Next Article in Journal
A Hierarchical Nano to Micro Scale Modelling of 3D Printed Nano-Reinforced Polylactic Acid: Micropolar Modelling and Molecular Dynamics Simulation
Previous Article in Journal
Novel Biological-Based Strategy for Synthesis of Green Nanochitosan and Copper-Chitosan Nanocomposites: Promising Antibacterial and Hematological Agents
Previous Article in Special Issue
Advanced Removal of Dyes with Tuning Carbon/TiO2 Composite Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Antibiotic Degradation via Fenton Process Assisted by a 3-Electron Oxygen Reduction Reaction Pathway Catalyzed by Bio-Carbon–Manganese Composites

by
Edgar Fajardo-Puerto
1,
Abdelhakim Elmouwahidi
1,*,
Esther Bailón-García
1,
María Pérez-Cadenas
1,2,
Agustín F. Pérez-Cadenas
1 and
Francisco Carrasco-Marín
1
1
UGR-Carbon, Materiales Polifuncionales Basados en Carbono, Dpto. de Química Inorgánica, Unidad de Excelencia de Química Aplicada a Biomedicina y Medioambiente, Universidad de Granada (UEQ-UGR), 18071 Granada, Spain
2
Dpto. Química Inorgánica y Química Técnica, Facultad de Ciencias, Universidad Nacional de Educación a Distancia (UNED), Av. de Esparta s/n, Las Rozas de Madrid, 28232 Madrid, Spain
*
Author to whom correspondence should be addressed.
Nanomaterials 2024, 14(13), 1112; https://doi.org/10.3390/nano14131112
Submission received: 11 May 2024 / Revised: 16 June 2024 / Accepted: 22 June 2024 / Published: 28 June 2024

Abstract

:
Bio-carbon–manganese composites obtained from olive mill wastewater were successfully prepared using manganese acetate as the manganese source and olive wastewater as the carbon precursor. The samples were characterized chemically and texturally by N2 and CO2 adsorption at 77 K and 273 K, respectively, by X-ray photoelectron spectroscopy (XPS) and X-ray diffraction. Electrochemical characterization was carried out by cyclic voltammetry (CV) and linear sweep voltammetry (LSV). The samples were evaluated in the electro-Fenton degradation of tetracycline in a typical three-electrode system under natural conditions of pH and temperature (6.5 and 25 °C). The results show that the catalysts have a high catalytic power capable of degrading tetracycline (about 70%) by a three-electron oxygen reduction pathway in which hydroxyl radicals are generated in situ, thus eliminating the need for two catalysts (ORR and Fenton).

1. Introduction

The contamination of water with antibiotics has become a global concern due to the associated risks to both the environment and public health. This includes the emergence of super-resistant bacteria that can cause new diseases and are able to tolerate treatments that were previously effective [1,2]. In general, the most widely used antibiotics worldwide are tetracycline (TC), quinolone, aminoglycoside, macrolide, and sulfonamide [3] TC, due to its low cost and broad antimicrobial spectrum, is used in both humans and animals; however, due to its low metabolization, it has even been found in drinking water [4]. Conventional methods for degrading this type of contaminant, such as biological processes, filtration, coagulation, flocculation, sedimentation, adsorption, and membrane processes, have proven to be inadequate or inefficient. Therefore, it is necessary to explore new processes capable of partially or completely mineralizing these molecules [5,6,7,8,9]
Advanced oxidation processes (AOPs) are a highly regarded option for reducing various persistent pollutants in water due to the generation of hydroxyl radicals (·OH) that have a high oxidative capacity (2.8 V vs. RH) capable of partially or completely mineralizing a broad range of pollutants. These radicals are produced on-site when using AOPs, which provides a powerful, alternative method for eliminating persistent pollutants in water [10,11,12]. The electro-Fenton (EF) process is a notable AOP due to its high rate of electro-generated H2O2, low production of iron sludge, and environmentally friendly nature [13]. In general, the EF process is based on the in situ production of H2O2 from the oxygen reduction reaction via two electrons (ORR 2e-) (Equation (1)) at the cathode, generation of ·OH by reacting the Fenton-like catalyst (usually Fe2+) with H2O2 (Equation (2)), and, finally, regeneration of Fe2+ in the cathode (Equation (3)) and by interaction with H2O2 (Equation (4)) [14,15,16,17].
O 2 + 2 H + + 2 e H 2 O 2
H 2 O 2 + F e 2 + F e 3 + + H O +   · O H
F e 3 + + e F e 2 +
H 2 O 2 + F e 3 + F e 2 + +   · O H 2 + H 2 O
As can be seen, the cathode material is a determining factor in the efficiency of the EF process [18]. So, recently, different types of materials have been used among which noble metals [19,20], metal oxides [21,22], and carbonaceous materials stand out [23,24,25]. However, carbonaceous materials stand out mainly for their low cost, good stability, and abundance [26,27]. Various carbonaceous materials, in order to reduce costs, have been made from different biomasses such as bamboo [28], black soybean [29], the inner layer of torreya grandis [30], waste wood (balsa wood [31], water hyacinth [32], cellulose [33], and, recently, in relation to the circular economy, sewage sludge has been used as a source of biocarbon [34]. However, the main drawback with this type of carbon material is the low activity and slow kinetics [35,36,37], so it is necessary to find a strategy to overcome these limitations and ameliorate the catalytic performances of carbon materials. To this end, the formation of a composite that exploits the advantages of carbon materials will be advantageous for use in electrocatalysis due to the synergistic effect between the metal catalyst and the catalytic support. In addition, the structure, composition, and concentration of the catalyst and support can be precisely modulated by controlling the stoichiometry of the precursors [38,39,40].
In general, doping with transition metals has resulted in a strategy to improve the ORR activity of carbonaceous materials, since, in addition to this, transition metal ions can serve as electron donors to generate ·OH from H2O2 [41]. Among the transition metals, MnOx oxides stand out, due to their different applications such as catalysis, energy storage, and, specifically, in electrochemical advanced oxidation processes (EAOP) [42]. MnO2 has been shown to have the ability to dissociate H2O2 to ·OH at a faster reaction rate [43]. However, the application of carbonaceous materials doped with Mn as possible bifunctional catalysts for the generation direct of ·OH by a three-electron pathway has not yet been studied.
In the present work, a biocarbon obtained from “alpechín” olive mill wastewater generated from the production of olive oil [44] was synthesized through a chemical activation process and treated with Manganese with three different methods and loadings. The synthesized composite materials were evaluated in an environmental remediation process as tetracycline degradation by the electro-Fenton process.

2. Experimental

2.1. Bio-Carbon–Manganese Composites Preparation

A bio-carbon-denominated CK2 was the base for the development of the composite series. The CK2 was prepared from olive mill wastewater by chemical activation with potassium hydroxide (KOH), the preparation details and textural characterization are reported in previous works [45]. Manganese acetate ((CH3COO)2Mn·4H2O, Sigma Aldrich, St. Louis, MI, USA) was used as a manganese precursor. Different amounts of the manganese acetate were dissolved in water and were added dropwise onto the corresponding amount of bio-carbon.
After impregnation, the prepared mixture was treated by three different methods. The composites were noted as CK2-Mn-X-Y; X corresponds to the method used for the preparation and Y traduces the theoretical manganese percentage.
The first sample was calcined (after impregnation with 10% of manganese) at 330 °C for 1 h under nitrogen gas (300 mL min−1) and 1 h under CO2 flow. The corresponding sample was noted as CK2-Mn-1-10. The second sample was prepared by adding H2O2 (25 mL g−1 carbon) to the impregnated samples and, after that, the samples were calcined at 330 °C for 2 h under 300 mL min−1 of nitrogen. The corresponding composites were noted as CK2-Mn-2-10, CK2-Mn-2-25, and CK2-Mn-2-60. The last sample was prepared by a few modifications of the second method, the impregnated sample was calcined at 330 °C for 2 h under 300 mL min−1 of nitrogen before being treated with H2O2 (25 mL g−1 carbon), CK2-Mn-3-10.

