Four kinds of manganese oxides were successfully prepared by hydrothermal and redox precipitation methods, and the obtained oxides were used for CIP removal from water by activating PMS. The microstructure and surface properties of four oxides were systematically characterized. The results showed that ε-MnO2 prepared by the redox precipitation method had large surface area, low crystallinity, high surface Mn(III)/Mn(Ⅳ) ratio and the highest activation efficiency for PMS, that is, when the concentration of PMS was 0.6 g/L, 0.2 g/L ε-MnO2 could degrade 93% of CIP within 30 min. Multiple active oxygen species, such as sulfate radical, hydroxyl radical and singlet oxygen, were found in CIP degradation, among which sulfate radical was the most important one. The degradation reaction mainly occurred on the surface of the catalyst, and the surface hydroxyl group played an important role in the degradation. The catalyst could be regenerated in situ through the redox reaction between Mn4+ and Mn3+. The ε-MnO2 had the advantages of simple preparation, good stability and excellent performance, which provided the potential for developing new green antibiotic removal technology.

  • Epsilon-MnO2 nanoflowers were simply prepared by the redox precipitation method.

  • Epsilon-MnO2 nanoflowers were used to activate PMS to degrade CIP in water.

  • The high activity of ε-MnO2 was due to its large surface area, low crystallinity and high surface Mn(III)/Mn(Ⅳ) ratio.

  • The –OH groups on the surface could promote the activation of PMS.

Ciprofloxacin (CIP) is one of the third-generation fluoroquinolone antibiotics. Due to its good clinical efficacy and strong antibacterial activity, it has become one of the most commonly prescribed antibiotics in medicine (Anirudh Gupta 2019; Alhaddad & Amin 2022; Enyoh & Wang 2023). However, more than 75% of CIP cannot be absorbed by humans and animals, resulting in a large amount of CIP residues being released into the water environment (Huang et al. 2018a). Recently, it has been detected frequently in surface water, groundwater and even drinking water, with the CIP concentrations ranging from ng L−1 to mg L−1 (Chun-Hui Shen 2019; Hu et al. 2020; Omuferen et al. 2022). Unfortunately, due to the poor biodegradability and high ecotoxicity of CIP, it is impossible to effectively remove it from water treatment by various traditional sewage technologies. In recent years, it has been found that the emergence of antibiotic-resistant bacteria and resistant genes due to the extensive use of antibiotics will pose a severe threat to human health and ecosystem security (Yu et al. 2020; Falyouna et al. 2022). When the World Health Organization revised antibiotic drugs in 2017, it classified some antibiotics as ‘prudent use’ in an attempt to reduce their use. In addition to effective control at the source, developing treatment technologies to remove such antibiotics from water and wastewater systems is essential.

At present, various methods have been investigated to reduce CIP from water, such as membrane filtration (Nasrollahi et al. 2022), adsorption (Arif et al. 2022), biodegradation (Shah et al. 2022) and advanced oxidation process (AOP) (Salari et al. 2022). Among them, the AOP based on sulfate radical has aroused much concern in recent years due to its advantages of economy, high efficiency, environmental protection, safety and stability (Tian et al. 2022). Persulfate (PS) includes peroxymonosulfate (PMS) and peroxydisulfate (PDS), both of which contain –O–O– bonds in their molecules. The method for activating PS degradation is to couple and excite PS through specific physical or chemical approaches to break its –O–O– bond and produce strong oxidizing , which reacts with organic substances through dehydrogenation, electron addition, electron transfer and other ways to trigger a series of free radical chain reactions, to realize the oxidative degradation of organic pollutants. Currently, PS can be activated in various ways, such as UV activation (Ghauch et al. 2017), thermal activation (He et al. 2021a), ultrasonic activation (Gujar et al. 2023), transition metal ion activation (Fe2+, Mn2+, Cu2+, Ag+, Co2+, Ce3+, etc.) (Huang et al. 2009; Gao et al. 2018; Yoon et al. 2022) and heterogeneous catalysts activation (Tian et al. 2022; Zhou et al. 2023). UV activation, thermal activation and ultrasonic activation all require external energy. Transition metal ion activation has difficulties in recovery and secondary pollution. In contrast, heterogeneous catalytic materials have become a research hotspot due to their high efficiency, low cost and low energy consumption (Zhang et al. 2021).

Transition metal-based heterogeneous materials have been reported to be effective in activating persulfates (Rama et al. 2018; Wang & Wang 2018), but cobalt-based or copper-based materials may cause health hazards due to the immersion of toxic ions in the aqueous phase, silver-based materials are expensive, which will limit their practical application. In contrast, manganese-based materials have obvious advantages such as many valence states, high oxygen migration efficiency, abundant resources, low price, friendly environment and strong safety (Huang & Zhang 2019). Manganese oxide has been reported as an effective catalyst for ozone oxidation (He et al. 2021b; Issaka et al. 2022) and also exhibited excellent performance in catalytic degradation of pollutants by hydrogen peroxide (Hussain et al. 2021; Zhen et al. 2023).

The synthesis route was of great importance to the physical and chemical properties and structural performance of the catalyst. For decades, a series of synthesis techniques have been used to synthesize manganese oxides, such as hydrothermal (Wang et al. 2015a), solvothermal (Ren et al. 2014), sol–gel (Zhang et al. 2019), template (Huang et al. 2018b) and redox precipitation (Wei et al. 2023) methods. Hydrothermal and solvothermal synthesis was characterized by high pressure, long period and strict requirements for reactants. There were many steps and uncontrollable factors in the process of sol–gel synthesis, which was difficult to industrialize; The purity of the products synthesized by the template method was not high, and the selection of template agent was very demanding. These shortcomings limited the practical application of many manganese oxides. The synthesis method with high efficiency, low cost and simple operation has become an urgent problem to be solved in the development of new manganese oxide catalysts. Redox precipitation method effectively avoided complex working procedures and special equipment requirements, and it was worth studying because of its simple process, fast reaction speed and diverse morphology of samples.

In recent years, manganese oxides, such as MnO2, Mn3O4 and MnO(OH), have attracted much attention in the field of catalysts because of their physical and chemical properties. Wang et al. (2015b) found that α-MnO2 showed shape dependence when used to degrade phenol; Saputra et al. (2013b) concluded through research that Mn3O4 had better performance than Fe3O4 and Co3O4 in activating PMS to degrade pollutants. He et al. (2020) synthesized MnO(OH) with the same crystal structure to degrade 2,4-dichlorophenol and found that different morphologies would lead to different physical and chemical properties, thus providing guidance for rational design of catalysts. However, there were few studies on the activity of a series of manganese oxides with different crystal structures in environmental catalysis. At the same time, we also found that the research on MnO2 was mostly focused on polycrystals. Epsilon-MnO2 showed low crystallinity due to its own structural fracture (i.e. De Wolff fracture) and structural defect (i.e. micro-twin) (Wang et al. 2015a), and its catalytic performance in water treatment was rarely touched.

