Next Article in Journal
All-Solid-State Carbon Black Paste Electrodes Modified by Poly(3-octylthiophene-2,5-diyl) and Transition Metal Oxides for Determination of Nitrate Ions
Previous Article in Journal
SERS Detection of the Anti-Epileptic Drug Perampanel in Human Saliva
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Peroxymonosulfate-Activation-Induced Phase Transition of Mn3O4 Nanospheres on Nickel Foam with Enhanced Catalytic Performance

1
School of Materials Science and Hydrogen Energy, Foshan University, Foshan 528000, China
2
Guangdong Key Laboratory for Hydrogen Energy Technologies, Foshan 528000, China
3
Guangdong Provincial Key Laboratory of Distributed Energy Systems, School of Chemical Engineering and Energy Technology, Dongguan University of Technology, Dongguan 523808, China
4
School of Light Industry and Materials, Guangdong Polytechnic, Foshan 528041, China
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(11), 4312; https://doi.org/10.3390/molecules28114312
Submission received: 9 April 2023 / Revised: 13 May 2023 / Accepted: 22 May 2023 / Published: 24 May 2023

Abstract

:
The transformations of physicochemical properties on manganese oxides during peroxymonosulfate (PMS) activation are vital factors to be concerned. In this work, Mn3O4 nanospheres homogeneously loaded on nickel foam are prepared, and the catalytic performance for PMS activation is evaluated by degrading a target pollutant, Acid Orange 7, in aqueous solution. The factors including catalyst loading, nickel foam substrate, and degradation conditions have been investigated. Additionally, the transformations of crystal structure, surface chemistry, and morphology on the catalyst have been explored. The results show that sufficient catalyst loading and the support of nickel foam play significant roles in the catalytic reactivity. A phase transition from spinel Mn3O4 to layered birnessite, accompanied by a morphological change from nanospheres to laminae, is clarified during the PMS activation. The electrochemical analysis reveals that more favorable electronic transfer and ionic diffusion occur after the phase transition so as to enhance catalytic performance. The generated SO4•− and •OH radicals through redox reactions of Mn are demonstrated to account for the pollutant degradation. This work will provide new understandings of PMS activation by manganese oxides with high catalytic activity and reusability.

1. Introduction

Advanced oxidation processes (AOPs) with free radicals have received much attention for the various applications in environmental remediation, including degradation and mineralization of organic molecules [1], detoxification [2], and disintegration of sludge [3]. The sulfate radical (SO4•−) provided tremendous potential as an advanced alternative to hydroxyl radical (•OH) due to its advantages including higher standard redox potential ( E SO 4 = 2.5–3.1 V, E OH = 1.8–2.7 V, vs. NHE) [4,5], better practicability under a wide pH range (pH = 2–8) [6,7], and longer lifespan (30–40 μs for SO4•− vs. 20 ns for •OH) [8,9]. It has been reported that SO4•− could be generated by the activation of peroxydisulfate (PDS) or peroxymonosulfate (PMS) with heating [10], UV irradiation [11], and ultrasound [12]. These physical activation methods require persistent energy inputs, resulting in higher costs for practical applications. Homogeneous transition-metal ions in solution, such as Co2+ [13], Fe2+ [14], and Cu2+ [15], can generate SO4•− through the change in valence state but without extraneous energy inputs. However, the homogeneous transition-metal ions are difficult to reclaim, and may cause secondary pollution. To overcome these problems, various heterogeneous catalysts, particularly transition-metal oxides, have been extensively investigated in PMS and PDS activation.
Among all kinds of transition-metal oxides, manganese oxides are promising catalysts for the effective activation of PMS due to their advantages of diverse valency, high natural abundance, and environmental benignity [16,17]. The reactivity of manganese oxides can be influenced by different structural and surface physicochemical properties. Manganese oxides with different chemical states, including MnO, MnO2, Mn2O3, and Mn3O4, displayed catalytic diversity on PMS activation [18]. In addition, the reactivity of different crystallographic MnO2 had a high correlation with the surface Mn valence state, and other properties, such as surface area, conductivity, or surface adsorbed oxygen, were secondary factors [19]. Therefore, it is essential to explore the diversify the oxidation states of Mn ions in manganese oxides. Thermodynamically stable hausmannite Mn3O4 composed of Mn(II) in tetrahedral sites and Mn(III) in octahedral sites has been reported as a promising catalyst alternative to Co3O4 and Fe3O4 for PMS-based AOPs [20]. However, Mn3O4 generally suffers from poor conductivity and serious agglomeration [21], which would hinder the electron transfer and exposure of active sites.
To tackle with above issues, the construction of novel microstructures for Mn3O4 is one of the strategies, such as hierarchical nanostructures [22] and yolk–shell microparticles [23]. In addition, the employment of matrixes (g-C3N4 [24] and reduced graphene oxide [25]) can effectively improve Mn3O4 spatial configuration. Previous studies have demonstrated nickel foam as a supporter for the uniform distribution of nano/micro materials, such as Co(OH)2 nanosheets [26], CuCoNi oxide nanowires [27], and Cu/CoS nano needles [28], which deliver high catalytic performance. Thus, the employment of nickel foam for the support of Mn3O4 is worthy of study. Moreover, the chemical phases of Mn3O4 are difficult to retain after PMS activation, affecting the reusability [29,30]. It has been reported that the chemical valence states of Mn3O4 are variational during PMS activation [29]. In addition to the valence states, the accompanying changes in crystal structure and morphology have not been considered in the reported literatures. Thus, detailed investigations of the transformations of physicochemical properties on Mn3O4 during PMS activation are vital for approaching high catalytic stability.
Herein, Mn3O4 nanospheres homogeneously loaded on nickel foam were constructed for PMS activation to degrade Acid Orange 7 (AO7, a typical dye). The influence of catalyst loading, nickel foam substrate, and some conditions, including PMS concentration, pH value, and reaction temperature, were investigated in detail. The cyclic performance accompanied by the changes in the microstructure and surface chemistry of Mn3O4 during PMS activation was explored. A solid-phase transition from spinel Mn3O4 to birnessite induced by PMS activation was clarified. Improved electronic transfer and ionic diffusion of the catalyst after repeated use were revealed, which were responsible for the promotion of the degradation rate. A mechanism involved in the solid-phase transition of Mn3O4 and the generation of SO4•− and •OH radicals was proposed.