2.2. Textural and Chemical Characterization

The textural characterization was obtained by using Autosorb equipment from Quantachrome system (Boynton Beach, FL, USA). N2 adsorption isotherm at −196 °C and CO2 at 0 °C were obtained. The B.E.T. and Dubinin–Radushkevich (DR) methods were applied to the nitrogen and carbon dioxide isotherms, respectively, to obtain the apparent surface area, total micropore volume (W0 with N2), narrow micropore volume (W0 with CO2), and mean micropore width (L0) [46,47,48]. The pore-size distribution (PSD) was determined by applying Quenched Solid Density Functional Theory (QSDFT) to the N2 adsorption isotherms, assuming slit-shaped pores.
The morphology of the samples was studied by scanning electron microscopy (SEM) in an AURIGA FIB-FESEM microscope provided by Carl Zeiss SMT (Jena, Germany). The inorganic content in the catalysts was determined by thermogravimetric analysis (TGA) with a Mettler-Toledo TGA/DSC1 thermogravimetric analyzer (Greifensee, Switzerland). The TGA was carried out in air flow from 20 °C to 900 °C with heating rate of 10 °C min−1. The crystallinity of manganese oxide particles of all CK2-Mn-X-Y composites was investigated by X-ray diffraction Bruker D8 (Karlsruhe, Germany) advance with Cu Kα radiation.
The surface chemistry of the samples was obtained by X-ray photoelectron spectroscopy using ESCA 5701 from Physical Electronics (PHI, Chanhassen, MN, USA) system (equipped with MgKα anode, model PHI 04-548. X-ray source (hν = 1253.6 eV) and hemispherical electron energy analyzer. For the analysis of the XPS peaks, the C1s peak position was set at 284.6 eV and used as reference to locate the other peaks. The XPS peaks were fitted using Gaussian–Lorentzian peak shapes and a Shirley background through the least squares method, utilizing XPS peaks 4.1.

2.3. Electrochemical Characterization

The prepared CK2-Mn-X-Y composites were electrochemically characterized in a Biologic VMP Multichannel potentiostat (Grenoble, France) using a Rotating Ring-Disk Electrode (RRDE) (Metrohm AUTOLAB RDE-2, 3 mm Glassy Carbon tip, Utrech, Holland) where the sample was deposited as working electrode. The working electrode was prepared by depositing on the RRDE tip 20 µL of an ink, consisting of 5 mg of sample dispersed in 1 mL of a Nafion solution (1/9 v:v Nafion 5%/water), and dried under infrared radiation.
The activity for ORR was studied using the RRDE technique by using an Autolab electrochemical system, Metrohm (Utrech, Holland), associated with a compact potentiostat/galvanostat (PGSTAT101). Electrochemical measurements were carried out in a standard three-electrode electrochemical cell at room temperature. Using a Pt sheet as a counter electrode and an Ag/AgCl reference electrode. The electrolyte was a solution of KOH 0.1 M prepared in deionized water. Before each electrochemical measurement, the electrolyte was saturated with N2 or O2 by purging the necessary gas into the KOH 0.1 M solution at room temperature for 30 min. Cyclic voltammetry (CV) was collected in the KOH 0.1 M solution saturated with N2 or O2 in a range of 0.40 to −0.80 V at 50 mV s−1 with the electrode at a rotational speed of 1000 rpm. Linear scanning voltammetry (LSV) for the ORR was carried out with a potential range between 0.40 V and −0.80 at 50 mV s−1 with the electrode at rotational speeds of 500, 1000, 1500, 2000, 2500, 3000, and 3500 rpm.
The number of electrons transferred and H2O2 selectivity were calculated using Equations (5) and (6), which are calculated from the data obtained in the RRDE during the experiment [49,50]
n = 4 × I D I D + I R N C
% H 2 O 2 = 200 × I R N C I D + I R N C
where I D and IR are the disk and ring current, respectively, and NC is the collection efficiency of the RRDE (0.249).
The current density (JK) was obtained by Koutecky–Levich equation (Equation (7)).
1 J = 1 J K + 1 B × w 1 / 2  
with B = 0.62 n × F × A × D 2 / 3 × v 1 / 6 × C , where F is Faraday’s constant (96,485,332 mC mol−1), n number of electrons transferred per oxygen molecule, A is the disc area of the RRDE (0.2475 cm2), D diffusion coefficient of oxygen (1.9 × 10−5 cm2 s−1), ν is kinematic viscosity (0.01 cm2 s−1), and, finally, C is the solubility of oxygen (1.2 × 10−6 mol cm−3) [51,52].

2.4. Electro-Fenton Processes

The electro-Fenton process was carried out using a standard three-electrode electrochemical cell with capacity for 150 mL of solution at room temperature. The TC concentration approached 40 mg L−1, using Na2SO4 [0.5 M] as a supporting electrolyte with continuous agitation. Potentiostat was maintained in potentiostatic mode at −0.6 V.
The working electrode was prepared by mixing 45 mg of CK2-Mn-X-Y with 8.5 mg of 60% PTFE, for subsequent drying at 100 °C for 12 h. Once dried, a paste was obtained that was deposited on a sheet of graphite (50 mg on each face). While the reference electrode was Ag/AgCl, and the counter electrode used was a platinum sheet. The working pH was 6.5, as is natural for the TC solution. TC concentrations in solution were determined by a UV–vis spectrophotometer at a wavelength of 356.5 nm.

3. Results and Discussions

3.1. Morphological and Textural Characterization

3.1.1. Morphology

The SEM micrographs (Figure 1) show the attack of the activator on the carbon surface. The microporous structure of the raw material is combined with the presence of mesopores formed by the reaction between KOH and carbon (A). After doping with manganese oxide, we can observe manganese oxide nanostructures that vary in morphology depending on the preparation method. Thus, we can see that in sample CK2-Mn-1-10, these nanostructures appear in the form of nanofilaments, whereas in samples CK2-Mn-2-10 and CK2-Mn-3-10, they appear in the form of spheres or hemispheres deposited on the bio-carbon support. This difference shows that the preparation method influences the final shape and size of the manganese oxide catalyst and consequently the electrocatalytic performance.

3.1.2. The Porosity and Surface Area of the Composites

Textural properties of the composites were examined by N2 adsorption–desorption isotherm measurements. Figure 2 shows the N2 adsorption–desorption isotherms of the CK2-Mn-X-Y samples. All samples show a hybrid-type I-IV isotherm, with a high N2 adsorption at low relative pressures and a hysteresis loop at intermediates ones, indicating the presence of micropores and mesopores [53]. Also, the samples CK2-Mn-3-10, CK2-Mn-1-10, and CK2-Mn-2-60 present a peak associated with capillary condensation. Note that the addition of Mn to the samples produces a blockage of the porosity in all cases; however, this porosity depletion depends on the final precursor decomposition treatment. From Figure 1, it is possible to observe that the H2O2 activation before carbonization (CK2-Mn-2-10) results in the better preservation of the porosity, due possibly to the chemical attack in the porous texture of the sample (CK2), which leads to the generation of new porosity. Finally, it is highlighted that the Mn amount directly affects the porosity; the higher the Mn content, the higher the porosity decrease, which is associated with the blockage of the pores of the carbon matrix by the manganese particles; this can be observed in the samples CK2-Mn-2 with different Mn charges (10, 25, and 60).
The B.E.T surface area (SBET), micropore surface area (Smicro), mesopore surface area (SDFT), and pore volume (VDFT) of the CK2-Mn-X-Y samples are listed in Table 1.
The same results can be observed by analyzing Table 1. The original activated carbon presents a surface area as high as 1672 m2/g. This porosity decreases after the incorporation of Mn; however, this porosity decrease is less pronounced when the pristine carbon is treated with H2O2 before the manganese decomposition at 330 °C. The physical activation with CO2 results in a micropore volume lower than with the other methods (chemical activation) decreasing from 0.38 (CK2) to 0.29 cm3 g−1, indicating the blockage of the ultramicroporosity by the Mn nanoparticles or an opening of such porosity by the CO2 treatment, which is in agreement with the reported in the literature [54]. On the other hand, the decrease in narrow micropore volume and total micropore volume, of the samples CK2-Mn-2-25 and Ck-2-Mn-60, can be attributed to pore enlargement and blockage of these pores with Mn particles, which result in a decrease in B.E.T surface area, indicating a lower development of microporosity [55]. The sample CK2-Mn-3-10 has an L0(CO2) greater than 0.7 nm and a W0(N2) < W0(CO2), indicating a diffusional problem of N2, which translates into a bad activation degree, due to restrictions in the microporosity, while the other samples have W0(N2) > W0(CO2), which is typical of an adequate activation degree.

3.2. TGA

Figure 3 shows the TGA analysis of the samples CK2-Mn-1-10, CK2-Mn-2-25, and CK2-Mn-2-60. From the final residual mass, the amount of manganese oxide was determined. The percentage of manganese oxide in the samples was 15.51, 30.02, and 66 wt.% for CK2-Mn-1-10, CK2-Mn-2-25, and CK2-Mn-2-60, respectively. The theoretical and TGA content of manganese in the samples are close, which indicates a good doping method.