Herein, four different manganese (hydroxyl) oxides were prepared by hydrothermal and redox coprecipitation methods, and the influence of catalyst structure on catalytic performance was studied by systematic characterization. The degradation of CIP by PMS activated by ε-MnO2 with low crystallinity was systematically studied to explore the green catalytic materials that were easy to prepare and highly active to degrade antibiotics.

Materials

All chemicals are used directly after purchase without any treatment. Manganese chloride (MnCl2·4H2O), manganese sulfate (MnSO4·H2O), potassium permanganate (KMnO4), L-histidine (C6H9N3O2) and ciprofloxacin (C17H18FN3O3) were bought from Shanghai Zhongqin Chemical Reagent Co., Ltd (Shanghai, China). Anhydrous ethanol (CH3CH2OH), sulfuric acid (H2SO4), methanol (CH3OH), tert-butanol (C4H10O), sodium hydroxide (NaOH), potassium nitrate (KNO3), phenol, potassium dihydrogen phosphate (KH2PO4), humic acid (HA), sodium chloride (NaCl), sodium carbonate (Na2CO3) and nitrobenzol (NB) were provided by Tianjin Damao Chemical Reagent Co., Ltd (Tianjin, China). PMS (KHSO5·0.5KHSO4·0.5 K2SO4, PMS), 5, 5-dimethyl-1-pyrrole Lino-N-oxide (DMPO), 2,2,6,6-tetramethylpiperidine (TEMP) were purchased from Aladdin Reagent Co., Ltd (Shanghai, China).

Manganese (hydroxyl) oxide synthesis

The valence heterostructure manganese oxides (MnOx) were prepared by hydrothermal method. Typically, 0.79 g (5 mmol) of KMnO4 was dissolved in 50 mL of a mixed solvent of ethanol and water, 0.99 g (5 mmol) of MnCl2·4H2O was added, stirred for about 30 min until the mixture was uniform and transferred to 100 mL of polytetrafluoroethylene (PTFE) liner stainless steel jacket reactor for 160° hydrothermal treatment for 12 h. The brown precipitates were washed three times with deionized water and anhydrous ethanol, respectively and then dried at 70° C for 24 h. In the experiment, the volume ratio of ethanol and water in the solvent was adjusted to 0:100, 50:50 and 100:0, respectively, to prepare nanomaterials with different crystal forms and morphologies, which were recorded as MnOx-1, MnOx-2 and MnOx-3.

Manganese oxide (MnOx-4) was synthesized by the redox coprecipitation method. 0.004 mol (0.6321 g) KMnO4 was added to 200 mL of 0.03 mol/L MnSO4·H2O solution and stirred for 30 min. Afterward, the solution was placed in a collector magnetic stirrer at 75 °C and went for 5 h to obtain a black precipitate. The precipitate was washed for several times alternately with deionized water and anhydrous ethanol until the supernatant was neutral, then it was put into a 60 °C constant temperature drying oven for 24 h after suction filtration and grounded to powder after cooling, which was denoted as MnOx-4.

Characterization

The microstructure of the catalysts was studied by X-ray diffractometer (XRD, X'pert pro) using the scanning angle (2θ) at 10–90° and Cu–Kα radiation at 40 kV and 40 mA. The microtopography of the catalysts was characterized by scanning electron microscopy (SEM, Apreo S) and transmission electron microscope (TEM, FEI Tecnai F 30). The sample elements and surface topography were analyzed using energy-dispersive methods (EDS, EDAX). The chemical bonds and functional groups of the catalysts were measured by a Fourier-transform infrared spectrometer (FTIR, VERTEX 70, Bruker). X-ray photoelectron spectroscopy (XPS, PHI 5702) was mainly used to characterize the elemental composition and valence state of the catalyst surface. The Brunauer–Emmett–Teller (BET) surface areas were determined by N2 adsorption tester (JW-BK200C). Electron paramagnetic resonance spectroscopy (EPR, Bruker EMX plus) was used to identify free radicals in solution and solid oxygen vacancies. The detection parameters were as follows: microwave frequency 9.8434 GHz, scanning width 100 G, central field 3,505.00 G, modulation amplitude 1 G, time constant 5.12 ms and scanning time 60 s.

Catalytic degradation

The catalytic performance of the prepared four kinds of MnOx was investigated by activating PMS to degrade CIP in water. All degradation experiments were carried out in 250 mL beakers on a constant temperature magnetic stirrer (speed 1,200 rpm). First, a certain amount of catalyst was added to the 200 mL prepared target pollutants (10 mg·L−1), after the catalyst reached adsorption equilibrium with the solution, a certain amount of PMS was added to start the timing. Five millilitres (5 mL) sample was taken from the beaker at regular intervals, filtered through a 0.45 μm microfiltration membrane and quenched with methanol. The CIP concentration was determined by measuring the absorbance at the maximum absorption wavelength of 277 nm with an ultraviolet spectrophotometer (UV-2800A).

When exploring the influence of different parameters on the degradation of pollutants, the single-variable principle was used for experimental analyses. During the investigation, 0.1 mol·L−1 H2SO4 and NaOH solutions were used to regulate the pH of the solution to study the influence of the initial pH value. To check the effect of quenchers on degradation experiments, MeOH, TBA, L-His, phenol, KH2PO4 and NB were selected as quenchers and set blank control at the same time. Samples were taken at regular intervals for testing after the degradation began. To explore the stability and reuse performance of the catalyst, the cyclic degradation test of the catalyst was carried out. After a single reaction, the catalyst was recovered and washed repeatedly, then put into a constant temperature drying oven at 80 °C for drying and reuse.

When the synthesized catalyst was further applied to the removal of CIP from actual water, the raw water obtained from Liujiaxia Reservoir in the upper reaches of the Yellow River was first filtered with 0.45 μm PTFE filter, which was used to dissolve CIP and configured into a solution of target concentration, and the removal effect of ε-MnO2/PMS on CIP was observed.