2. Results and Discussion

2.1. Characterization of the As-Prepared Mn3O4 Sample

The XRD pattern of the as-prepared Mn3O4 sample scraped from nickel foam is shown in Figure 1a. All the diffraction peaks are well indexed to hausmannite Mn3O4 (JCPDS 24-0734), the most prominent of which at 2θ values of 32.7° and 36.2° can be attributed to the (103) and (211) planes of the tetragonal spinel structure. Raman and FT-IR measurements were carried out to analyze the phase purity of the sample. As shown in Figure 1b, the strong Raman band located at 655 cm−1 is characteristic of hausmannite Mn3O4, which arose due to the breathing mode of Mn-O vibration from Mn2+ ions on the tetrahedral site [31]. Two main absorption bands located at 630 and 511 cm−1 are observed from the FT-IR spectrum shown in Figure S1. This result agrees with the previously reported Mn3O4, which is attributed to Mn-O stretching vibrations on octahedral and tetrahedral sites [22]. The broad absorption band at around 3400 cm−1 could be assigned to the stretching vibration of the adsorbed hydroxyl group. These spectroscopic results further confirm the Mn3O4 phase.
The SEM images of the pristine nickel foam and the as-prepared Mn3O4 sample are present. Figure 2a,b display the 3D porous structure of the nickel foam and its rather smooth surface. After the hydrothermal reaction, the evenly coated Mn3O4 nanospheres roughen the nickel foam surface (Figure 2c,d). The nanospheres stacked on top of each other forming an irregular thickness (Figure 2e). The enlarged SEM image in Figure 2f reveals the irregularly spherical shape of Mn3O4 nanospheres with a diameter of about 100–200 nm. The EDS results show the homogeneous distribution of Ni, Mn, and O elements on the sample (Figure 2g). The TEM image exhibits Mn3O4 nanospheres formed by the aggregation of small nanoparticles (Figure 2h). The nanoparticles are approximately 12 nm in size. The high-resolution TEM (HRTEM) image in Figure 2i displays lattice fringes with interplanar distances of 0.49 and 0.27 nm, which correspond to the crystal planes of (101) and (103) of hausmannite Mn3O4, respectively. The selected area electron diffraction (SAED) pattern obtained from Figure 2j demonstrates a polycrystalline feature and can be well indexed to hausmannite Mn3O4.

2.2. Degradation Performances in Different Systems

2.2.1. Effect of Catalyst Loading

The samples with different catalyst loading were prepared using different dosages of manganese salt (Mn(CH3COO)2∙4H2O). The prepared samples were denoted as Mn3O4/NF-M (M means the dosage of manganese salt). The mass loadings of Mn3O4/NF-0.12, Mn3O4/NF-0.24, and Mn3O4/NF-0.48 are about 0.3, 0.6, and 0.9 mg cm−2, respectively. The SEM images of Mn3O4/NF-0.24 and Mn3O4/NF-0.12 are present in Figure S2. It can be observed that the Mn3O4 nanospheres aggregated by nanoparticles were uniformly distributed on the nickel foam surface. In contrast with the morphology of Mn3O4/NF-0.48, the Mn3O4 nanospheres exhibited decreasingly smaller dimensions with lower dosages of manganese salt. Figure 3a shows the AO7 degradation of different samples under the same conditions. The Mn3O4/NF-0.48 presented a better PMS activation performance than Mn3O4/NF-0.24 and Mn3O4/NF-0.12. A degradation efficiency of 94% could be achieved for Mn3O4/NF-0.48 within 30 min, which was 83.5% and 63% for Mn3O4/NF-0.24 and Mn3O4/NF-0.12, respectively. The results of AO7 degradation with the Mn3O4/NF system could be fitted to pseudo-first-order kinetics, as shown in Figure 3b. The degradation rate constant (k) of Mn3O4/NF-0.48 was estimated to be 0.095 min−1, higher than those of Mn3O4/NF-0.24 (0.060 min−1) and Mn3O4/NF-0.12 (0.033 min−1). The better degradation performance of Mn3O4/NF-0.48 could be attributed to its raised catalyst loading.

2.2.2. Effect of Nickel Foam

The effect of nickel foam was evaluated by the comparison of the performance using blank nickel foam, Mn3O4 powder, and Mn3O4/NF-0.48. Figure 3c displays AO7 degradation under different systems. A low degradation efficiency of about 6% was evaluated after 60 min with the presence of PMS or both PMS and blank nickel foam, suggesting that the blank nickel foam shows negligible effectiveness on the removal of AO7. The degradation efficiency of Mn3O4 powder with PMS (87%) is lower than that of Mn3O4/NF-0.48 with PMS (98%) after 60 min. Furthermore, no adsorption of AO7 on Mn3O4/NF-0.48 was detected (Figure S3) within 30 min, confirming the contribution of nickel foam to the catalytic activity of Mn3O4. Figure 3d shows the corresponding UV-vis spectra of the reaction system as a function of degradation time. The consecutive reduction in the absorbance at around 484 nm proved the destruction of azo chromophore.

