, 626, 564–580. [Google Scholar] [CrossRef]
  50. Yang, Z.; Wang, Z.; Liang, G.; Zhang, X.; Xie, X. Catalyst bridging-mediated electron transfer for nonradical degradation of bisphenol A via natural manganese ore-cornstalk biochar composite activated peroxymonosulfate. Chem. Eng. J. 2021, 426, 131777. [Google Scholar] [CrossRef]
  51. Xu, L.S.; Sun, X.B.; Hong, J.-M.; Zhang, Q. Peroxymonosulfate activation by α-MnO2/MnFe2O4 for norfloxacin degradation: Efficiency and mechanism. J. Phys. Chem. Solids 2021, 153, 110029. [Google Scholar] [CrossRef]
Figure 1. XRD patterns (a), N2 adsorption/desorption isotherm (b), FTIR spectra (c), and cyclic voltammetry curves, (d) of α-MnO2 under different preparation conditions.
Figure 1. XRD patterns (a), N2 adsorption/desorption isotherm (b), FTIR spectra (c), and cyclic voltammetry curves, (d) of α-MnO2 under different preparation conditions.
Molecules 28 04388 g001
Figure 2. SEM images of α-MnO2-6 h 1:1 (ac) and α-MnO2-12 h 1:1 (df), and elemental mapping results of α-MnO2-6 h 1:1.
Figure 2. SEM images of α-MnO2-6 h 1:1 (ac) and α-MnO2-12 h 1:1 (df), and elemental mapping results of α-MnO2-6 h 1:1.
Molecules 28 04388 g002
Figure 3. Effect of different α-MnO2 preparation parameters on RhB degradation (a) and corresponding pseudo-first-order kinetics (b) (reaction condition: α-MnO2 = 0.2 g/L, RhB = 20 mg/L, PMS = 0.05 g/L, initial pH = 7, temperature = 25 ± 2 °C).
Figure 3. Effect of different α-MnO2 preparation parameters on RhB degradation (a) and corresponding pseudo-first-order kinetics (b) (reaction condition: α-MnO2 = 0.2 g/L, RhB = 20 mg/L, PMS = 0.05 g/L, initial pH = 7, temperature = 25 ± 2 °C).
Molecules 28 04388 g003
Figure 4. Effect of the PMS dosage on the RhB degradation efficiency (a) and the corresponding pseudo-first-order kinetic curves (b) (reaction condition: α-MnO2 = 0.2 g/L, RhB = 20 mg/L, initial pH = 7, temperature = 25 ± 2 °C).
Figure 4. Effect of the PMS dosage on the RhB degradation efficiency (a) and the corresponding pseudo-first-order kinetic curves (b) (reaction condition: α-MnO2 = 0.2 g/L, RhB = 20 mg/L, initial pH = 7, temperature = 25 ± 2 °C).
Molecules 28 04388 g004
Figure 5. Effect of the α-MnO2 dosage on the RhB degradation efficiency (a) and the corresponding pseudo-first-order kinetic fitting curves (b) (reaction condition: PMS = 0.02 g/L, RhB = 20 mg/L, initial pH = 7, temperature = 25 ± 2 °C).
Figure 5. Effect of the α-MnO2 dosage on the RhB degradation efficiency (a) and the corresponding pseudo-first-order kinetic fitting curves (b) (reaction condition: PMS = 0.02 g/L, RhB = 20 mg/L, initial pH = 7, temperature = 25 ± 2 °C).
Molecules 28 04388 g005
Figure 6. Influence of the RhB initial concentration (a) and the pseudo-first-order kinetic curves (b) (reaction condition: α-MnO2 = 0.2 g/L, RhB = 20 mg/L, PMS = 0.05 g/L, initial pH = 7, temperature = 25 ± 2 °C).
Figure 6. Influence of the RhB initial concentration (a) and the pseudo-first-order kinetic curves (b) (reaction condition: α-MnO2 = 0.2 g/L, RhB = 20 mg/L, PMS = 0.05 g/L, initial pH = 7, temperature = 25 ± 2 °C).
Molecules 28 04388 g006
Figure 7. Influence of pH value (a) and cation (b) on the RhB degradation efficiency (reaction condition: α-MnO2 = 0.2 g/L, RhB = 20 mg/L, PMS = 0.05 g/L, temperature = 25 ± 2 °C).
Figure 7. Influence of pH value (a) and cation (b) on the RhB degradation efficiency (reaction condition: α-MnO2 = 0.2 g/L, RhB = 20 mg/L, PMS = 0.05 g/L, temperature = 25 ± 2 °C).
Molecules 28 04388 g007
Figure 8. Influence of different radical scavengers (a) and the pseudo-first-order kinetic curves (b) (reaction condition: α-MnO2 = 0.2 g/L, PMS = 0.05 g/L, initial pH = 7, RhB = 20 mg/L, temperature = 25 ± 2 °C).
Figure 8. Influence of different radical scavengers (a) and the pseudo-first-order kinetic curves (b) (reaction condition: α-MnO2 = 0.2 g/L, PMS = 0.05 g/L, initial pH = 7, RhB = 20 mg/L, temperature = 25 ± 2 °C).
Molecules 28 04388 g008
Figure 9. Repeatability of α-MnO2 for 4 cycling tests (reaction condition: α-MnO2 = 0.2 g/L, PMS = 0.05 g/L, initial pH = 7, RhB = 20 mg/L, temperature = 25 ± 2 °C).
Figure 9. Repeatability of α-MnO2 for 4 cycling tests (reaction condition: α-MnO2 = 0.2 g/L, PMS = 0.05 g/L, initial pH = 7, RhB = 20 mg/L, temperature = 25 ± 2 °C).
Molecules 28 04388 g009
Figure 10. UV-vis scanning spectrum during the RhB degradation process (reaction condition: α-MnO2 = 0.2 g/L, PMS = 0.05 g/L, initial pH = 7, RhB = 20 mg/L, temperature = 25 ± 2 °C).
Figure 10. UV-vis scanning spectrum during the RhB degradation process (reaction condition: α-MnO2 = 0.2 g/L, PMS = 0.05 g/L, initial pH = 7, RhB = 20 mg/L, temperature = 25 ± 2 °C).
Molecules 28 04388 g010
Figure 11. Proposed RhB degradation mechanism by α-MnO2 activating PMS.
Figure 11. Proposed RhB degradation mechanism by α-MnO2 activating PMS.
Molecules 28 04388 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