Catalyst characterization

The crystal structure of the prepared catalyst could be judged by the XRD diffraction pattern of the sample and its matching with the standard pattern Joint Committee on Powder Diffraction Standards (JCPDS). The XRD spectrum of the four manganese oxides is shown in Figure 1(a). It was observed that the change in the content of ethanol in the solvent directly affected the phase of the product. When the solvent was water, the diffraction peaks of MnOx-1 appeared at 2θ of 18.107°, 28.841°, 37.522°, 41.968°, 56.372° and 69.711°, corresponding to (200) (310) (211) (301) (600) and (541) planes of the Tetragonal α-MnO2 (PDF #44-0141). The diffraction peaks of MnOx-2 prepared by 50% hydrothermal water and 50% ethanol appeared at 2θ of 26.149°, 35.538°, 37.168°, 51.222°, 54.864°, 61.797°, 71.152°, 81.503° and 85.478°, corresponding to Monoclinic γ-MnO(OH) (11-1), (111), (002), (022), (11-3), (131), (113), (2-2-4) and the (242) plane (PDF #41-1379). The diffraction peaks of MnOx-3 prepared with ethanol appeared at 2θ of 18.00°, 28.88°, 32.315°, 36.449°, 37.983°, 44.44°, 50.709°, 58.510°, 59.84°, 64.651° and 74.151°, corresponding to Tetragonal Mn3O4 (101), (112), (103), (202), (004), (220), (105), (321), (224), (400) and (413) planes (PDF #24-0734). This might be due to the fact that when the solvent was all water, KMnO4 was reduced by Mn2+ to MnO2, as shown in Equation (1). When the solvent was all ethanol, KMnO4 was reduced by ethanol to produce Mn3O4, as shown in Equation (2). The production of Mn3O4 was accompanied by the production of KOH, and the solution was alkaline. The increase of ethanol in the solvent caused a large amount of OH to be generated in the solution, which combined with Mn2+ to produce Mn(OH)2. The unstable Mn(OH)2 was rapidly oxidized to MnO(OH) (Mn has a +3 valence). The diffraction peaks of MnOx-4 prepared by the redox coprecipitation method appeared at 2θ of 37.12°, 42.401°, 56.027° and 66.761°, corresponding to (100), (101), (102) and (110) crystal planes of hexagonal ε-MnO2 (PDF #30-0820), respectively. The diffraction peak positions of the four catalysts could be matched with the PDF card, and there were almost no impurity peaks. The diffraction peaks of the first three manganese oxides were sharp, indicating that the catalyst prepared by the hydrothermal method had good crystallinity, while the fourth catalyst prepared by the redox precipitation method had wide diffraction peaks and low crystallinity.
(1)
(2)
Figure 1

XRD patterns (a) and FTIR spectra (b) of four manganese (hydrogen) oxides.

Figure 1

XRD patterns (a) and FTIR spectra (b) of four manganese (hydrogen) oxides.

Close modal

The FTIR of the four manganese oxides is shown in Figure 1(b). It could be seen that the absorption peaks of the four manganese oxides appeared at 3,425 and 1,633 cm−1, which could be related to the stretching vibration and bending vibration of hydroxyl. The absorption peaks of γ-MnO(OH) at 1,150 and 1,085 cm−1 were attributed to the bending vibration of hydroxyl, and the absorption peaks at 2,674 and 2,080 cm−1 were assigned to the stretching vibration of hydroxyl. This indicated that there were abundant hydroxyl groups on the surface of the four catalysts. The absorption peaks in the 400–800 cm−1 corresponded to the vibrational absorption of Mn–O metal bonds in manganese oxides. FTIR showed that the Mn–O metal bonds had been formed in the four catalysts, and a large number of hydroxyl groups existed on the surface of the catalysts.

Figure 2(a)–2(f) shows the morphological characteristics of the four manganese oxides. The α-MnO2 and γ-MnO(OH) prepared by solvent containing water were both nanorods with smooth surfaces and a small number of nanoparticles. The diameter of the latter was larger than that of the former, which was about 200 nm (Figure 2(b)), and the former was about 50 nm (Figure 2(a)). Combined with XRD, it was speculated that the two rods were α-MnO2 and γ-MnO(OH). From Figure 2(c), it could be seen that Mn3O4 presented nano-spherical particles with uniform particle size, the spherical surface was smooth, the diameter was about 50–100 nm and the particles tended to agglomerate. According to the XRD pattern, it was speculated to be Mn3O4 spherical particles. With the increase of the proportion of ethanol in the hydrothermal solvent, the morphology of the catalyst prepared under the same conditions gradually changed from a slender rod to a larger diameter and finally transited to a smaller spherical diameter, with little change in the surface roughness. In these reactions, ethanol not only acted as a solvent to provide the reaction environment, but also participated in the reaction as a reducing agent, thereby affecting the composition, phase and morphology of the products. The surface of ε-MnO2 was rough in Figure 2(d), showing cluster-like spherical nanoflowers with a small number of nanoparticles attached. The diameter of the nanoflowers was about 1 μm, which was speculated to be ε-MnO2 in combination with XRD. TEM images (Figure 2(e)) further confirmed the mesoporous nanoflower structure of ε-MnO2, showing clear lattice streaks in Figure 2(f) with a spacing of 0.242 nm, corresponding to the (100) face of ε-MnO2. As shown in EDS (Figure 2(g)), ε-MnO2 nanomaterials were composed of Mn and O elements, and the coexisting element ratio Mn/O was 2.1, which was close to the theoretical stoichiometric value of ε-MnO2.
Figure 2

SEM images of α-MnO2 (a),γ-MnO(OH) (b), Mn3O4 (c), ε-MnO2 (d), TEM images of ε-MnO2 (e,f); EDS of ε-MnO2 (g), N2 adsorption–desorption isotherms of ε-MnO2 (h).

Figure 2

SEM images of α-MnO2 (a),γ-MnO(OH) (b), Mn3O4 (c), ε-MnO2 (d), TEM images of ε-MnO2 (e,f); EDS of ε-MnO2 (g), N2 adsorption–desorption isotherms of ε-MnO2 (h).

Close modal

According to the measurement of N2 adsorption and desorption, the four samples in Figure 2(h) presented a typical type-IV isotherm, which indicated that all four catalysts were mesoporous materials with average pore diameters ranging from 2 to 50 nm, and there were obvious H3-type hysteresis loops between the adsorption and desorption isotherms of Mn3O4 and ε-MnO2 in the range of P/P0 of 0.4–1.0 and 0.8–1.0, respectively. It could be inferred that these two catalysts were composed of slit-like pores of flaky particles (Yang et al. 2001). As shown in Table S1, the BET surface area of ε-MnO2 was 156.43 m2/g, which was 8.25 times that of α-MnO2, 25.35 times of γ-MnO(OH) and 6.66 times of Mn3O4. The pore volume of ε-MnO2 was 0.42 cm3/g, which was 7.5 times that of α-MnO2, 28.49 times of γ-MnO(OH) and 2.43 times of Mn3O4.