3.3. X-ray Diffraction

Figure 4 shows the XRD diffractograms of the CK2-Mn-X-Y samples, where we can observe different peaks associated with some species of Mn, indicating the presence of manganese in different oxidation states. It is possible to determine the peaks to approximately 18°, 29°, 32°, 36°, 38°, 44°, 58°, 60°, and 65°, associated with the planes (101), (112), (103), (211), (004), (220), (321), (224), and (400) characteristics of spinel structure Mn3O4 (CK2-Mn-3-10; ICSD card number 76088) [56,57,58]. While the peaks closer to 35°, 40°, and 59° correspond to the crystalline planes (111), (100), and (220) of MnO (CK2-Mn-3-60; ICSD card number 657304) [59,60,61]. The samples CK2-Mn-1-10 and CK2-Mn-2-10 present spectra corresponding to practically amorphous specimens, so raising the dispersion of manganese particles was better in these two materials. By increasing the manganese content, a new crystalline phase emerges, finding well-defined peaks of MnO crystalline phase form for the CK2-Mn-2-25 and CK2-Mn-2-60 samples.
From these results, it would be expected that better catalytic results will be obtained with the samples with 10% of Mn, due to the presence of Mn3O4, a species considered active for promoting ·OH generation by ORR [62].

4. Surface Chemistry of Composites

4.1. XPS Spectra

Figure 5 and Figure 6 show the XPS spectra for the signals C1s, O1S, and Mn2p. The surface composition of the samples was determined from the deconvolution of the XPS spectra, and the results are summarized in Table S1 (Supplementary Materials). The peaks obtained from the deconvolution of the C1s spectrum were related to C=C (284.6 eV), C-O (285.7 eV), C=O (287 eV), O-C=O (288.4 eV), CO2 or π-π* bonds and plasmon (290 and 291.6 eV) [63,64,65,66,67]. The peaks from O1s, were deconvoluted in O-Mn (530.0 eV), O=C (531.4 eV), and O-C (533.4 eV) [68,69]. From Mn2p the peaks detected were Mn2+ (641.6 eV) and Mn3+ (643.4 eV) [70,71,72,73,74].
From the results obtained, it can be concluded that all samples present similar composition profiles, with the same chemical species being found in all cases (see Table S1 Supplementary Materials). However, several significant facts are worth noting. The first one is the increase in the C peak width at 284.6 eV, which is indicative of an increase in the defects present in the graphitic crystals or a decrease in the size of the microcrystals, both facts will influence the electrical conductivity of the catalysts. Secondly, there is a noticeable increase in the O content when Mn is introduced into the catalysts, this is due to the oxidation produced by the reduction of Manganese acetate during the heat treatment. With respect to the Mn spectrum, two clearly differentiated species Mn2+ and Mn3+ are detected in all samples. However, a clear difference appears for sample CK2-Mn-3-10 for which the Mn3+/Mn2+ ratio is higher than for the rest of the samples, which is indicative of the formation of Mn3O4 due to the oxidation treatment carried out with H2O2, which agrees with the data obtained by XRD. For the rest of the samples, this ratio varies very little, between 0.64 and 0.71, indicating that part of the Mn2+ formed by the reduction treatment can be oxidized superficially to Mn3+ by exposure to air.

4.2. Electrochemical Characterization

The CV results are shown in Figure 7, where is possible to observe that all samples have ORR activity. The sample CK2-Mn-2-10 showed the higher capacitance attributed principally to the type of porosity and large specific surface [75]. These better behaviors could be related to the improved hydrophilicity of the surface due to the increase in the oxygen surface groups and the presence of MnOx, which improves the electrolyte–surface contact favoring the diffusion of the ions, and the optimization of the pore structure by the presence of MnOx. This leads to a decrease in the resistance of the electrodes, enhancing their performance for energy storage.
Based on its chemical composition and textural characteristics, sample CK2-Mn-2-10 is expected to have better ORR activity and higher capacitance compared to the other samples. There is no visible OPR peak when the electrolyte is saturated with N2. In contrast, a well-defined ORR peak is observed in the O2-saturated 0.1 M KOH solution, demonstrating excellent electrocatalytic activity for ORR. More interestingly, all CV curves (both N2 and O2 saturated) show strong redox peaks, the peak at ca. 0 V vs. Ag/AgCl was attributed to Mn2+/Mn3+ redox because the area of the peak at ca. 0 V vs. Ag/AgCl increases with greater Mn concentration (Mn content changes from 10 to 60) [76], and especially with the amount of Mn2+ and Mn3+ present in the sample.
In this line, Table 2 shows that the highest JK value was obtained with the pre-activated sample with H2O2 (CK2-Mn-2-10), which possibly is due in one part to the well-dispersed manganese phase and also to the highest surface area, which allows more current to pass into the matrix, for the same could explain the JK value lower of CK2-Mn-2-25 and CK2-Mn-2-60.
Figure 8 and Figure 9 show the correlation between Jk and the Mn3+/Mn2+ amount and SB.E.T., respectively. The correlation coefficient of R2 was closer to 1 in SB.E.T. (0.98) than with Mn3+/Mn2+ amount (0.94), indicating that although the manganese can be an active site for the ORR, the current density is affected principally for the SB.E.T. [77]. In this case, we can say that physical activation gives as a result a higher surface area that is traduced in a better JK, which is an important parameter for the different electrochemical applications.
Figure 10 shows the number of electrons transferred and selectivity to H2O2. Where it is possible to observe that the presence of Mn increases the number of electrons transferred, which is favorable for the ORR 4e; however, for the case of H2O2 generation, it is necessary that the values are closer to two electrons, with a high selectivity to H2O2; for this, it is possible to suppose that the best catalyst for electro-Fenton is obtained by using the samples CK2-Mn-1-10 and CK2-Mn-3-10, which have higher amounts of manganese on the external surface area compared to the sample CK2-Mn-2-10:14.9 and 16.6, respectively, and also, a higher surfaces area compared to other samples with higher amounts of manganese CK2-Mn-2-25 and CK2-Mn-2-60. In this case, we can observe that both the amount of Mn and the surface porosity could be important parameters for the number of electrons transferred. Nevertheless, according to DRX results, with the increase from 10 to 60 wt.% of Mn on the same series, significant chemical changes occurred: sample CK2-Mn-2-10 only could show some Mn3O4 phases, while the presence of MnO appears in sample CK2-Mn-2-25 and in sample CK2-Mn-2-60 as the main phase. On the other hand, sample CK2-Mn-3-10, which shows higher and better-defined peaks of Mn3O4, is the one that produces a larger amount of H2O2 and shows an n-value closer to three.

4.3. Electro Fenton

Since the presence of an MnO phase does not favor the desired ORR performance, only samples with a 10 wt.% of Mn were checked in the electro-Fenton tests. In this way, Figure 9 shows the tetracycline degradation with the above-mentioned samples as well as with graphite only (an electrode-support material on which the prepared samples are pasted for the electro-Fenton tests).
Additionally, were made electro-Fenton tests in N2 with H2O2 and N2 without H2O2 (Figure 11), with the objective of evaluating the effect of hydrogen peroxide and of the support.
The TC degradation in the presence only of O2 followed an order of CK2-Mn-1-10 > CK2 > CK2-Mn-3-10 > Graphite > CK2-Mn-2-10, whereby the sample CK2-Mn1-10 was used to evaluate its activity in presence of N2 with H2O2. The TC degradation in the presence of N2 showed a lower value compared with O2 (Figure 12), with this experiment having established that O2 presence has an important effect on the system, much more than the direct addition of H2O2 and obviously more than with only the presence of N2 in the total absence of O2 and H2O2.
These results could hide a probable three electrons ORR (Equation (8)) pathway with the direct formation of hydroxyl radicals (·OH), which enhances tetracycline degradation, as is observed in Figure 12 by comparing the effect of H2O2 addition vs. O2 bubbling [78].
O 2 + 3 e + 2 H +   · OH + H O
However, it is not possible to dismiss the route by electro-Fenton traditional with the ORR pathway with two electrons with an initial reduction of O2 to H2O2 (Equation (9) [79].
O 2 + 2 e + 2 H + H 2 O 2
H 2 O 2 + e + H + H 2 O +   · OH
In this process, the Mn3+/Mn2+ redox cycle in the spinel structure is crucial since electrons can be provided. Moreover, the following way would be working in parallel:
M n 2 + + H 2 O 2 M n 3 + +   · OH + H O
M n 3 + + e M n 2 +
However, with our results, it is proposed that the surface area B.E.T has a more significant effect in TC degradation (R2 > 0.6) by affecting the value of Jk, which could be attributed to greater degree of access to the active sites.
In fact, it seems that a relationship between the Mn2+ values and tetracycline degradation occurs: CK2-Mn-1-10 (8.42 Mn2+; 70% TC removal), CK2-Mn-3-10 (5.48 Mn2+; 47% TC removal), and CK2-Mn-2-10 (4.40Mn2+; 39% TC removal) (Figure 13). The relevance of Mn2+ can be explained by the interaction between Mn2+ and H2O2, which can generate ·OH radicals (Fenton traditional), and by the possible direct generation of ·OH radicals (ORR 3e proposed) with Mn3+, which is proposed as active site for this route. We propose that a good relationship between the Mn2+ percentage and the surface area is key in the catalytic activity for degrading the tetracycline. So, we can conclude that the tetracycline degradation depends on Jk, which depends on the surface area and the percentage of manganese, which interacts with H2O2, generating ·OH radicals.