2.2.3. Effects of Degradation Conditions

The effects of PMS concentration, initial pH value, and reaction temperature have been investigated in the Mn3O4/NF-0.48 system. As shown in Figure 4a, PMS concentration promotes the degradation efficiency significantly in the range of 0.2 to 1 mM oxone, which could be ascribed to the involvement of more reactive species activated from PMS. Once the concentration of PMS is higher than 1 mM, the positive effect on degradation tends to be slight. In the current reaction system, the PMS dosage is in plenty relative to the catalyst loaded on nickel foam. In consideration of the degradation efficiency and operation cost, 1 mM of PMS concentration can be more adequate for AO7 degradation.
The influence of the initial pH value on the degradation of AO7 was examined. As shown in Figure 4b, a weakening trend of degradation performance with an increasing initial pH value can be observed. It is reported that manganese oxides are generally negatively charged under neutral and basic conditions [30,32]. The electrostatic repulsion between the catalyst and PMS anion or anionic dye AO7 might be one of the reasons. In addition, the deprotonation of HSO5 into SO42− and O2 under basic conditions would cause invalid PMS consumption [33]. Nonetheless, the system of Mn3O4/NF-0.48 with PMS shows efficiency in the pH range from 3 to 9, since the degradation efficiencies remain above 90%.
The effect of the reaction temperature on AO7 degradation is displayed in Figure 4c, showing an accelerated degradation at higher temperature. The degradation rate constant at 30, 40, 50, and 60 °C can be deduced to be 0.085, 0.112, 0.139, and 0.178 min−1. The relationship between the rate constants and reaction temperatures obeys the Arrhenius equation closely as shown in Figure 4d. The activation energy (Ea) is estimated to be 20.52 kJ/mol, which is at a lower level than those of AO7 degradation over some reported catalysts, for example, Mn3O4-rGO (49.5 kJ/mol) [25], MnFe2O4 (31.7 kJ/mol) [34], Co/Bi25FeO40 (51.3 kJ/mol) [35], and Co3O4/N-doped graphene (41.6 kJ/mol) [36]. It suggests that the Mn3O4/NF exhibits a low chemical reaction energy barrier as a promising catalyst for PMS activation. The influence of the stirring rate on AO7 degradation was explored. As shown in Figure S4, a degradation rate constant (k) of 0.094 min−1 could be achieved under the stirring rate of 200 rpm, which was 0.024 and 0.04 min−1 under the stirring rates of 50 and 400 rpm, respectively. Thus, a stirring rate of 200 rpm is suitable for mass transfer.

2.3. Catalyst Reusability and Structural Change

Figure 5a displays the recycling performance of AO7 degradation with Mn3O4/NF-0.48. As the results show, the catalyst exhibited progressively enhanced degradation performance within the four repeated cycles. The AO7 degradation rate escalated from 0.086 to 0.105 min−1 after four runs (Figure 5b). These results are noteworthy as the heterogeneous catalysts usually suffer from the problem of stability.
The catalyst after four cycles of repetitive usages was further investigated to make clear the reason for enhanced degradation performance. The XRD pattern in Figure 6a demonstrates a set of characteristic peaks that can be indexed to the standard contour of birnessite MnO2 (JCPDS no. 18-0802). The diffraction peaks at 2θ values of 18.7°, 36.8°, and 65.7° can be attributed to the (101), (006), and (119) planes of the layered structure. These weak diffraction peaks indicate the low crystallinity of the catalyst. The broad peak at the range of 20° to 30° may be ascribed to the adsorbed organic intermediates on the catalyst surface [37]. Figure 6b displays deconvoluted Raman bands at 508, 578, 652, and 720 cm−1, which could be assigned to the vibrational features of birnessite manganese oxide [38]. Additionally, the band at 620 cm−1 may be raised due to the presence of MnOOH on the catalyst surface [39]. These results attest to a phase transition from spinel Mn3O4 to birnessite manganese oxide after PMS activation on AO7 degradation.
XPS was employed to analyze the evolution of the surface chemistry on the catalyst sample. From the Mn 3s core-level XPS spectra (Figure 7a), the energy separation of the two peaks was measured to be 5.70, 5.06, and 4.58 eV for pristine Mn3O4/NF-0.48 and the samples after one cycle and four cycles, respectively. This value has been proved to be linearly corelated to the mean valence state of Mn [40], which then can be estimated to be +2.50, +3.23, and +3.77 for the three samples, respectively. This verifies the increasing trend of the mixed Mn valence states during the repeated PMS activation on AO7 degradation. The shifts of the broad Mn 2p3/2 peak toward higher binding energy of the samples after one cycle and four cycles (Figure 7b) further confirm the oxidation of Mn. The O 1s spectra (Figure 7c) are deconvoluted into three bands which, respectively, correspond to the Mn-O-Mn bond (529.6 ± 0.1 eV) for anhydrous Mn oxide, Mn-O-H bond (531.1 ± 0.2 eV) for Mn hydroxide, and H-O-H bond (532.5 ± 0.2 eV) for structure water. Increasing content of the surface hydroxyl group and structure water were detected during the repeated cycles. In particular, more than twice the content of the structure water on the surface was obtained after four cycles of degradation.
The surface morphologies of the catalyst samples collected after the first and fourth cycles were observed using SEM. As shown in Figure 8a–c, the pristine Mn3O4 nanospheres in situ grew into a lamella structure. With an increasing cycle number, the lamellae exhibited a larger lateral diameter, and interconnected with each other assembling porous architecture (Figure 8d–f). The corresponding TEM image (Figure 8g) displays the interconnected lamellae, which are 5–20 nm in thickness and 50–200 nm in lateral size. The HRTEM image (Figure 8h) shows the typical lattice fringes with spacings of 0.14, 0.24, and 0.46 nm, which correspond to the (119), (006), and (101) planes of birnessite MnO2, respectively. The SAED pattern (Figure 8i) further confirms the poor crystallinity of the birnessite structure, which agrees well with the XRD result.
Based on the abovementioned results, it can be seen that Mn3O4 catalyst samples suffer a chemical oxidation process after AO7 degradation with PMS activation, leading to a solid-phase transition from spinel to birnessite and morphological evolution from nanospheres to a lamella structure. The enhanced degradation performance of the catalyst with repeated use is most probably associated with the phase and morphology changes. The properties of electron transport and ion diffusion resistivity of the catalysts are analyzed using electrochemical impedance spectroscopy (EIS). Nyquist plots and the fitting lines of the Mn3O4/NF-0.48 and the sample after four cycles of AO7 degradation are compared in Figure 9a. The approximate semicircle at high frequency represents a charge-transfer-controlled process. The series resistance (Rs) and charge-transfer resistance (Rct) are fitted to be 1.906 Ω and 9.858 Ω for Mn3O4/NF-0.48, while they change to be 1.797 Ω and 0.01 Ω after four cycles. The phase transition introduces minimal change to the intrinsic resistance of the catalyst but a great extent of reduction on the Rct. This indicates better efficient electron transfer at the interface, which could promote the reaction between the catalyst and PMS. The straight line at low frequency represents an ion-diffusion-controlled process. The higher slop of the line implies a more favorable ionic diffusion of the catalyst after four cycles. The real part of impedance (Z’) is plotted versus the reciprocal of the square root of frequency (ω−0.5) in the intermediate frequency range, which can derive the ion diffusion resistance (σ) through the slope of linear fitting (Figure 9b). The σ value is found to decrease considerably from 9.418 Ω/s−0.5 to 0.253 Ω/s−0.5 after four cycles. The decrease in σ may be ascribed to the lamella structure, which offers open channels for facile ion diffusion [22]. This allows more active surface sites for the adsorption and reactions of the reactants, improving the catalyst activity.