The elemental composition and valence state of the catalyst were investigated by XPS. It could be seen from the full spectrum diagram of catalyst XPS (Figure 3(a)) that the four catalysts mainly contain two elements, Mn and O (element C comes from the test instrument). The XPS spectrum of Mn 2p3/2 in the sample is shown in Figure 3(b). The binding energy positions of Mn4+, Mn3+ and Mn2+ were determined by comparing literatures (Saputra et al. 2013b; He et al. 2020; Lin & Chen 2020; Liu et al. 2021), and the relative contents of Mn in different valence states could be obtained by the proportion of peak areas after peak fitting. The relative contents of Mn3+ in α-MnO2, γ-MnO(OH), Mn3O4, ε-MnO2 were 71.21, 70.68, 21.6 and 85.4%, respectively. The relative contents of Mn4+ in α-MnO2, γ-MnO(OH), Mn3O4, ε-MnO2 were 28.79, 29.32, 5.8 and 14.6%, respectively, and the content of Mn2+ in Mn3O4 was 72.6%. The XPS spectrum of sample O1s was shown in Figure 3(c), the peaks in 529.0–530.0 eV, 531.0–532.0 eV and around 533 eV belonged to lattice oxygen (Olatt), surface adsorbed oxygen (Oads) and surface hydroxyl oxygen (OadsO-H), respectively. The relative content of Olatt was 51.03, 26.92, 56.20 and 36.2% in α-MnO2, γ-MnO(OH), Mn3O4 and ε-MnO2, and the content of Oads was 37.04, 66.53, 43.80 and 63.8%, respectively. OadsO-H was found only in α-MnO2 and γ-MnO(OH) with relative contents of 11.93 and 6.55%, respectively.
Figure 3

XPS spectra of survey (a), Mn 2p (b), O 1s (c) of α-MnO2, γ-MnO(OH), Mn3O4 and ε-MnO2 before and after reaction; EPR of ε-MnO2(d).

Figure 3

XPS spectra of survey (a), Mn 2p (b), O 1s (c) of α-MnO2, γ-MnO(OH), Mn3O4 and ε-MnO2 before and after reaction; EPR of ε-MnO2(d).

Close modal

CIP degradation in different reaction systems

It could be seen from Figure 4(a) that the degradation of CIP by α-MnO2, γ-MnO(OH), Mn3O4 and ε-MnO2 in the absence of PMS was 28.65, 9.75, 18.71 and 43.66%, respectively, which might be caused by adsorption or direct oxidation. Only 10.53% of CIP was removed in a separate PMS system, indicating that PMS was difficult to degrade pollutants directly. After adding catalysts to the system, the degradation efficiency of α-MnO2, γ-MnO(OH), Mn3O4 and ε-MnO2 in the PMS system for CIP reached 41.52, 32.55, 71.73 and 94.35%, respectively. ε-MnO2 also showed excellent activity compared to other MnO2 in Table S2 that had been reported. All the reaction characteristics conformed to the first-order reaction kinetics (R2 > 0.95) (Figure 4(b)), and the average degradation rate constant k of CIP in ε-MnO2/PMS system was 0.09429, which was 5.20, 7.04 and 2.2 times higher than that in α-MnO2/PMS, γ-MnO(OH)/PMS and Mn3O4/PMS systems, respectively. The results showed that all four catalysts could activate PMS to degrade pollutants, and ε-MnO2 had the best effect, which might be related to its structure and surface physicochemical properties.
Figure 4

Degradation of CIP (a) and reaction rate constants (b) in different reaction systems. Reaction conditions: [CIP]0 = 10 mg/L, [catalyst] = 0.2 g/L, [PMS] = 0.6 g/L, temperature = 25 °C, pH was not adjusted.

Figure 4

Degradation of CIP (a) and reaction rate constants (b) in different reaction systems. Reaction conditions: [CIP]0 = 10 mg/L, [catalyst] = 0.2 g/L, [PMS] = 0.6 g/L, temperature = 25 °C, pH was not adjusted.

Close modal

It has been reported that the catalytic activity of manganese oxides of the same type was affected by the structure, especially the tunnel size and crystallinity (Saputra et al. 2013b; Huang et al. 2019). For instance, α-MnO2 with (2 × 2) tunnel structure had a larger size and higher degradation efficiency than β-MnO2 with (1 × 1) tunnel structure and γ-MnO2 with (1 × 1) and (1 × 2) tunnel structure. The four catalysts prepared were α-MnO2 with (2 × 2) tunnel structure, γ-MnO(OH) with (1 × 1) tunnel structure, Mn3O4 with spinel-like structure and ε-MnO2 with (1 × 1) tunnel and (2 × 1) tunnel in a matrix of random coexistence, respectively. The degradation efficiency followed the order ε-MnO2 > Mn3O4 > α-MnO2 > γ-MnO(OH), which was inconsistent with the reported results. It might be that the mesoporous nanostructures of ε-MnO2 exhibited more active sites due to its defects. The ε-MnO2 with the highest reaction efficiency was found to have the lowest crystallinity among the four catalysts, which was also different from the conclusions in the literature. Combined with the EPR spectrum of ε-MnO2 in Figure 3(d), a sharp symmetric peak appears at g = 2.011, which could be considered as the solid defect existing in the structure of ε-MnO2 led to the formation of oxygen vacancy (Zhao et al. 2020), thus improving the catalytic activity.

By comparing the SEM of four nanoparticles, it was found that the reactivity of nanorods and spherical nanoparticles was higher than that of nanorods and spherical nanoparticles, which was consistent with the findings of Wang and Jing et al. (Wang et al. 2015b; Jing et al. 2017), that is, the nanoflower shape shows the greatest ability to activate PMS, which was conducive to the heterogeneous reaction on the catalyst surface.

The order of BET surface area and pore volume was ε-MnO2 > Mn3O4 > α-MnO2 > γ-MnO(OH), which was consistent with the results of CIP degradation by activated PMS. The larger surface area meant more exposed active sites, and the larger void volume could reduce the diffusion resistance, thus showing better adsorption performance and catalytic performance (Liu et al. 2020), which was beneficial to the activation of PMS.