Li, J.; Shi, Q.; Sun, M.; Liu, J.; Zhao, R.; Chen, J.; Wang, X.; Liu, Y.; Gong, W.; Liu, P.; et al. Peroxymonosulfate Activation by Facile Fabrication of α-MnO2 for Rhodamine B Degradation: Reaction Kinetics and Mechanism. Molecules 2023, 28, 4388. https://doi.org/10.3390/molecules28114388

AMA Style

Li J, Shi Q, Sun M, Liu J, Zhao R, Chen J, Wang X, Liu Y, Gong W, Liu P, et al. Peroxymonosulfate Activation by Facile Fabrication of α-MnO2 for Rhodamine B Degradation: Reaction Kinetics and Mechanism. Molecules. 2023; 28(11):4388. https://doi.org/10.3390/molecules28114388

Chicago/Turabian Style

Li, Juexiu, Qixu Shi, Maiqi Sun, Jinming Liu, Rui Zhao, Jianjing Chen, Xiangfei Wang, Yue Liu, Weijin Gong, Panpan Liu, and et al. 2023. "Peroxymonosulfate Activation by Facile Fabrication of α-MnO2 for Rhodamine B Degradation: Reaction Kinetics and Mechanism" Molecules 28, no. 11: 4388. https://doi.org/10.3390/molecules28114388

Article Metrics

Back to TopTop