2.4. Identification of Radicals

It has been reported that the organic pollutant degradation with PMS usually follows two different mechanisms, including radical and nonradical pathways. EPR was employed to identify the active species for catalysis. The radical signals captured by DMPO are shown in Figure 10. Compared with the weak signals without catalyst addition, high intensities of DMPO-•OH and DMPO- SO4•− signals could be observed in the Mn3O4/NF-0.48/PMS system. Both signals exhibit enhanced intensities with the increase in reaction time from 5 min to 10 min. It suggests that both •OH and SO4•− could be derived via catalysis during PMS activation and would be accumulated within the reaction time. In addition, the nonradical 1O2 signal captured using TEMP is present in Figure S5. The TEMP-1O2 signal in the Mn3O4/NF-0.48/PMS system showed roughly the same intensity as that of the raw PMS system without catalyst addition. It indicates that nonradical 1O2 formed through spontaneous PMS decomposition, which presented little contribution to the AO7 degradation based on the above results under the raw PMS condition. Therefore, radical •OH and SO4•− were the contributing species for the AO7 degradation in the Mn3O4/NF-0.48/PMS system.

2.5. The Mechanism of PMS Activation and Phase Transition

Based on above results, the phase transition induced by PMS activation was schematically proposed in Figure 11. Firstly, Mn(II) can produce SO4•− by reacting with HSO5 (Equation (1)). The oxidation reaction of Mn(II) may induce its extraction from tetrahedral sites, which is similar to the dissolution of Mn(II) form Mn3O4 matrix under electrochemical oxidation conditions in aqueous solution [41,42]. Meanwhile, Mn(III) at the octahedral sites can be oxidized to Mn(IV), while one fourth of octahedron Mn cations migrate to the (101) plane of spinel, leading to the in situ formation of a layered structure [43]. Additionally, Mn at the higher valence states can then be reduced by HSO5 to generate SO4•− (Equations (2) and (3)). The generated SO4•− can be readily converted to •OH via the oxidation of water (Equation (4)). Based on the surface chemistry of the catalyst after repeated cycles, the extracted Mn(II) and free H3O+ may fill into the MnO6 layer gap to recover the charge balance after the ion extraction and stabilize the structure [44]. These atomic movements during the solid-phase transition produce a large strain between the phase edges, which could cause the grown birnessite lamella to peel off from the Mn3O4 surface [45], leading to morphological change.
HSO 5 + Mn ( II )     SO 4 + OH + Mn ( III )
HSO 5 + Mn ( III / IV )     SO 5 + H + + Mn ( II / III )
2 SO 5     2 SO 4 + O 2
SO 4 + H 2 O     HSO 4 + OH

3. Materials and Methods

3.1. Synthesis of Mn3O4 Nanospheres on Nickel Foam

Mn3O4 nanospheres were fabricated on nickel foam via a facile hydrothermal route, which was modified based on a reported recipe [46]. In brief, a precursor solution was firstly prepared through the dissolution of 0.48 g Mn(CH3COO)2∙4H2O into 3 g ethanol with 1 g deionized water under vigorous stirring for 15 min, followed by the addition of 13 mL ethylene glycol with continuous stirring for 30 min. A piece of pretreated nickel foam (~320 g m−2, 4 cm × 3.5 cm × 0.1 cm) was dipped into the above precursor solution. The pretreatment of nickel foam was conducted via drying after successive ultrasonic cleaning in acetone, ethanol, and deionized water, respectively. The solution and the nickel foam were statically aged for 3 days and then transferred into an autoclave for a hydrothermal reaction at 170 °C for 5 h. Finally, the nickel foam sample was removed and rinsed with deionized water and ethanol, and dried in air at 50 °C. For comparison, blank nickel foam and Mn3O4 powder were prepared using the same route but without the addition of manganese salt or nickel foam, respectively.

3.2. Characterization

The X-ray diffraction (XRD, SmartLab, Rigaku, Tokyo, Japan) was conducted using Cu-Kα radiation (λ = 1.5406 Å). Raman scattering was recorded using a LabRAM HR Evolution Raman spectrometer with a 532 nm laser (Horiba Jobin-Yvon, Villeneuve d’Ascq, France). Fourier transform infrared (FT-IR) was performed using a Fourier transform infrared spectrometer (Shimadzu IRAffinity-1S, Kyoto, Japan). The morphology was observed via field emission scanning electron microscopy (FESEM, Hitachi S4800, Kyoto, Japan) and field-emission transmission electron microscopy (TEM, FEI Tecnai TF20, Hillsboro, OR, USA). The elemental energy dispersive spectroscopy (EDS) analysis was conducted using an Oxford EDS detector. X-ray photoelectron spectroscopy (XPS) was performed using a monochromatic Al Kα X-ray source with photon energy of 1486.6 eV (Thermo Scientific K-Alpha, ThermoFisher, Waltham, MA, USA). Electrochemical impedance spectroscopy (EIS) was performed on a CHI760E electrochemical workstation (Chenhua, China) with a Pt wire counter electrode and an Ag/AgCl (3 M KCl) reference electrode. An electron paramagnetic resonance spectrometer (EPR, JEOL JES FA200, Tokyo, Japan) was used to identify the radical species with DMPO and TEMP spin-trapping agents.