5. Conclusions

Two series of bio-carbon–manganese composites were prepared with bifunctional behavior in the electro-Fenton process: H2O2 generation from ORR, and hydroxyl radical formation. Moreover, the hydroxyl radicals can be formed in two different ways, (i) by H2O2 decomposition on Mn phases as heterogeneous Fenton, or (ii) by H2O2 reduction in one-pot way directly from the ORR following a pathway or three electrons. Moreover, the current density during the ORR is directly proportional to the surface area of these materials. The development of the Mn3O4 phase by a proper preparation method in these composites seems to be the clue that both mentioned processes can take place, as a consequence of the Mn3+/Mn2+ redox cycles that can occur in the spinel structure. This bifunctional behavior makes possible the antibiotic degradation in water solutions by a process in which H2O2 does not need to be added because it is in situ produced in such any way.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano14131112/s1, Table S1. Compositional characterization by XPS analysis.

Author Contributions

Methodology, A.F.P.-C.; Validation, M.P.-C.; Formal analysis, F.C.-M.; Investigation, E.F.-P. and A.E.; Writing—original draft, E.F.-P. and A.E.; Writing—review & editing, E.B.-G. and M.P.-C.; Supervision, E.B.-G., A.F.P.-C. and F.C.-M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Consejería de Universidad, Investigación e Innovación and ERDF Andalusia Program 2021–2027, under Grant C-EXP-247-UGR23, and by MICIU/AEI/10.13039/501100011033/ and “ERDF A way of making Europe”, under Project PID2021-127803OB-I00.

Data Availability Statement

Data are contained within the article and supplementary materials.