In addition, it was reported that Mn(III) content played an important role in Mn-based catalysts (Meng et al. 2014), because it was generally believed that was generated by Mn (III) reaction (Equation (3)), while was generated by Mn (IV) reaction (Equation (4)), and was higher than oxidation–reduction potential (Wen-Da et al. 2016). Unfortunately, the content of Mn(III) in the four catalysts did not show a significant correlation with the degradation efficiency of pollutants. However, it was found that the Mn(III)/Mn(Ⅳ) ratio in the catalyst was positively correlated with the degradation efficiency of CIP, which might be related to the redox reaction occurring during the degradation of manganese oxides.

(3)
(4)

In summary, it was not difficult to see that the chemical properties and reactivity of manganese oxides were highly dependent on their structure. High BET surface area could provide more reaction sites; ε-MnO2 with low crystallinity might have the best activation performance because of its own defects leading to the formation of oxygen vacancy. The oxidation capacity of manganese oxides was positively correlated with the ratio of Mn(III)/Mn(Ⅳ) content, which might indicate that high Mn(III)/Mn(Ⅳ) content could promote the redox reaction in the process of manganese oxides activation of PMS.

Effect of reaction conditions

The process of ε-MnO2 activating PMS to degrade CIP was influenced by many reaction factors. The effects of catalyst dosage, PMS dosage, initial pH of solution, reaction temperature, common anions in water and actual water on CIP degradation were discussed, and the optimal reaction parameters of the system were determined to provide a reference for engineering practice.

Figure s1a showed when the dosage of ε-MnO2 was 0.05, 0.1, 0.2, 0.4, 0.6 and 1.0 g·L−1, the degradation efficiencies of CIP after 30 min of reaction reached 66.7, 83.8, 92.4, 95.3, 95.9 and 97.1%, respectively. Obviously, the increase of catalyst concentration could promote the degradation of CIP, because with the increase of catalyst content in the reaction system, more active sites would be generated in the system, and more active components would be produced to remove more CIP. However, when the catalyst dosage was higher than 0.2 g/L, the removal rate of CIP increased slowly. Therefore, 0.2 g/L was selected as the catalyst concentration of the system.

As the source of in the reaction system (Equation (5)), the concentration of PMS would have an immediate impact on the concentration of free radicals and thus the degradation of CIP. In this experiment, the influence of different PMS concentrations on the CIP removal rate was discussed, and the results are shown in Figure s1b. With the concentration of PMS increasing from 0.1 to 0.6 g/L, the removal rate of CIP rose from 69.0 to 90.8% after 30 min of reaction. However, when the concentration of PMS continued to increase to 1.0 g/L, the removal rate of CIP decreased by 0.2%, which was mainly due to the autolysis (Equation (6)) and quenching caused by excess PMS (Equation (7)). Therefore, the concentration of PMS was selected as 0.6 g/L in this experiment.
(5)
(6)
(7)
According to relevant literature (Chen et al. 2017), AOPs based on had better pH adaptability than traditional AOPs. Therefore, under the conditions of determining the amount of catalyst and PMS, the effect of pH on the degradation of CIP was investigated. As shown in Figure S1c that the ε-MnO2/PMS system could maintain a high degradation efficiency in the range of pH 2–11, in which more than 90% of CIP could be removed within the range of initial solution pH 3–7, and the removal rate of CIP was close to 90% (85.6 and 88.5%) at pH 2 and 9, indicating that the system had good adaptability in both acidic and weak alkaline environments. When the pH was 11, the removal rate of CIP was reduced to 78.4%. This aspect was that when the PH of the solution was higher than the zero-point charge (PZC) of ε-MnO2 (Lin & Chen 2020), the surface of ε-MnO2 was negatively charged, generating charge repulsion with , which hindered the contact between them and reduced the activity. In addition, the transformation of free radicals (Equation (8)) under alkaline conditions also affected the degradation efficiency. The redox potential of •OH was lower than that of , thus reducing the reaction rate. Therefore, ε-MnO2 could activate PMS under conventional conditions without additional pH adjustment, which provided convenience for practical applications.
(8)
Figure S1d shows the influence of different temperatures on the degradation of CIP by ε-MnO2-activated PMS. It could be seen from the figure that the removal rate of CIP increased with the temperature rise. When the reaction temperature was 15, 25 and, 35 °C, the CIP removal rates were 83.5, 90.8 and, 92.0%, respectively, after 30 min reaction. This result could be attributed to the fact that PMS could be decomposed into (Equation (9)) through thermal activation. In addition, high temperatures could speed up the diffusion reaction in the system, thus accelerating the degradation of CIP. The Arrhenius formula (Equation (10)) was used to calculate the activation energy (Ea) of the reaction to estimate the effect of thermal energy. The calculated result was 12.52 kJ mol−1, indicating that the reaction was not only controlled by diffusion but also by heterogeneous chemical reaction. Compared with other manganese-based catalysts for activating PS degradation reactions (as shown in Table s3), the Ea of the ε-MnO2 reaction was significantly reduced, indicating that the reaction could be carried out under low energy conditions, thus demonstrating its superior catalytic performance.
(9)
(10)
where k is the rate constant, the fitting result was displayed in Figure s1d inset. A is the prefactor, R is the gas constant (8.314 J mol−1 K−1) and T is the absolute temperature (K).

There were often various inorganic anions and natural organic matter in actual water, such as Cl, , , etc., which would affect the activation efficiency of PMS by quenching free radicals and competing with the active site of the catalyst. Humic acid (HA) was the most common natural organic matter in water, mainly consisting of aromatic ring, hydroxyl group and phenolic hydroxyl group, which might adsorb on the surface of the material to occupy the active site or compete for free radicals in water, thus affecting the removal effect of pollutants. Figure S1e shows the effect of 5 mmol/L Cl, , and HA, as well as the Yellow River water as solvents on the degradation of CIP by ε-MnO2-activated PMS. Only and HA among them had inhibitory effects on the system, the other had little effect. As shown in the figure, significantly inhibited the degradation of CIP in the system, which might be due to the competitive reaction of both and CIP, resulting in the formation of intermediates (Equation (11)) with low or no oxidation performance (Lin & Tsai 2019). In addition, the addition of would increase the alkalinity of the solution, which was not conducive to the degradation of CIP according to the research conclusion of the influence of PH value. HA had a certain inhibitory effect on CIP degradation, which might be that HA occupied the active site of MnO2 and affected its binding with PMS to form reaction intermediates, or that HA competed with CIP to react with active substances, leading to a decrease in CIP degradation efficiency. In addition, ɛ-MnO2/PMS also had excellent removal performance in CIP solution with Yellow River water as solvent, which indicated that ε-MnO2/PMS system had a good application prospect.