3.3. Evaluation of Catalytic Activity

The catalytic activity was evaluated using the degradation of AO7 under a constant stirring of 200 rpm with the nickel foam sample at room temperature. In a typical procedure, the nickel foam sample was firstly immersed into 100 mL 20 mg L−1 of AO7 solution in a 250 mL beaker. After constant stirring for 30 min, 1 mmol/L oxone (2KHSO5•KHSO4•K2SO4) was added to initiate the reaction. At given time intervals, 2 mL solution samples were collected, followed by the addition of 2 mL ethanol for quenching of the reaction. The concentration of AO7 was detected using a UV-vis spectrophotometer (Shimadzu UV-1800, Japan) at 485 nm. For comparison, the catalytic activity of blank nickel foam and Mn3O4 powder was evaluated through the same route. The effect of pH was investigated using 1 M HCl or 1 M NaOH. The stability of the catalyst was evaluated through reutilization of the nickel foam sample after washing and drying within four successive degradation cycles.

4. Conclusions

In summary, Mn3O4 nanospheres homogeneously loaded on nickel foam are prepared using a simple hydrothermal route for PMS activation to degrade AO7. The Mn3O4/NF/PMS system displays favorable catalytic activity and an enhanced degradation rate with repeated usage. A transition from spinel nanospheres to birnessite laminae driven by PMS is revealed, which can lead to more favorable electronic transfer and ionic diffusion. The phase transition is proposed to be comprised of Mn extraction, rearrangement, and the insertion of cations between layers, accompanied by redox reactions of Mn to generate active radicals of SO4•− and •OH. This work will provide new insights into PMS activation by manganese oxides.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/molecules28114312/s1, Figure S1: FT-IR spectrum of the as-prepared Mn3O4 sample; Figure S2: SEM images of (a1–a3) Mn3O4/NF-0.12 and (b1–b3) Mn3O4/NF-0.24; Figure S3: The adsorption of AO7 on Mn3O4/NF-0.48; Figure S4: AO7 degradation efficiency (a) and kinetic curves (b) of Mn3O4/NF-0.48/PMS system under different stirring rates. Conditions: [AO7]0 = 20 mg/L, [oxone]0 = 1 mM; Figure S5: EPR spectra of Mn3O4/NF-0.48/PMS system captured by TEMP.

Author Contributions

Conceptualization, C.L. and J.L.; Formal analysis, Y.C.; Investigation, Z.W., H.Z. and J.D.; Methodology, J.L.; Supervision, H.R.; Validation, Y.C.; Visualization, X.Z.; Writing—original draft, Z.W.; Writing—review and editing, C.L. and H.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the Guangdong Basic and Applied Basic Research Foundation (Nos. 2019A1515110535, 2019A1515110098, and 2019A1515110528), National Natural Science Foundation of China (No. 22208051), Innovation Team of Universities of Guangdong Province (2022KCXTD030 and 2020KCXTD011), Engineering Research Center of Universities of Guangdong Province (2019GCZX002), and Guangdong Key Laboratory for Hydrogen Energy Technologies (2018B030322005).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are available with requirements.

Acknowledgments

The authors would like to thank Shiyanjia Lab (www.shiyanjia.com, accessed on 29 May 2022) for the XPS analysis.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not applicable.