Acknowledgments

A. Elmouwahidi acknowledges the support of the grant María Zambrano (RD 289/2021), funded by the European Union—Next Generation EU programme. E. Fajardo-Puerto acknowledged “MINCIENCIAS” for supporting his PhD studies. E. Bailón-García is acknowledged to MICINN for her postdoctoral fellowship (RYC2020-029301-I). M. Pérez-Cadenas acknowledges to UNED the economic support to the program of the “Ayudas para la recualificación del sistema universitario español 2021–2023”.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Garza-Campos, B.; Morales-Acosta, D.; Hernández-Ramírez, A.; Guzmán-Mar, J.L.; Hinojosa-Reyes, L.; Manríquez, J.; Ruiz-Ruiz, E.J. Air Diffusion Electrodes Based on Synthetized Mesoporous Carbon for Application in Amoxicillin Degradation by Electro-Fenton and Solar Photo Electro-Fenton. Electrochim. Acta 2018, 269, 232–240. [Google Scholar] [CrossRef]
  2. Ahmadi, A.; Zarei, M.; Hassani, A.; Ebratkhahan, M.; Olad, A. Facile Synthesis of Iron(II) Doped Carbonaceous Aerogel as a Three-Dimensional Cathode and Its Excellent Performance in Electro-Fenton Degradation of Ceftazidime from Water Solution. Sep. Purif. Technol. 2021, 278, 119559. [Google Scholar] [CrossRef]
  3. Pan, Y.; Zhang, Y.; Zhou, M.; Cai, J.; Tian, Y. Enhanced Removal of Antibiotics from Secondary Wastewater Effluents by Novel UV/Pre-Magnetized Fe0/H2O2 Process. Water Res. 2019, 153, 144–159. [Google Scholar] [CrossRef] [PubMed]
  4. Qin, J.; Ruan, Y.; Yi, L.; Sun, H.; Qi, Q.; Zhao, L.; Xiong, Y.; Wang, J.; Fang, D. Tetracycline (TC) Degradation via Hydrodynamic Cavitation (HC) Combined Fenton’s Reagent: Optimizing Geometric and Operation Parameters. Chem. Eng. Process.-Process Intensif. 2022, 172, 108801. [Google Scholar] [CrossRef]
  5. Du, X.; Fu, W.; Su, P.; Zhang, Q.; Zhou, M. FeMo@porous Carbon Derived from MIL-53(Fe)@MoO3 as Excellent Heterogeneous Electro-Fenton Catalyst: Co-Catalysis of Mo. J. Environ. Sci. (China) 2023, 127, 652–666. [Google Scholar] [CrossRef] [PubMed]
  6. Yang, C.; Fan, Y.; Shang, S.; Li, P.; Li, X. yan Fabrication of a Permeable SnO2-Sb Reactive Anodic Filter for High-Efficiency Electrochemical Oxidation of Antibiotics in Wastewater. Environ. Int. 2021, 157. [Google Scholar] [CrossRef] [PubMed]
  7. Akbari, M.Z.; Xu, Y.; Lu, Z.; Peng, L. Review of Antibiotics Treatment by Advance Oxidation Processes. Environ. Adv. 2021, 5, 100111. [Google Scholar] [CrossRef]
  8. Yuan, Q.; Qu, S.; Li, R.; Huo, Z.Y.; Gao, Y.; Luo, Y. Degradation of Antibiotics by Electrochemical Advanced Oxidation Processes (EAOPs): Performance, Mechanisms, and Perspectives. Sci. Total Environ. 2023, 856, 159092. [Google Scholar] [CrossRef] [PubMed]
  9. Wang, J.; Zhuan, R. Degradation of Antibiotics by Advanced Oxidation Processes: An Overview. Sci. Total Environ. 2020, 701, 135023. [Google Scholar] [CrossRef]
  10. Midassi, S.; Bedoui, A.; Bensalah, N. Efficient Degradation of Chloroquine Drug by Electro-Fenton Oxidation: Effects of Operating Conditions and Degradation Mechanism. Chemosphere 2020, 260, 127558. [Google Scholar] [CrossRef]
  11. Cao, P.; Zhao, K.; Quan, X.; Chen, S.; Yu, H. Efficient and Stable Heterogeneous Electro-Fenton System Using Iron Oxides Embedded in Cu, N Co-Doped Hollow Porous Carbon as Functional Electrocatalyst. Sep. Purif. Technol. 2020, 238, 116424. [Google Scholar] [CrossRef]
  12. Le, T.T.; Hoang, V.C.; Zhang, W.; Kim, J.M.; Kim, J.; Moon, G.H.; Kim, S.H. Mesoporous Sulfur-Modified Metal Oxide Cathodes for Efficient Electro-Fenton Systems. Chem. Eng. J. Adv. 2022, 12, 100371. [Google Scholar] [CrossRef]
  13. Puga, A.; Rosales, E.; Pazos, M.; Sanromán, M.A. Prompt Removal of Antibiotic by Adsorption/Electro-Fenton Degradation Using an Iron-Doped Perlite as Heterogeneous Catalyst. Process Saf. Environ. Prot. 2020, 144, 100–110. [Google Scholar] [CrossRef]
  14. Arellano, M.; Oturan, N.; Oturan, M.A.; Pazos, M.; Sanromán, M.Á.; González-Romero, E. Differential Pulse Voltammetry as a Powerful Tool to Monitor the Electro-Fenton Process. Electrochim. Acta 2020, 354. [Google Scholar] [CrossRef]
  15. Ganzenko, O.; Trellu, C.; Oturan, N.; Huguenot, D.; Péchaud, Y.; van Hullebusch, E.D.; Oturan, M.A. Electro-Fenton Treatment of a Complex Pharmaceutical Mixture: Mineralization Efficiency and Biodegradability Enhancement. Chemosphere 2020, 253. [Google Scholar] [CrossRef] [PubMed]
  16. Chai, Y.; Qin, P.; Zhang, J.; Li, T.; Dai, Z.; Wu, Z. Simultaneous Removal of Fe(II) and Mn(II) from Acid Mine Wastewater by Electro-Fenton Process. Process Saf. Environ. Prot. 2020, 143, 76–90. [Google Scholar] [CrossRef]
  17. Deng, F.; Olvera-Vargas, H.; Zhou, M.; Qiu, S.; Sirés, I.; Brillas, E. Critical Review on the Mechanisms of Fe2+ Regeneration in the Electro-Fenton Process: Fundamentals and Boosting Strategies. Chem. Rev. 2023, 123, 4635–4662. [Google Scholar] [CrossRef] [PubMed]
  18. Wang, J.; Qin, J.; Liu, B.; Song, S. Reaction Mechanisms and Toxicity Evolution of Sulfamethoxazole Degradation by CoFe-N Doped C as Electro-Fenton Cathode. Sep. Purif. Technol. 2022, 298, 121655. [Google Scholar] [CrossRef]
  19. Verdaguer-Casadevall, A.; Deiana, D.; Karamad, M.; Siahrostami, S.; Malacrida, P.; Hansen, T.W.; Rossmeisl, J.; Chorkendorff, I.; Stephens, I.E.L. Trends in the Electrochemical Synthesis of H2O2: Enhancing Activity and Selectivity by Electrocatalytic Site Engineering. Nano Lett. 2014, 14, 1603–1608. [Google Scholar] [CrossRef]
  20. He, D.; Zhong, L.; Gan, S.; Xie, J.; Wang, W.; Liu, Z.; Guo, W.; Yang, X.; Niu, L. Hydrogen Peroxide Electrosynthesis via Regulating the Oxygen Reduction Reaction Pathway on Pt Noble Metal with Ion Poisoning. Electrochim. Acta 2021, 371, 137721. [Google Scholar] [CrossRef]
  21. Choe, Y.J.; Kim, J.; Byun, J.Y.; Kim, S.H. An Electro-Fenton System with Magnetite Coated Stainless Steel Mesh as Cathode. Catal. Today 2021, 359, 16–22. [Google Scholar] [CrossRef]
  22. Gholizadeh, A.M.; Zarei, M.; Ebratkhahan, M.; Hasanzadeh, A. Phenazopyridine Degradation by Electro-Fenton Process with Magnetite Nanoparticles-Activated Carbon Cathode, Artificial Neural Networks Modeling. J. Environ. Chem. Eng. 2021, 9, 104999. [Google Scholar] [CrossRef]
  23. Zarei, M.; Beheshti Nahand, F.; Khataee, A.; Hasanzadeh, A. Removal of Nalidixic Acid from Aqueous Solutions Using a Cathode Containing Three-Dimensional Graphene. J. Water Process Eng. 2019, 32, 100978. [Google Scholar] [CrossRef]
  24. Scialdone, O.; Galia, A.; Sabatino, S. Electro-Generation of H2O2 and Abatement of Organic Pollutant in Water by an Electro-Fenton Process in a Microfluidic Reactor. Electrochem. Commun. 2013, 26, 45–47. [Google Scholar] [CrossRef]
  25. Nair, K.M.; Kumaravel, V.; Pillai, S.C. Carbonaceous Cathode Materials for Electro-Fenton Technology: Mechanism, Kinetics, Recent Advances, Opportunities and Challenges. Chemosphere 2021, 269, 129325. [Google Scholar] [CrossRef] [PubMed]
  26. Ahmad, A.; Priyadarshini, M.; Yadav, S. The Potential of Biochar-Based Catalysts in Removal of Persistent Organic Pollutants from Wastewater: A Review. Chem. Eng. Res. Des. 2022, 187, 470–496. [Google Scholar] [CrossRef]
  27. Xin, S.; Huo, S.; Zhang, C.; Ma, X.; Liu, W.; Xin, Y.; Gao, M. Coupling Nitrogen/Oxygen Self-Doped Biomass Porous Carbon Cathode Catalyst with CuFeO2/Biochar Particle Catalyst for the Heterogeneous Visible-Light Driven Photo-Electro-Fenton Degradation of Tetracycline. Appl. Catal. B 2022, 305, 121024. [Google Scholar] [CrossRef]
  28. Zhou, W.; Li, F.; Su, Y.; Li, J.; Chen, S.; Xie, L.; Wei, S.; Meng, X.; Rajic, L.; Gao, J.; et al. O-Doped Graphitic Granular Biochar Enables Pollutants Removal via Simultaneous H2O2 Generation and Activation in Neutral Fe-Free Electro-Fenton Process. Sep. Purif. Technol. 2021, 262, 118327. [Google Scholar] [CrossRef]
  29. Hu, X.; Deng, Y.; Zhou, J.; Liu, B.; Yang, A.; Jin, T.; Fai Tsang, Y. N- and O Self-Doped Biomass Porous Carbon Cathode in an Electro-Fenton System for Chloramphenicol Degradation. Sep. Purif. Technol. 2020, 251, 117376. [Google Scholar] [CrossRef]
  30. Gao, M.; Wang, W.K.; Zheng, Y.M.; Zhao, Q.B.; Yu, H.Q. Hierarchically Porous Biochar for Supercapacitor and Electrochemical H2O2 Production. Chem. Eng. J. 2020, 402, 126171. [Google Scholar] [CrossRef]
  31. Deng, F.; Olvera-Vargas, H.; Garcia-Rodriguez, O.; Zhu, Y.; Jiang, J.; Qiu, S.; Yang, J. Waste-Wood-Derived Biochar Cathode and Its Application in Electro-Fenton for Sulfathiazole Treatment at Alkaline PH with Pyrophosphate Electrolyte. J. Hazard. Mater. 2019, 377, 249–258. [Google Scholar] [CrossRef] [PubMed]
  32. Liang, J.; Tang, D.; Huang, L.; Chen, Y.; Ren, W.; Sun, J. High Oxygen Reduction Reaction Performance Nitrogen-Doped Biochar Cathode: A Strategy for Comprehensive Utilizing Nitrogen and Carbon in Water Hyacinth. Bioresour. Technol. 2018, 267, 524–531. [Google Scholar] [CrossRef] [PubMed]
  33. Sun, C.; Chen, T.; Huang, Q.; Duan, X.; Zhan, M.; Ji, L.; Li, X.; Wang, S.; Yan, J. Biochar Cathode: Reinforcing Electro-Fenton Pathway against Four-Electron Reduction by Controlled Carbonization and Surface Chemistry. Sci. Total Environ. 2021, 754, 142136. [Google Scholar] [CrossRef] [PubMed]