(11)

Catalytic stability

In order to explore the stability of the catalyst, the prepared catalyst ε-MnO2 was tested for cyclic degradation. It was apparent from Figure S2 that ε-MnO2 was relatively stable and still exhibited excellent catalytic performance after four cycles of degradation experiments. The removal rate of CIP in the fourth cycle in 30 min was 77.4%, 15.7% lower than that in the first cycle. The reasons could be attributed to two aspects: with the increase in catalyst cycle degradation times, the active components on the catalyst surface might be lost, resulting in a gradual reduction of active sites. In addition, the small molecule intermediates generated in the process of organic degradation gathered on the catalyst surface, which hindered the contact of the catalyst with PMS and CIP, thus reducing the removal rate of CIP.

Exploration of the catalytic mechanism

To explore the main reaction active species in the system, tert-butanol (TBA), methanol (MeOH) and L-histidine (L-His) were used as quenchers to investigate the effect of CIP degradation in the ε-MnO2/PMS system. According to the literature (Hao et al. 2019), the reaction rate constants of MeOH and TBA with •OH and were shown in Table s4, in which the reaction rate constant of MeOH and •OH was 50 times that of , while the reaction rate constant of TBA and •OH was 1,000 times that of . When the solution contains two free radicals simultaneously, it could be considered that TBA only quenched •OH, while MeOH quenched •OH and . The contribution of the two free radicals was judged by comparing the effects of two kinds of quenching agents on the degradation of CIP in the system. L-histidine could be used to quench singlet oxygen (Chen et al. 2018) with a second-order rate constant of 3 × 107 (D2O, pD = 7) (Wilkinsan 1981). So MeOH (3 M), TBA (3 M) and, L-His (0.2 mM) were used to identify the presence of active species such as , •OH and, 1O2 in the system, respectively.

As described in Figure 5(a), after TBA was added to the system, the removal rate of CIP decreased from 91.9 to 79.2% 30 min later, indicating the presence of •OH in the system. The CIP removal rate was further reduced to 72.6% after the addition of MeOH, which meant that not only •OH but also existed in the ε-MnO2/PMS system. Compared with the case of no quencher, the removal rate of CIP decreased by 13.3% after adding L-His to the system, which indirectly indicated that there was still 1O2 in the ε-MnO2/PMS system to promote the degradation of CIP. The inhibition effect of MeOH on CIP was less than that of TBA, indicating that •OH was not the main free radical in the system. The addition of MeOH could consume the generated and •OH, but these two quenchers had no significant effect on CIP degradation, so it was speculated that the generated free radicals were mainly attached to the surface of the catalyst, and MeOH and TBA were difficult to quench. In order to verify this statement, phenol was added to the reaction system as a quenching agent. The reaction rate constant of phenol with and •OH was high and the polarity was weak, so it could effectively quench the free radicals on the catalyst surface (Yan et al. 2017). After the addition of phenol, the degradation efficiency decreased to 51.77%, and the inhibition of CIP degradation by phenol was significantly higher than that by MeOH and TBA, indicating that free radicals were mainly produced on the surface of the catalyst. In addition, NB could be used to quench free radicals on the catalyst surface due to its hydrophobicity. The reaction rate of NB with •OH was 3,000 times that of (fast reaction) (Neta et al. 1988). Therefore, the effect of •OH on the catalyst surface could be evaluated by adding NB to this system. The inhibition effect of 100 mM NB on CIP degradation efficiency was much smaller than that of other quenchers, indicating that the main free radical on the catalyst surface was rather than •OH.
Figure 5

The degradation of CIP in the presence of MeOH (3 M), TBA (3 M), L-His (0.2 mM), phenol (0.1 M), NB (0.1 M) and (0.01 M) (a) and EPR spectrum of solution, reaction conditions: [CIP]0 = 10 mg/L, [catalyst] = 0.2 g/L, [PMS] = 0.6 g/L, temperature = 25 °C, pH was not adjusted.

Figure 5

The degradation of CIP in the presence of MeOH (3 M), TBA (3 M), L-His (0.2 mM), phenol (0.1 M), NB (0.1 M) and (0.01 M) (a) and EPR spectrum of solution, reaction conditions: [CIP]0 = 10 mg/L, [catalyst] = 0.2 g/L, [PMS] = 0.6 g/L, temperature = 25 °C, pH was not adjusted.

Close modal

It has been reported that the –OH group was essential in the heterogeneous activation of PMS by metal-based catalysts (Liu et al. 2015). In order to determine the role of the –OH group on the catalyst surface, KH2PO4 was added to the system as a masking agent to study its effect on CIP degradation. KH2PO4 could easily combine with transition metals to mask the –OH group on the catalyst surface (Khan et al. 2017), preventing the interaction of the –OH group with . After 10.0 mM was added to the ε-MnO2/PMS system, the degradation efficiency of CIP decreased from 91.9 to 15.9%, which indicated that the –OH group on the catalyst surface played an important role in controlling the interaction between peroxides and catalysts.

To further determine the production of reactive oxygen species, EPR tests were performed on ε-MnO2/PMS system and PMS system alone using DMPO and TEMP as spin trapping agents. and •OH could form stable adducts with DMPO, which could be detected by electron paramagnetic resonance (EPR) spectroscopy, so DMPO was used as a trapping agent for and •OH, and TEMP was a spin capture agent for 1O2 (Xu et al. 2021). As shown in Figure 5(b), the mixed signals of DMPO- and DMPO-•OH and the triplex signal of TEMP-1O2 (1:1:1) were observed in the spectrum of the ε-MnO2/PMS system, but no obvious signal was detected when only PMS existed indicating that PMS produced , •OH and 1O2 in solution under the activation of ε-MnO2, and the quantity of was higher than that of •OH and 1O2 in terms of relative spectral intensity.

By comparing the XPS spectra before and after the ε-MnO2 reaction (Figure 3(a)–3(c)), it was found that the manganese content of different valence states changed, Mn(III) decreased with Mn(Ⅳ) increasing, Mn(III)/Mn(Ⅳ) decreased from 5.85 to 4.82. The relative content of Oads also decreased, and the Oads of ε-MnO2 participated in the reaction. The proportion of Mn and O species changed slightly before and after ε-MnO2 reaction, indicating that ε-MnO2 could regenerate itself and maintain high activity while participating in the reaction.