References

  1. Lu, C.; Zhang, S.; Wang, J.; Zhao, X.; Zhang, L.; Tang, A.; Dong, X.; Fu, L.; Yang, H. Efficient activation of peroxymonosulfate by iron-containing mesoporous silica catalysts derived from iron tailings for degradation of organic pollutants. Chem. Eng. J. 2022, 446, 137044. [Google Scholar] [CrossRef]
  2. Babu, D.S.; Srivastava, V.; Nidheesh, P.V.; Kumar, M.S. Detoxification of water and wastewater by advanced oxidation processes. Sci. Total Environ. 2019, 696, 133961. [Google Scholar] [CrossRef]
  3. Lin, W.; Liu, X.; Ding, A.; Ngo, H.H.; Zhang, R.; Nan, J.; Ma, J.; Li, G. Advanced oxidation processes (AOPs)-based sludge conditioning for enhanced sludge dewatering and micropollutants removal: A critical review. J. Water Process Eng. 2022, 45, 102468. [Google Scholar] [CrossRef]
  4. Buxton, G.V.; Greenstock, C.L.; Helman, W.P.; Ross, A.B. Critical Review of rate constants for reactions of hydrated electrons, hydrogen atoms and hydroxyl radicals (OH/O) in Aqueous Solution. J. Phys. Chem. Ref. Data 1988, 17, 513–886. [Google Scholar] [CrossRef]
  5. Neta, P.; Huie, R.E.; Ross, A.B. Rate Constants for Reactions of Inorganic Radicals in Aqueous Solution. J. Phys. Chem. Ref. Data 1988, 17, 1027–1284. [Google Scholar] [CrossRef]
  6. Li, Y.; Zhu, W.; Guo, Q.; Wang, X.; Zhang, L.; Gao, X.; Luo, Y. Highly efficient degradation of sulfamethoxazole (SMX) by activating peroxymonosulfate (PMS) with CoFe2O4 in a wide pH range. Sep. Purif. Technol. 2021, 276, 119403. [Google Scholar] [CrossRef]
  7. Xu, X.; Zhan, F.; Pan, J.; Zhou, L.; Su, L.; Cen, W.; Li, W.; Tian, C. Engineering single-atom Fe-Pyridine N4 sites to boost peroxymonosulfate activation for antibiotic degradation in a wide pH range. Chemosphere 2022, 294, 133735. [Google Scholar] [CrossRef]
  8. Janzen, E.G.; Kotake, Y.; Randall, D.H. Stabilities of hydroxyl radical spin adducts of PBN-type spin traps. Free. Radic. Biol. Med. 1992, 12, 169–173. [Google Scholar] [CrossRef]
  9. 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]
  10. Huang, D.; Chen, N.; Zhu, C.; Fang, G.; Zhou, D. The overlooked oxidative dissolution of silver sulfide nanoparticles by thermal activation of persulfate: Processes, mechanisms, and influencing factors. Sci. Total Environ. 2021, 760, 144504. [Google Scholar] [CrossRef]
  11. Xie, R.; Ji, J.; Guo, K.; Lei, D.; Fan, Q.; Leung, D.Y.C.; Huang, H. Wet scrubber coupled with UV/PMS process for efficient removal of gaseous VOCs: Roles of sulfate and hydroxyl radicals. Chem. Eng. J. 2019, 356, 632–640. [Google Scholar] [CrossRef]
  12. Deng, D.; Lin, X.; Ou, J.; Wang, Z.; Li, S.; Deng, M.; Shu, Y. Efficient chemical oxidation of high levels of soil-sorbed phenanthrene by ultrasound induced, thermally activated persulfate. Chem. Eng. J. 2015, 265, 176–183. [Google Scholar] [CrossRef]
  13. Chen, M.; Zhu, L.; Liu, S.; Li, R.; Wang, N.; Tang, H. Efficient degradation of organic pollutants by low-level Co2+ catalyzed homogeneous activation of peroxymonosulfate. J. Hazard. Mater. 2019, 371, 456–462. [Google Scholar] [CrossRef] [PubMed]
  14. Yang, Q.; Yan, Y.; Yang, X.; Liao, G.; He, J.; Wang, D. The effect of complexation with metal ions on tetracycline degradation by Fe2+/3+ and Ru3+ activated peroxymonosulfate. Chem. Eng. J. 2022, 429, 132178. [Google Scholar] [CrossRef]
  15. Bouzayani, B.; Rosales, E.; Pazos, M.; Elaoud, S.C.; Sanromán, M.A. Homogeneous and heterogeneous peroxymonosulfate activation by transition metals for the degradation of industrial leather dye. J. Clean. Prod. 2019, 228, 222–230. [Google Scholar] [CrossRef]
  16. Zhou, X.; Shen, S.; Wang, P.; Wei, X.; Zhang, R.; Tang, H.; Wang, H.; Lu, W.; Wang, J. Research progress on catalytic activation of peroxymonosulfate based on manganese oxides. J. Environ. Chem. Eng. 2022, 10, 108937. [Google Scholar] [CrossRef]
  17. Jia, Z.; Lv, Y.; Guo, Y.; Gao, Q.; Lan, H.; Khan, A.; Xu, A.; Li, X. Synergistic effects between graphite nanosheets and OMS-2 to enhance the direct electron transfer process-based peroxymonosulfate activation for pollutants degradation. Appl. Surf. Sci. 2022, 590, 153065. [Google Scholar] [CrossRef]
  18. Saputra, E.; Muhammad, S.; Sun, H.; Ang, H.-M.; Tadé, M.O.; Wang, S. Manganese oxides at different oxidation states for heterogeneous activation of peroxymonosulfate for phenol degradation in aqueous solutions. Appl. Catal. B 2013, 142–143, 729–735. [Google Scholar] [CrossRef]
  19. Huang, J.; Zhong, S.; Dai, Y.; Liu, C.-C.; Zhang, H. Effect of MnO2 Phase Structure on the Oxidative Reactivity toward Bisphenol A Degradation. Environ. Sci. Technol. 2018, 52, 11309–11318. [Google Scholar] [CrossRef]
  20. Saputra, E.; Muhammad, S.; Sun, H.; Ang, H.-M.; Tadé, M.O.; Wang, S. A comparative study of spinel structured Mn3O4, Co3O4 and Fe3O4 nanoparticles in catalytic oxidation of phenolic contaminants in aqueous solutions. J. Colloid Interface Sci. 2013, 407, 467–473. [Google Scholar] [CrossRef]
  21. Mu, J.; Gu, Z.; Sun, H.; Wei, Q. Low Temperature Synthesis of Mn3O4 Nanoparticles Using Starch as Capping Agent. J. Dispers. Sci. Technol. 2006, 27, 307–309. [Google Scholar] [CrossRef]
  22. Wang, Q.; Li, Y.; Shen, Z.; Liu, X.; Jiang, C. Facile synthesis of three-dimensional Mn3O4 hierarchical microstructures for efficient catalytic phenol oxidation with peroxymonosulfate. Appl. Surf. Sci. 2019, 495, 143568. [Google Scholar] [CrossRef]
  23. Zhang, L.; Tong, T.; Wang, N.; Ma, W.; Sun, B.; Chu, J.; Lin, K.A.; Du, Y. Facile Synthesis of Yolk–Shell Mn3O4 Microspheres as a High-Performance Peroxymonosulfate Activator for Bisphenol A Degradation. Ind. Eng. Chem. Res. 2019, 58, 21304–21311. [Google Scholar] [CrossRef]
  24. Chen, C.; Xie, M.; Kong, L.; Lu, W.; Feng, Z.; Zhan, J. Mn3O4 nanodots loaded g-C3N4 nanosheets for catalytic membrane degradation of organic contaminants. J. Hazard. Mater. 2020, 390, 122146. [Google Scholar] [CrossRef] [PubMed]
  25. Yao, Y.; Xu, C.; Yu, S.; Zhang, D.; Wang, S. Facile Synthesis of Mn3O4-Reduced Graphene Oxide Hybrids for Catalytic Decomposition of Aqueous Organics. Ind. Eng. Chem. Res. 2013, 52, 3637–3645. [Google Scholar] [CrossRef]
  26. Yuan, R.; Jiang, M.; Gao, S.; Wang, Z.; Wang, H.; Boczkaj, G.; Liu, Z.; Ma, J.; Li, Z. 3D mesoporous α-Co(OH)2 nanosheets electrodeposited on nickel foam: A new generation of macroscopic cobalt-based hybrid for peroxymonosulfate activation. Chem. Eng. J. 2020, 380, 122447. [Google Scholar] [CrossRef]
  27. Wang, M.-M.; Liu, L.-J.; Wen, J.-T.; Ding, Y.; Xi, J.-R.; Li, J.-C.; Lu, F.-Z.; Wang, W.-K.; Xu, J. Multimetallic CuCoNi Oxide Nanowires In Situ Grown on a Nickel Foam Substrate Catalyze Persulfate Activation via Mediating Electron Transfer. Environ. Sci. Technol. 2022, 56, 12613–12624. [Google Scholar] [CrossRef] [PubMed]
  28. Chen, H.-H.; Park, Y.-K.; Kwon, E.; Fai Tsang, Y.; Xuan Thanh, B.; Cong Khiem, T.; You, S.; Hu, C.; Lin, K.-Y.A. Nanoneedle-Assembled Copper/Cobalt sulfides on nickel foam as an enhanced 3D hierarchical catalyst to activate monopersulfate for Rhodamine b degradation. J. Colloid Interface Sci. 2022, 613, 168–181. [Google Scholar] [CrossRef]
  29. Wang, F.; Xiao, M.; Ma, X.; Wu, S.; Ge, M.; Yu, X. Insights into the transformations of Mn species for peroxymonosulfate activation by tuning the Mn3O4 shapes. Chem. Eng. J. 2021, 404, 127097. [Google Scholar] [CrossRef]
  30. Liu, Y.; Luo, J.; Tang, L.; Feng, C.; Wang, J.; Deng, Y.; Liu, H.; Yu, J.; Feng, H.; Wang, J. Origin of the Enhanced Reusability and Electron Transfer of the Carbon-Coated Mn3O4 Nanocube for Persulfate Activation. ACS Catal. 2020, 10, 14857–14870. [Google Scholar] [CrossRef]
  31. Julien, C.M.; Massot, M.; Poinsignon, C. Lattice vibrations of manganese oxides: Part I. Periodic structures. Spectrochim. Acta Part A 2004, 60, 689–700. [Google Scholar] [CrossRef] [PubMed]