  34. Gopinath, A.; Divyapriya, G.; Srivastava, V.; Laiju, A.R.; Nidheesh, P.V.; Kumar, M.S. Conversion of Sewage Sludge into Biochar: A Potential Resource in Water and Wastewater Treatment. Environ. Res. 2021, 194, 110656. [Google Scholar] [CrossRef]
  35. Dessalle, A.; Quílez-Bermejo, J.; Fierro, V.; Xu, F.; Celzard, A. Recent Progress in the Development of Efficient Biomass-Based ORR Electrocatalysts. Carbon N. Y. 2023, 203, 237–260. [Google Scholar] [CrossRef]
  36. Quílez-Bermejo, J.; Morallón, E.; Cazorla-Amorós, D. On the Deactivation of N-Doped Carbon Materials Active Sites during Oxygen Reduction Reaction. Carbon N. Y. 2022, 189, 548–560. [Google Scholar] [CrossRef]
  37. Zhang, T.; Wang, H.; Zhang, J.; Ma, J.; Wang, Z.; Liu, J.; Gong, X. Carbon Charge Population and Oxygen Molecular Transport Regulated by Program-Doping for Highly Efficient 4e-ORR. Chem. Eng. J. 2022, 444, 136560. [Google Scholar] [CrossRef]
  38. Abdelwahab, A.; Castelo-Quibén, J.; Vivo-Vilches, J.F.; Pérez-Cadenas, M.; Maldonado-Hódar, F.J.; Carrasco-Marín, F.; Pérez-Cadenas, A.F. Electrodes Based on Carbon Aerogels Partially Graphitized by Doping with Transition Metals for Oxygen Reduction Reaction. Nanomaterials 2018, 8, 266. [Google Scholar] [CrossRef]
  39. Fajardo-Puerto, E.; Elmouwahidi, A.; Bailón-García, E.; Pérez-Cadenas, A.F.; Carrasco-Marín, F. From Fenton and ORR 2e−-Type Catalysts to Bifunctional Electrodes for Environmental Remediation Using the Electro-Fenton Process. Catalysts 2023, 13, 674. [Google Scholar] [CrossRef]
  40. Xu, X.; Pan, Y.; Ge, L.; Chen, Y.; Mao, X.; Guan, D.; Li, M.; Zhong, Y.; Hu, Z.; Peterson, V.K.; et al. High-Performance Perovskite Composite Electrocatalysts Enabled by Controllable Interface Engineering. Small 2021, 17, 2101573. [Google Scholar] [CrossRef]
  41. Li, Y.; Wang, C.; Pan, S.; Zhao, X.; Liu, N. Mn Doping Improves In-Situ H2O2 Generation and Activation in Electro-Fenton Process by Fe/Mn@CC Cathode Using High-Temperature Shock Technique. Chemosphere 2022, 307, 136074. [Google Scholar] [CrossRef] [PubMed]
  42. Abbas, Z.I.; Abbas, A.S. Oxidative Degradation of Phenolic Wastewater by Electro-Fenton Process Using MnO2-Graphite Electrode. J. Environ. Chem. Eng. 2019, 7, 103108. [Google Scholar] [CrossRef]
  43. Li, D.; Sun, T.; Wang, L.; Wang, N. Enhanced Electro-Catalytic Generation of Hydrogen Peroxide and Hydroxyl Radical for Degradation of Phenol Wastewater Using MnO2/Nano-G|Foam-Ni/Pd Composite Cathode. Electrochim. Acta 2018, 282, 416–426. [Google Scholar] [CrossRef]
  44. Rubio, J.A.; Fdez-Güelfo, L.A.; Romero-García, L.I.; Wilkie, A.C.; García-Morales, J.L. Co-Digestion of Two-Phase Olive-Mill Waste and Cattle Manure: Influence of Solids Content on Process Performance. Fuel 2022, 322, 124187. [Google Scholar] [CrossRef]
  45. Elmouwahidi, A.; Bailón-garcía, E.; Pérez-cadenas, A.F.; Maldonado-hódar, F.J.; Carrasco-marín, F. Activated Carbons from KOH and H3PO4 Activation of Olive Residues and Its Application as Supercapacitor Electrodes. Electrochim. Acta 2017, 229, 219–228. [Google Scholar] [CrossRef]
  46. Walton, K.S.; Snurr, R.Q. Applicability of the BET Method for Determining Surface Areas of Microporous Metal-Organic Frameworks. J. Am. Chem. Soc. 2007, 129, 8552–8556. [Google Scholar] [CrossRef]
  47. Raja, P.M.V.; Barron, A.R. 2.3: BET Surface Area Analysis of Nanoparticles; 2019; Vol. i. Available online: https://chem.libretexts.org/Bookshelves/Analytical_Chemistry/Physical_Methods_in_Chemistry_and_Nano_Science_(Barron)/02%3A_Physical_and_Thermal_Analysis/2.03%3A_BET_Surface_Area_Analysis_of_Nanoparticles (accessed on 10 May 2024).
  48. Nguyen, C.; Do, D.D. The Dubinin-Radushkevich Equation and the Underlying Microscopic Adsorption Description. Carbon N. Y. 2001, 39, 1327–1336. [Google Scholar] [CrossRef]
  49. Zhou, R.; Zheng, Y.; Jaroniec, M.; Qiao, S.Z. Determination of the Electron Transfer Number for the Oxygen Reduction Reaction: From Theory to Experiment. ACS Catal. 2016, 6, 4720–4728. [Google Scholar] [CrossRef]
  50. Jiang, K.; Back, S.; Akey, A.J.; Xia, C.; Hu, Y.; Liang, W.; Schaak, D.; Stavitski, E.; Nørskov, J.K.; Siahrostami, S.; et al. Highly Selective Oxygen Reduction to Hydrogen Peroxide on Transition Metal Single Atom Coordination. Nat. Commun. 2019, 10. [Google Scholar] [CrossRef]
  51. Xu, S.; Kim, Y.; Higgins, D.; Yusuf, M.; Jaramillo, T.F.; Prinz, F.B. Building upon the Koutecky-Levich Equation for Evaluation of Next-Generation Oxygen Reduction Reaction Catalysts. Electrochim. Acta 2017, 255, 99–108. [Google Scholar] [CrossRef]
  52. Brocato, S.; Lau, C.; Atanassov, P. Mechanistic Study of Direct Electron Transfer in Bilirubin Oxidase. Electrochim. Acta 2012, 61, 44–49. [Google Scholar] [CrossRef]
  53. Alcañiz-Monge, J.; Pérez-Cadenas, M.; Lozano-Castelló, D. Influence of Pore Size Distribution on Water Adsorption on Silica Gels. J. Porous Mater. 2010, 17, 409–416. [Google Scholar] [CrossRef]
  54. Hong, S.M.; Jang, E.; Dysart, A.D.; Pol, V.G.; Lee, K.B. CO2 Capture in the Sustainable Wheat-Derived Activated Microporous Carbon Compartments. Sci. Rep. 2016, 6, 34590. [Google Scholar] [CrossRef]
  55. Alcañiz-Monge, J.; Linares-Solano, A.; Rand, B. Mechanism of Adsorption of Water in Carbon Micropores as Revealed by a Study of Activated Carbon Fibers. J. Phys. Chem. B 2002, 106, 3209–3216. [Google Scholar] [CrossRef]
  56. Madhuri, S.; Shilpa Chakra, C.; Sadhana, K.; Divya, V. Sketchy Synthesis of Mn3O4, Mn3O4/AC and Mn3O4/CNT Composites for Application of/in Energy Cache. Mater. Today Proc. 2022, 65, 2812–2818. [Google Scholar] [CrossRef]
  57. Wang, B.; Yu, J.; Lu, Q.; Xiao, Z.; Ma, X.; Feng, Y. Preparation of Mn3O4 Microspheres via Glow Discharge Electrolysis Plasma as a High-Capacitance Supercapacitor Electrode Material. J. Alloys Compd. 2022, 926, 166775. [Google Scholar] [CrossRef]
  58. Moses Ezhil Raj, A.; Victoria, S.G.; Jothy, V.B.; Ravidhas, C.; Wollschläger, J.; Suendorf, M.; Neumann, M.; Jayachandran, M.; Sanjeeviraja, C. XRD and XPS Characterization of Mixed Valence Mn3O4 Hausmannite Thin Films Prepared by Chemical Spray Pyrolysis Technique. Appl. Surf. Sci. 2010, 256, 2920–2926. [Google Scholar] [CrossRef]
  59. Feng, T.; Li, H.; Tan, J.; Liang, Y.; Zhu, W.; Zhang, S.; Wu, M. Synthesis, Characterization and Electrochemical Behavior of Zn-Doped MnO/C Submicrospheres for Lithium Ion Batteries. J. Alloys Compd. 2022, 897, 163153. [Google Scholar] [CrossRef]
  60. Li, Z.; Xiao, S.; Wang, Y.; Zhu, Z.; Li, X.; Niu, X.; Kong, Q.; Jiang, J.; Chen, J.S. Fast and Stable Na Insertion/Deinsertion in Double-Shell Hollow MnO Nanospheres. J. Alloys Compd. 2022, 920, 165449. [Google Scholar] [CrossRef]
  61. Ruan, Y.; Lei, H.; Xue, W.; Wang, T.; Song, S.; Xu, H.; Yu, Y.; Zhang, G.R.; Mei, D. Cu-N-C Assisted MnO Nanorods as Bifunctional Electrocatalysts for Superior Long-Cycle Zn-Air Batteries. J. Alloys Compd. 2023, 934, 167781. [Google Scholar] [CrossRef]
  62. Weng, Z.; Li, J.; Weng, Y.; Feng, M.; Zhuang, Z.; Yu, Y. Surfactant-Free Porous Nano-Mn3O4 as a Recyclable Fenton-like Reagent That Can Rapidly Scavenge Phenolics without H2O2. J. Mater. Chem. A Mater. 2017, 5, 15650–15660. [Google Scholar] [CrossRef]
  63. Huang, Z.; Wang, H.; Wang, X.; Ma, X.; Ren, J.; Wang, R. Heterostructured Mn Clusters/N-Doped Carbon Foams Enhancing the Polysulfides Immobilization and Catalytic Conversion. J. Alloys Compd. 2022, 924, 166514. [Google Scholar] [CrossRef]
  64. Zhou, C.; Liang, Y.; Xia, W.; Almatrafi, E.; Song, B.; Wang, Z.; Zeng, Y.; Yang, Y.; Shang, Y.; Wang, C.; et al. Single Atom Mn Anchored on N-Doped Porous Carbon Derived from Spirulina for Catalyzed Peroxymonosulfate to Degradation of Emerging Organic Pollutants. J. Hazard. Mater. 2023, 441, 129871. [Google Scholar] [CrossRef] [PubMed]
  65. Song, S.; Wang, Y.; Tian, P.; Zang, J. One-Step Complexation and Self-Template Strategy to Synthesis Bimetal Fe/Mn–N Doped Interconnected Hierarchical Porous Carbon for Enhancing Catalytic Oxygen Reduction Reaction. Int. J. Hydrogen Energy 2022, 47, 24728–24737. [Google Scholar] [CrossRef]
  66. Gao, H.; Wang, Y.; Li, W.; Zhou, S.; Song, S.; Tian, X.; Yuan, Y.; Zhou, Y.; Zang, J. One-Step Carbonization of ZIF-8 in Mn-Containing Ambience to Prepare Mn, N Co-Doped Porous Carbon as Efficient Oxygen Reduction Reaction Electrocatalyst. Int. J. Hydrogen Energy 2021, 46, 36742–36752. [Google Scholar] [CrossRef]
  67. Moreno-Castilla, C.; López-Ramón, M.V.; Carrasco-Marín, F. Changes in Surface Chemistry of Activated Carbons by Wet Oxidation. Carbon N. Y. 2000, 38, 1995–2001. [Google Scholar] [CrossRef]
  68. Xie, Z.; Lyu, Z.; Wang, J.; Li, A.; François-Xavier Corvini, P. Ultrafine-Mn2O3@N-Doped Porous Carbon Hybrids Derived from Mn-MOFs: Dual-Reaction Centre Catalyst with Singlet Oxygen-Dominant Oxidation Process. Chem. Eng. J. 2022, 429, 132299. [Google Scholar] [CrossRef]
  69. Zhou, H.; Yang, J.; Cao, W.; Chen, C.; Jiang, C.; Wang, Y. Hollow Meso-Crystalline Mn-Doped Fe3O4 Fenton-like Catalysis for Ciprofloxacin Degradation: Applications in Water Purification on Wide PH Range. Appl. Surf. Sci. 2022, 590, 153120. [Google Scholar] [CrossRef]