Based on the above experimental results, the reaction mechanism of ε-MnO2-activated PMS to degrade CIP was speculated in Figure 6. On the one hand, PMS combined with ε-MnO2 in the reaction process, reducing Mn4+ to Mn3+, and simultaneously generating SO5 (Equation (3)), and Mn3+ reacted with PMS to produce . In addition, the hydroxyl group on the catalyst surface could act as a bridge between the catalyst and PMS, which could promote the activation of PMS. First, the hydroxyl group combines with Mn4+ (Mn3+) to form Mn4+ (Mn3+) –OH and Mn4+ (Mn3+) –OH reacted with to form (Equation (12)), and then was decomposed to form (Equation (13)). could also react with OH to form •OH (Equation (14)), and •OH could be decomposed into 1O2 during the dissolution process (Equation (15)). These active substances in the solution eventually attacked the chemical bonds in CIP to make it decomposed and mineralized (Equation (16)). The oxidation–reduction reaction between Mn4+ and Mn3+ occurred in the catalytic process of MnO2 (Equations (3) and (4)), and the catalyst reacted and self-repaired simultaneously during the degradation process, which could effectively reduce its wastage and extend its working life. This result was of great significance to the actual water treatment engineering.
(12)
(13)
(14)
(15)
(16)
Figure 6

The mechanism of ε-MnO2 activating PMS to degrade CIP.

Figure 6

The mechanism of ε-MnO2 activating PMS to degrade CIP.

Close modal

An environmentally friendly catalytic material, ε-MnO2 was synthesized by the redox precipitation method and used to activate PMS to degrade the antibiotic CIP. The results showed that the CIP degradation rate could reach 93% within 30 min after the optimization of reaction parameters. The high activity of ε-MnO2 was related to its physicochemical properties of the surface, such as the specific surface area, the oxygen vacancy formed by the defect and the Mn(III)/Mn(Ⅳ) content ratio. After four cycles of degradation experiments, the efficiency decreased by 13%, which confirmed that the catalyst had good stability. The quenching experiment and EPR result showed that sulfate radicals, hydroxyl radicals and singlet oxygen jointly participated in the degradation of CIP. The main free radical was . The hydroxyl group on the surface of ε-MnO2 was helpful to the contact between PMS and the catalyst, which could promote the activation of PMS. The redox reaction between Mn4+ and Mn3+ occurred during the activation of PMS by ε-MnO2, which could realize the self-repair in situ and recycle of the catalyst, which provided the theoretical basis and technical support for the removal of antibiotics and had great potential for practical engineering application.

This study was financially supported by Gansu Province Higher Education Industry Support Plan Project (2022CYZC-40), Natural Science Foundation of Gansu Provincial Department of Science and Technology (22JR5RA316), Major Scientific and Technological Achievements Cultivation Project of Lanzhou Jiaotong University (2022CG07) and Tianyou Innovation Team of Lanzhou Jiaotong University (TY202005).

All relevant data are included in the paper or its Supplementary Information.

The authors declare there is no conflict.