  32. Zhu, S.; Xiao, P.; Wang, X.; Liu, Y.; Yi, X.; Zhou, H. Efficient peroxymonosulfate (PMS) activation by visible-light-driven formation of polymorphic amorphous manganese oxides. J. Hazard. Mater. 2022, 427, 127938. [Google Scholar] [CrossRef]
  33. Lin, K.-Y.A.; Zhang, Z.-Y. Degradation of Bisphenol A using peroxymonosulfate activated by one-step prepared sulfur-doped carbon nitride as a metal-free heterogeneous catalyst. Chem. Eng. J. 2017, 313, 1320–1327. [Google Scholar] [CrossRef]
  34. Yao, Y.; Cai, Y.; Lu, F.; Wei, F.; Wang, X.; Wang, S. Magnetic recoverable MnFe2O4 and MnFe2O4-graphene hybrid as heterogeneous catalysts of peroxymonosulfate activation for efficient degradation of aqueous organic pollutants. J. Hazard. Mater. 2014, 270, 61–70. [Google Scholar] [CrossRef] [PubMed]
  35. Li, W.; Zhang, Y.; Liu, Y.; Cheng, X.; Tang, W.; Zhao, C.; Guo, H. Kinetic performance of peroxymonosulfate activated by Co/Bi25FeO40: Radical and non-radical mechanism. J. Taiwan Inst. Chem. Eng. 2019, 100, 56–64. [Google Scholar] [CrossRef]
  36. Yuan, R.; Hu, L.; Yu, P.; Wang, Z.; Wang, H.; Fang, J. Co3O4 nanocrystals/3D nitrogen-doped graphene aerogel: A synergistic hybrid for peroxymonosulfate activation toward the degradation of organic pollutants. Chemosphere 2018, 210, 877–888. [Google Scholar] [CrossRef]
  37. Xu, Y.; Lin, H.; Li, Y.; Zhang, H. The mechanism and efficiency of MnO2 activated persulfate process coupled with electrolysis. Sci. Total Environ. 2017, 609, 644–654. [Google Scholar] [CrossRef] [PubMed]
  38. Post, J.E.; McKeown, D.A.; Heaney, P.J. Raman spectroscopy study of manganese oxides: Layer structures. Am. Mineral. 2021, 106, 351–366. [Google Scholar] [CrossRef]
  39. Bernard, M.C.; Hugot-Le Goff, A.; Thi, B.V.; de Torresi, S.C. Electrochromic reactions in manganese oxides: I. Raman analysis. J. Electrochem. Soc. 1993, 140, 3065–3070. [Google Scholar] [CrossRef]
  40. Nagaraju, G.; Ko, Y.H.; Cha, S.M.; Im, S.H.; Yu, J.S. A facile one-step approach to hierarchically assembled core–shell-like MnO2@MnO2 nanoarchitectures on carbon fibers: An efficient and flexible electrode material to enhance energy storage. Nano Res. 2016, 9, 1507–1522. [Google Scholar] [CrossRef]
  41. Jabeen, N.; Hussain, A.; Xia, Q.; Sun, S.; Zhu, J.; Xia, H. High-performance 2.6 V aqueous asymmetric supercapacitors based on in situ formed Na0.5MnO2 nanosheet assembled nanowall arrays. Adv. Mater. 2017, 29, 1700804. [Google Scholar] [CrossRef]
  42. Zhao, N.; Fan, H.; Zhang, M.; Ma, J.; Wang, C.; Yadav, A.K.; Li, H.; Jiang, X.; Cao, X. Beyond intercalation-based supercapacitors: The electrochemical oxidation from Mn3O4 to Li4Mn5O12 in Li2SO4 electrolyte. Nano Energy 2020, 71, 104626. [Google Scholar] [CrossRef]
  43. Li, Y.-F.; Liu, Z.-P. Active site revealed for water oxidation on electrochemically induced δ-MnO2: Role of spinel-to-layer phase transition. J. Am. Chem. Soc. 2018, 140, 1783–1792. [Google Scholar] [CrossRef]
  44. Yang, L.; Cheng, S.; Ji, X.; Jiang, Y.; Zhou, J.; Liu, M. Investigations into the origin of pseudocapacitive behavior of Mn3O4 electrodes using in operando Raman spectroscopy. J. Mater. Chem. A 2015, 3, 7338–7344. [Google Scholar] [CrossRef]
  45. Kim, S.; Nam, K.W.; Lee, S.; Cho, W.; Kim, J.-S.; Kim, B.G.; Oshima, Y.; Kim, J.-S.; Doo, S.-G.; Chang, H.; et al. Direct observation of an anomalous spinel-to-layered phase transition mediated by crystal water intercalation. Angew. Chem. Int. Ed. 2015, 54, 15094–15099. [Google Scholar] [CrossRef]
  46. Sun, Z.; Liao, T.; Dou, Y.; Hwang, S.M.; Park, M.-S.; Jiang, L.; Kim, J.H.; Dou, S.X. Generalized self-assembly of scalable two-dimensional transition metal oxide nanosheets. Nat. Commun. 2014, 5, 3813. [Google Scholar] [CrossRef] [PubMed]
Figure 1. As-prepared Mn3O4 sample: (a) XRD pattern; (b) Raman spectrum.
Figure 1. As-prepared Mn3O4 sample: (a) XRD pattern; (b) Raman spectrum.
Molecules 28 04312 g001
Figure 2. (a,b) SEM images of the pristine nickel foam. As-prepared Mn3O4 sample: (cf) SEM images; (g) EDS element mappings; (h,i) TEM images; (j) SAED pattern.
Figure 2. (a,b) SEM images of the pristine nickel foam. As-prepared Mn3O4 sample: (cf) SEM images; (g) EDS element mappings; (h,i) TEM images; (j) SAED pattern.
Molecules 28 04312 g002
Figure 3. AO7 degradation efficiency (a) and kinetic curves (b) of Mn3O4/NF-0.12, Mn3O4/NF-0.24, and Mn3O4/NF-0.48. (c) AO7 degradation efficiency of Mn3O4/NF-0.48 under different systems. (d) UV-vis spectra of AO7 solution under given time intervals in the presences of Mn3O4/NF-0.48 and PMS. Conditions: [AO7]0 = 20 mg/L, [oxone]0 = 1 mM.
Figure 3. AO7 degradation efficiency (a) and kinetic curves (b) of Mn3O4/NF-0.12, Mn3O4/NF-0.24, and Mn3O4/NF-0.48. (c) AO7 degradation efficiency of Mn3O4/NF-0.48 under different systems. (d) UV-vis spectra of AO7 solution under given time intervals in the presences of Mn3O4/NF-0.48 and PMS. Conditions: [AO7]0 = 20 mg/L, [oxone]0 = 1 mM.
Molecules 28 04312 g003
Figure 4. The influences of PMS concentration (a), initial pH value (b), and reaction temperature (c) on AO7 degradation with Mn3O4/NF-0.48. (d) The corresponding Arrhenius curve at different reaction temperatures. Conditions: (a) [AO7]0 = 20 mg/L; (b,c) [AO7]0 = 20 mg/L, [oxone]0 = 1 mM.
Figure 4. The influences of PMS concentration (a), initial pH value (b), and reaction temperature (c) on AO7 degradation with Mn3O4/NF-0.48. (d) The corresponding Arrhenius curve at different reaction temperatures. Conditions: (a) [AO7]0 = 20 mg/L; (b,c) [AO7]0 = 20 mg/L, [oxone]0 = 1 mM.
Molecules 28 04312 g004
Figure 5. (a) Recycling study of AO7 degradation. (b) Pseudo-first-order kinetic constant (k, min−1) versus cycle number.
Figure 5. (a) Recycling study of AO7 degradation. (b) Pseudo-first-order kinetic constant (k, min−1) versus cycle number.
Molecules 28 04312 g005
Figure 6. Catalyst sample after four cycles of PMS activation on AO7 degradation: (a) XRD pattern; (b) Raman spectrum.
Figure 6. Catalyst sample after four cycles of PMS activation on AO7 degradation: (a) XRD pattern; (b) Raman spectrum.
Molecules 28 04312 g006
Figure 7. Comparisons of Mn 3s (a), Mn 2p (b), and O 1s (c) core-level XPS spectra of Mn3O4/NF-0.48 (1) and the catalyst samples after one cycle (2) and four cycles (3) of AO7 degradation.
Figure 7. Comparisons of Mn 3s (a), Mn 2p (b), and O 1s (c) core-level XPS spectra of Mn3O4/NF-0.48 (1) and the catalyst samples after one cycle (2) and four cycles (3) of AO7 degradation.
Molecules 28 04312 g007
Figure 8. SEM images of the catalyst sample after one cycle (ac) and four cycles (df) of AO7 degradation. Sample after four cycles: (g,h) TEM images; (i) SAED pattern.
Figure 8. SEM images of the catalyst sample after one cycle (ac) and four cycles (df) of AO7 degradation. Sample after four cycles: (g,h) TEM images; (i) SAED pattern.
Molecules 28 04312 g008
Figure 9. (a) Nyquist plots of the Mn3O4/NF-0.48 and the sample after four cycles of AO7 degradation. (b) Z’ versus the reciprocal of the square root of frequency (ω−0.5) in the intermediate frequency range.
Figure 9. (a) Nyquist plots of the Mn3O4/NF-0.48 and the sample after four cycles of AO7 degradation. (b) Z’ versus the reciprocal of the square root of frequency (ω−0.5) in the intermediate frequency range.
Molecules 28 04312 g009
Figure 10. EPR spectra of Mn3O4/NF-0.48/PMS system captured by DMPO.
Figure 10. EPR spectra of Mn3O4/NF-0.48/PMS system captured by DMPO.
Molecules 28 04312 g010
Figure 11. Scheme for the phase transition induced by PMS activation.
Figure 11. Scheme for the phase transition induced by PMS activation.
Molecules 28 04312 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