  70. Wirecka, R.; Lachowicz, D.; Berent, K.; Marzec, M.M.; Bernasik, A. Ion Distribution in Iron Oxide, Zinc and Manganese Ferrite Nanoparticles Studied by XPS Combined with Argon Gas Cluster Ion Beam Sputtering. Surf. Interfaces 2022, 30, 101865. [Google Scholar] [CrossRef]
  71. Chen, J.; Li, J.; Zeng, Q.; Li, H.; Chen, F.; Hou, H.; Lan, J. Efficient Removal of Tetracycline from Aqueous Solution by Mn-N-Doped Carbon Aerogels: Performance and Mechanism. J. Mol. Liq. 2022, 358, 119153. [Google Scholar] [CrossRef]
  72. Gong, Z.; Wang, B.; Chen, W.; Ma, S.; Jiang, W.; Jiang, X. Waste Straw Derived Mn-Doped Carbon/Mesoporous Silica Catalyst for Enhanced Low-Temperature SCR of NO. Waste Manag. 2021, 136, 28–35. [Google Scholar] [CrossRef] [PubMed]
  73. Cen, H.; Wu, C.; Chen, Z. N, S Co-Doped Carbon Coated MnS/MnO/Mn Nanoparticles as a Novel Corrosion Inhibitor for Carbon Steel in CO2-Saturated NaCl Solution. Colloids Surf. A Physicochem. Eng. Asp. 2021, 630, 127528. [Google Scholar] [CrossRef]
  74. Tang, F.; Zhao, Y.W.; Ge, Y.; Sun, Y.G.; Zhang, Y.; Yang, X.L.; Cao, A.M.; Qiu, J.H.; Lin, X.J. Synergistic Effect of Mn Doping and Hollow Structure Boosting Mn-CoP/Co2P Nanotubes as Efficient Bifunctional Electrocatalyst for Overall Water Splitting. J. Colloid. Interface Sci. 2022, 628, 524–533. [Google Scholar] [CrossRef] [PubMed]
  75. Xu, B.; Wu, F.; Chen, R.; Cao, G.; Chen, S.; Zhou, Z.; Yang, Y. Highly Mesoporous and High Surface Area Carbon: A High Capacitance Electrode Material for EDLCs with Various Electrolytes. Electrochem. Commun. 2008, 10, 795–797. [Google Scholar] [CrossRef]
  76. Kuznetsov, D.A.; Han, B.; Yu, Y.; Rao, R.R.; Hwang, J.; Román-Leshkov, Y.; Shao-Horn, Y. Tuning Redox Transitions via Inductive Effect in Metal Oxides and Complexes, and Implications in Oxygen Electrocatalysis. Joule 2018, 2, 225–244. [Google Scholar] [CrossRef]
  77. Elmouwahidi, A.; Bailón-García, E.; Pérez-Cadenas, A.F.; Castelo-Quibén, J.; Carrasco-Marín, F. Carbon-Vanadium Composites as Non-Precious Catalysts for Electro-Reduction of Oxygen. Carbon N. Y. 2019, 144, 289–300. [Google Scholar] [CrossRef]
  78. Xiao, F.; Wang, Z.; Fan, J.; Majima, T.; Zhao, H.; Zhao, G. Selective Electrocatalytic Reduction of Oxygen to Hydroxyl Radicals via 3-Electron Pathway with FeCo Alloy Encapsulated Carbon Aerogel for Fast and Complete Removing Pollutants. Angew. Chem.-Int. Ed. 2021, 60, 10375–10383. [Google Scholar] [CrossRef]
  79. Li, N.; Huang, C.; Wang, X.; Feng, Y.; An, J. Electrosynthesis of Hydrogen Peroxide via Two-Electron Oxygen Reduction Reaction: A Critical Review Focus on Hydrophilicity/Hydrophobicity of Carbonaceous Electrode. Chem. Eng. J. 2022, 450, 138246. [Google Scholar] [CrossRef]
Figure 1. SEM microphotographs of CK2 (A), CK2-Mn-1-10 (B), CK2-Mn-2-10 (C), and CK2-Mn-3-10 (D).
Figure 1. SEM microphotographs of CK2 (A), CK2-Mn-1-10 (B), CK2-Mn-2-10 (C), and CK2-Mn-3-10 (D).
Nanomaterials 14 01112 g001
Figure 2. Isotherms N2 adsorption of samples CK2, Nanomaterials 14 01112 i001; CK2-Mn-1-10, Nanomaterials 14 01112 i002; CK2-Mn-2-10, Nanomaterials 14 01112 i003; CK2-Mn-3-10, Nanomaterials 14 01112 i004; CK2-Mn-2-25, Nanomaterials 14 01112 i005; CK2-Mn-2-60, Nanomaterials 14 01112 i006.
Figure 2. Isotherms N2 adsorption of samples CK2, Nanomaterials 14 01112 i001; CK2-Mn-1-10, Nanomaterials 14 01112 i002; CK2-Mn-2-10, Nanomaterials 14 01112 i003; CK2-Mn-3-10, Nanomaterials 14 01112 i004; CK2-Mn-2-25, Nanomaterials 14 01112 i005; CK2-Mn-2-60, Nanomaterials 14 01112 i006.
Nanomaterials 14 01112 g002
Figure 3. TGA analysis of samples CK2-Mn-2-10, Nanomaterials 14 01112 i002; CK2-Mn-2-25, Nanomaterials 14 01112 i007; CK2-Mn-2-60, Nanomaterials 14 01112 i006.
Figure 3. TGA analysis of samples CK2-Mn-2-10, Nanomaterials 14 01112 i002; CK2-Mn-2-25, Nanomaterials 14 01112 i007; CK2-Mn-2-60, Nanomaterials 14 01112 i006.
Nanomaterials 14 01112 g003
Figure 4. X-ray diffraction of the composites.
Figure 4. X-ray diffraction of the composites.
Nanomaterials 14 01112 g004
Figure 5. XPS spectra of (a) C1s, (b) O1s, and (c) Mn2p for original activated carbon CK2 and samples prepared with different impregnation methods and 10% of manganese loading.
Figure 5. XPS spectra of (a) C1s, (b) O1s, and (c) Mn2p for original activated carbon CK2 and samples prepared with different impregnation methods and 10% of manganese loading.
Nanomaterials 14 01112 g005
Figure 6. XPS spectra of (a) C1s, (b) O1s, and (c) Mn2p for samples prepared with different amounts of manganese.
Figure 6. XPS spectra of (a) C1s, (b) O1s, and (c) Mn2p for samples prepared with different amounts of manganese.
Nanomaterials 14 01112 g006
Figure 7. CV of (A) CK2-Mn-1-10, (B) CK2-Mn-2-10, (C) CK2-Mn-3-10, (D) CK2-Mn-2-25, (E) CK2-Mn-2-60 in N2 (blue) and O2 (red).
Figure 7. CV of (A) CK2-Mn-1-10, (B) CK2-Mn-2-10, (C) CK2-Mn-3-10, (D) CK2-Mn-2-25, (E) CK2-Mn-2-60 in N2 (blue) and O2 (red).
Nanomaterials 14 01112 g007
Figure 8. Correlation between the current density and the surface area B.E.T. of CK2-Mn-1-10, Nanomaterials 14 01112 i002; CK2-Mn-2-10, Nanomaterials 14 01112 i003; CK2-Mn-3-10, Nanomaterials 14 01112 i004; CK2-Mn-2-25, Nanomaterials 14 01112 i005; CK2-Mn-2-60, Nanomaterials 14 01112 i006.
Figure 8. Correlation between the current density and the surface area B.E.T. of CK2-Mn-1-10, Nanomaterials 14 01112 i002; CK2-Mn-2-10, Nanomaterials 14 01112 i003; CK2-Mn-3-10, Nanomaterials 14 01112 i004; CK2-Mn-2-25, Nanomaterials 14 01112 i005; CK2-Mn-2-60, Nanomaterials 14 01112 i006.
Nanomaterials 14 01112 g008
Figure 9. Correlation between the current density and the Mn3+/Mn2+ ratio of CK2-Mn-2-10, Nanomaterials 14 01112 i003; CK2-Mn-2-25, Nanomaterials 14 01112 i005; CK2-Mn-2-60, Nanomaterials 14 01112 i006.
Figure 9. Correlation between the current density and the Mn3+/Mn2+ ratio of CK2-Mn-2-10, Nanomaterials 14 01112 i003; CK2-Mn-2-25, Nanomaterials 14 01112 i005; CK2-Mn-2-60, Nanomaterials 14 01112 i006.
Nanomaterials 14 01112 g009
Figure 10. (A) Number of electrons transferred, (B) Selectivity to H2O2, (C) Current Ring and Disc of CK2, Nanomaterials 14 01112 i001; CK2-Mn-1-10, Nanomaterials 14 01112 i002; CK2-Mn-2-10, Nanomaterials 14 01112 i003; CK2-Mn-3-10, Nanomaterials 14 01112 i004; CK2-Mn-2-25, Nanomaterials 14 01112 i005; CK2-Mn-2-60, Nanomaterials 14 01112 i006.
Figure 10. (A) Number of electrons transferred, (B) Selectivity to H2O2, (C) Current Ring and Disc of CK2, Nanomaterials 14 01112 i001; CK2-Mn-1-10, Nanomaterials 14 01112 i002; CK2-Mn-2-10, Nanomaterials 14 01112 i003; CK2-Mn-3-10, Nanomaterials 14 01112 i004; CK2-Mn-2-25, Nanomaterials 14 01112 i005; CK2-Mn-2-60, Nanomaterials 14 01112 i006.
Nanomaterials 14 01112 g010
Figure 11. TC degradation by electro-Fenton with Graphite, Nanomaterials 14 01112 i008; CK2, Nanomaterials 14 01112 i001; CK2-Mn-1-10, Nanomaterials 14 01112 i002; CK2-Mn-2-10, Nanomaterials 14 01112 i003; CK2-Mn-3-10, Nanomaterials 14 01112 i004.
Figure 11. TC degradation by electro-Fenton with Graphite, Nanomaterials 14 01112 i008; CK2, Nanomaterials 14 01112 i001; CK2-Mn-1-10, Nanomaterials 14 01112 i002; CK2-Mn-2-10, Nanomaterials 14 01112 i003; CK2-Mn-3-10, Nanomaterials 14 01112 i004.
Nanomaterials 14 01112 g011
Figure 12. TC degradation by electro Fenton with CK2-Mn-1-10 in O2, Nanomaterials 14 01112 i002; CK2-Mn-1-10 in Fenton traditional, Nanomaterials 14 01112 i009; CK2-Mn-1-10 in N2 and H2O2, Nanomaterials 14 01112 i010; H2O2 without catalyst, Nanomaterials 14 01112 i011.
Figure 12. TC degradation by electro Fenton with CK2-Mn-1-10 in O2, Nanomaterials 14 01112 i002; CK2-Mn-1-10 in Fenton traditional, Nanomaterials 14 01112 i009; CK2-Mn-1-10 in N2 and H2O2, Nanomaterials 14 01112 i010; H2O2 without catalyst, Nanomaterials 14 01112 i011.
Nanomaterials 14 01112 g012
Figure 13. Correlation between the TTC degradation and the JK value (a) and % of Mn2+ (b), of CK2-Mn-1-10, Nanomaterials 14 01112 i002; CK2-Mn-2-10, Nanomaterials 14 01112 i003; CK2-Mn-3-10, Nanomaterials 14 01112 i004.
Figure 13. Correlation between the TTC degradation and the JK value (a) and % of Mn2+ (b), of CK2-Mn-1-10, Nanomaterials 14 01112 i002; CK2-Mn-2-10, Nanomaterials 14 01112 i003; CK2-Mn-3-10, Nanomaterials 14 01112 i004.
Nanomaterials 14 01112 g013
Table 1. Textural characteristics of all the samples obtained by N2 adsorption isotherms at −196 °C and CO2 at 0 °C.
Table 1. Textural characteristics of all the samples obtained by N2 adsorption isotherms at −196 °C and CO2 at 0 °C.
Samples N2 CO2
SBETW0L0W0L0
m2 g−1cm3 g−1nmcm3 g−1nm
CK216720.661.330.380.68
CK2-Mn-1-1012320.501.320.290.67
CK2-Mn-2-1013800.561.330.310.67
CK2-Mn-3-1012570.501.310.871.28
CK2-Mn-2-259450.371.330.220.82
CK2-Mn-2-604120.161.400.100.73
Table 2. Current density Jk (mA cm−2).
Table 2. Current density Jk (mA cm−2).
SampleJK
(mA cm−2)
CK27.39
CK2-Mn-1-108.99
CK2-Mn-2-1011.37
CK2-Mn-3-1010.68
CK2-Mn-2-257.44
CK2-Mn-2-602.68
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