Arif
M.
,
Liu
G.
,
Zia Ur Rehman
M.
,
Yousaf
B.
,
Ahmed
R.
,
Mian
M. M.
,
Ashraf
A.
,
Mujtaba Munir
M. A.
,
Rashid
M. S.
&
Naeem
A.
2022
Carbon dioxide activated biochar-clay mineral composite efficiently removes ciprofloxacin from contaminated water – reveals an incubation study
.
Journal of Cleaner Production
332
,
130079
.
Chen
Y.
,
Yan
J.
,
Ouyang
D.
,
Qian
L.
,
Han
L.
&
Chen
M.
2017
Heterogeneously catalyzed persulfate by CuMgFe layered double oxide for the degradation of phenol
.
Applied Catalysis A: General
538
,
19
26
.
Chun-Hui Shen
X. W. Z. F.
2019
Visible-light-driven activation of peroxymonosulfate for accelerating ciprofloxacin degradation using CeO2/CO3O4 p-n heterojunction photocatalysts
.
Chemical Engineering Journal
391, 123612.
Falyouna
O.
,
Maamoun
I.
,
Bensaida
K.
,
Tahara
A.
,
Sugihara
Y.
&
Eljamal
O.
2022
Encapsulation of iron nanoparticles with magnesium hydroxide shell for remarkable removal of ciprofloxacin from contaminated water
.
Journal of Colloid and Interface Science
605
,
813
827
.
Gao
F.
,
Li
Y.
&
Xiang
B.
2018
Degradation of bisphenol A through transition metals activating persulfate process
.
Ecotoxicology and Environmental Safety
158
,
239
247
.
Ghauch
A.
,
Baalbaki
A.
,
Amasha
M.
,
El Asmar
R.
&
Tantawi
O.
2017
Contribution of persulfate in UV-254 nm activated systems for complete degradation of chloramphenicol antibiotic in water
.
Chemical Engineering Journal
317
,
1012
1025
.
Gujar
S. K.
,
Divyapriya
G.
,
Gogate
P. R.
&
Nidheesh
P. V.
2023
Environmental applications of ultrasound activated persulfate/peroxymonosulfate oxidation process in combination with other activating agents
.
Critical Reviews in Environmental Science and Technology
53
,
780
802
.
Hao
S.
,
Yu
M.
,
Zhang
Y.
,
Abdelkrim
Y.
&
Qu
J.
2019
Hierarchical mesoporous cobalt silicate architectures as high-performance sulfate-radical-based advanced oxidization catalysts
.
Journal of Colloid and Interface Science
545
,
128
137
.
He
L.
,
Chen
H.
,
Wu
L.
,
Zhang
Z.
,
Ma
Y.
,
Zhu
J.
,
Liu
J.
,
Yan
X.
,
Li
H.
&
Yang
L.
2021a
Synergistic heat/UV activated persulfate for the treatment of nanofiltration concentrated leachate
.
Ecotoxicology and Environmental Safety
208
,
111522
.
Hu
X.
,
Hu
X.
,
Peng
Q.
,
Zhou
L.
,
Tan
X.
,
Jiang
L.
,
Tang
C.
,
Wang
H.
,
Liu
S.
,
Wang
Y.
&
Ning
Z.
2020
Mechanisms underlying the photocatalytic degradation pathway of ciprofloxacin with heterogeneous TiO2
.
Chemical Engineering Journal
380
,
122366
.
Huang
W.
,
Chen
J.
&
Zhang
J.
2018a
Removal of ciprofloxacin from aqueous solution by rabbit manure biochar
.
Environmental Technology
41, 1380–1390.
Huang
Y.
,
Tian
X.
,
Nie
Y.
,
Yang
C.
&
Wang
Y.
2018b
Enhanced peroxymonosulfate activation for phenol degradation over MnO2 at pH 3.5–9.0 via Cu(ii) substitution
.
Journal of Hazardous Materials
360
,
303
310
.
Huang
J.
,
Dai
Y.
,
Singewald
K.
,
Liu
C.
,
Saxena
S.
&
Zhang
H.
2019
Effects of MnO2 of different structures on activation of peroxymonosulfate for bisphenol A degradation under acidic conditions
.
Chemical Engineering Journal (Lausanne, Switzerland: 1996)
370
,
906
915
.
Issaka
E.
,
Amu-Darko
J. N.
,
Yakubu
S.
,
Fapohunda
F. O.
,
Ali
N.
&
Bilal
M.
2022
Advanced catalytic ozonation for degradation of pharmaceutical pollutants-a review
.
Chemosphere
289
,
133208
.
Jing
D.
,
Yongjian
G.
,
Chaoqun
T.
,
Hongyu
W.
,
Qingsong
L.
,
Shiqing
Z.
&
Kejia
Z.
2017
Degradation of ciprofloxacin using α-MnO2 activated peroxymonosulfate process: effect of water constituents, degradation intermediates and toxicity evaluation
.
Chemical Engineering Journal
330
, 1390–1400.
Khan
A.
,
Wang
H.
,
Liu
Y.
,
Jawad
A.
,
Ifthikar
J.
,
Liao
Z.
,
Wang
T.
&
Chen
Z.
2017
Highly efficient α-mn2o3@α-MnO2-500 nanocomposite for peroxymonosulfate activation: comprehensive investigation of manganese oxides
.
Journal of Materials Chemistry A.
6, 1590–1600.
Lin
C.
&
Tsai
C.
2019
Degradation of isopropyl alcohol using UV and persulfate in a large reactor
.
Separation and Purification Technology
209
,
88
93
.
Nasrollahi
N.
,
Vatanpour
V.
&
Khataee
A.
2022
Removal of antibiotics from wastewaters by membrane technology: limitations, successes, and future improvements
.
Science of the Total Environment
838
,
156010
.
Neta
P.
,
Huie
R. E.
&
Huie
P. N. A. R.
1988
Rate constants for reactions of inorganic radicals in aqueous solution
.
Journal of Physical and Chemical Reference Data
17
,
1027
1284
.
Rama
P.
,
Roggy
D.
,
Satinder
K. B.
,
Patrick
D.
,
Francois
P.
,
Mausam
V.
&
Rao
Y. S.
2018
Activation of persulfate by homogeneous and heterogeneous iron catalyst to degrade chlortetracycline in aqueous solution
.
Chemosphere
207
,
543
551
.
Salari
M.
,
Nikoo
M. R.
,
Al-Mamun
A.
,
Rakhshandehroo
G. R.
&
Mooselu
M. G.
2022
Optimizing Fenton-like process, homogeneous at neutral pH for ciprofloxacin degradation: comparing RSM-CCD and ANN-GA
.
Journal of Environmental Management
317
,
115469
.
Saputra
E.
,
Muhammad
S.
,
Sun
H.
,
Ang
H. M.
,
Tadé
M. O.
&
Wang
S.
2013a
Different crystallographic one-dimensional MnO2 nanomaterials and their superior performance in catalytic phenol degradation
.
Environmental Science & Technology
47
,
5882
5887
.
Saputra
E.
,
Muhammad
S.
,
Sun
H.
,
Ang
H.
,
Tadé
M. O.
&
Wang
S.
2013b
A comparative study of spinel structured Mn3O4, Co3O4 and Fe3O4 nanoparticles in catalytic oxidation of phenolic contaminants in aqueous solutions
.
Journal of Colloid and Interface Science
407
,
467
473
.
Wang
Y.
,
Indrawirawan
S.
,
Duan
X.
,
Sun
H.
,
Ang
H. M.
,
Tadé
M. O.
&
Wang
S.
2015b
New insights into heterogeneous generation and evolution processes of sulfate radicals for phenol degradation over one-dimensional α-MnO2 nanostructures
.
Chemical Engineering Journal
266
,
12
20
.
Wilkinsan
F.
1981
Rate constants for the decay and reactions of the lowest electronically excited singlet state of molecular oxygen in solution
.
Journal of Physical and Chemical Reference Data
4
(10), 809–999.
Xu
X.
,
Zhang
Y.
,
Zhou
S.
,
Huang
R.
,
Huang
S.
,
Kuang
H.
,
Zeng
X.
&
Zhao
S.
2021
Activation of persulfate by MnOOH: degradation of organic compounds by nonradical mechanism
.
Chemosphere
272
,
129629
.
Yan
S.
,
Xiong
W.
,
Xing
S.
,
Shao
Y.
,
Guo
R.
&
Zhang
H.
2017
Oxidation of organic contaminant in a self-driven electro/natural maghemite/peroxydisulfate system: efficiency and mechanism
.
Science of the Total Environment
599–600
,
1181
1190
.
Yang
Q.
,
Hou
P.
,
Bai
S.
,
Wang
M.
&
Cheng
H.
2001
Adsorption and capillarity of nitrogen in aggregated multi-walled carbon nanotubes
.
Chemical Physics Letters
345
,
18
24
.
Zhang
L.
,
Zhao
X.
,
Niu
C.
,
Tang
N.
,
Guo
H.
,
Wen
X.
,
Liang
C.
&
Zeng
G.
2019
Enhanced activation of peroxymonosulfate by magnetic Co3MnFeO6 nanoparticles for removal of carbamazepine: efficiency, synergetic mechanism and stability
.
Chemical Engineering Journal
362
,
851
864
.
Zhen
Y.
,
Sun
Z.
,
Jia
Z.
,
Liu
C.
,
Zhu
S.
,
Li
X.
,
Wang
W.
&
Ma
J.
2023
Facile preparation of α-MnO2 nanowires for assembling free-standing membrane with efficient Fenton-like catalytic activity
.
Chinese Chemical Letters
34
,
107664
.
This is an Open Access article distributed under the terms of the Creative Commons Attribution Licence (CC BY 4.0), which permits copying, adaptation and redistribution, provided the original work is properly cited (http://creativecommons.org/licenses/by/4.0/).

Supplementary data