Liu, C.; Wang, Z.; Chen, Y.; Zeng, X.; Long, H.; Rong, H.; Zou, H.; Ding, J.; Li, J. Peroxymonosulfate-Activation-Induced Phase Transition of Mn3O4 Nanospheres on Nickel Foam with Enhanced Catalytic Performance. Molecules 2023, 28, 4312. https://doi.org/10.3390/molecules28114312

AMA Style

Liu C, Wang Z, Chen Y, Zeng X, Long H, Rong H, Zou H, Ding J, Li J. Peroxymonosulfate-Activation-Induced Phase Transition of Mn3O4 Nanospheres on Nickel Foam with Enhanced Catalytic Performance. Molecules. 2023; 28(11):4312. https://doi.org/10.3390/molecules28114312

Chicago/Turabian Style

Liu, Cuiyin, Ziyan Wang, Yanfeng Chen, Xinjuan Zeng, Hangyu Long, Haibo Rong, Hongtao Zou, Jinpeng Ding, and Jingling Li. 2023. "Peroxymonosulfate-Activation-Induced Phase Transition of Mn3O4 Nanospheres on Nickel Foam with Enhanced Catalytic Performance" Molecules 28, no. 11: 4312. https://doi.org/10.3390/molecules28114312

Article Metrics

Back to TopTop