Fajardo-Puerto, E.; Elmouwahidi, A.; Bailón-García, E.; Pérez-Cadenas, M.; Pérez-Cadenas, A.F.; Carrasco-Marín, F. Antibiotic Degradation via Fenton Process Assisted by a 3-Electron Oxygen Reduction Reaction Pathway Catalyzed by Bio-Carbon–Manganese Composites. Nanomaterials 2024, 14, 1112. https://doi.org/10.3390/nano14131112

AMA Style

Fajardo-Puerto E, Elmouwahidi A, Bailón-García E, Pérez-Cadenas M, Pérez-Cadenas AF, Carrasco-Marín F. Antibiotic Degradation via Fenton Process Assisted by a 3-Electron Oxygen Reduction Reaction Pathway Catalyzed by Bio-Carbon–Manganese Composites. Nanomaterials. 2024; 14(13):1112. https://doi.org/10.3390/nano14131112

Chicago/Turabian Style

Fajardo-Puerto, Edgar, Abdelhakim Elmouwahidi, Esther Bailón-García, María Pérez-Cadenas, Agustín F. Pérez-Cadenas, and Francisco Carrasco-Marín. 2024. "Antibiotic Degradation via Fenton Process Assisted by a 3-Electron Oxygen Reduction Reaction Pathway Catalyzed by Bio-Carbon–Manganese Composites" Nanomaterials 14, no. 13: 1112. https://doi.org/10.3390/nano14131112

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

Article Metrics

Back to TopTop