Next Article in Journal
DSS-OSM: An Integrated Decision Support System for Offshore Oil Spill Management
Next Article in Special Issue
Comparison of Adsorptive Removal of Fluoride from Water by Different Adsorbents under Laboratory and Real Conditions
Previous Article in Journal
Using the Change Point Model (CPM) Framework to Identify Windows for Water Resource Management Action in the Lower Colorado River Basin of Texas, USA
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhanced Ozone Oxidation by a Novel Fe/Mn@γ−Al2O3 Nanocatalyst: The Role of Hydroxyl Radical and Singlet Oxygen

1
Guangzhou Higher Education Mega Centre, School of Environment and Energy, South China University of Technology, Guangzhou 510006, China
2
Guangzhou Higher Education Mega Centre, The Key Lab of Pollution Control and Ecosystem Restoration in Industry Clusters, Ministry of Education, Guangzhou 510006, China
*
Author to whom correspondence should be addressed.
Water 2022, 14(1), 19; https://doi.org/10.3390/w14010019
Submission received: 12 November 2021 / Revised: 15 December 2021 / Accepted: 20 December 2021 / Published: 22 December 2021
(This article belongs to the Special Issue Advanced Technologies and Materials for Polluted Water Remediation)

Abstract

:
Catalytic ozonation is a potential alternative to address the dye wastewater effluent, and developing an effective catalyst for catalyzing ozone is desired. In this study, a novel Fe/Mn@γ−Al2O3 nanomaterial was prepared and successfully utilized for catalytic ozonation toward dye wastewater effluent components (dimethyl phthalate and 1−naphthol). The synthesized Fe/Mn@γ−Al2O3 exhibited superior activity in catalytic ozonation of dimethyl phthalate and 1−naphthol in contrast to Fe@γ−Al2O3 and Mn@γ−Al2O3. Quench and probe tests indicated that HO° contributed to almost all removal of dimethyl phthalate, whereas O3, HO°, and singlet oxygen participated in the degradation of 1−naphthol in the Fe/Mn@γ−Al2O3/O3 system. The results of XPS, FT−IR, and EPR suggested that HO° and singlet oxygen were generated from the valence variations of Fe(II/III)and Mn(III/IV). Moreover, the Fe/Mn@γ−Al2O3/O3 system could also have excellent efficacy in actual water samples, including dye wastewater effluent. This study presents an efficient ozone catalyst to purify dye wastewater effluent and deepens the comprehension of the role and formation of reactive species involved in the catalytic ozonation system.

Graphical Abstract

1. Introduction

Advanced oxidation processes (AOPs) gained much attention as powerful techniques to mineralize the pollutants in water treatments [1]. Catalytic ozonation is a promising technology for efficiently removing refractory pollutants, especially in the advanced treatment of wastewater. Heterogeneous catalytic ozonation can minimize the dissolving of toxic metal cations in contrast to homogeneous catalyzing.
Many metal oxides have been utilized in the heterogeneous catalytic ozonation process, including manganese oxides, iron oxides/oxyhydroxide, aluminum oxides, and bimetallic/polymetallic oxides. MnOX supported by granular activated carbon or alumina catalyzed ozone to generate hydroxyl radical (HO°) and efficiently degrade benzenes [2,3]. The surface hydroxyl groups of hydroxylated synthetic ɑ−FeOOH can promote catalyzing ozone to yield HO° [4], and surface MeO−H weak bonds were the favorable sites for accelerating HO° generation. The complex of oxalic acid and iron (Fe2O3/Al2O3) reacted with ozone and accelerated the degradation of oxalic acid [5]. In addition to Mn and Fe, TiO2 and MgO nanoparticles were also investigated and used to catalyze ozone [6,7].
Bimetallic nanostructured materials hold promise for improving catalyst activity and selectivity [8]. Mesoporous bimetallic catalysts, such as Ru−Cu/SBA−15, Co−Mn−MCM−41 catalysts, accelerated catalytic ozonation to generate more HO° to degrade dye industry effluent or dimethyl phthalate (DMP) in contrast to single metal loading [9,10]. Fe−Ni/activated carbon also exhibited high performance for degrading 2,4−dichlorophenoxyacetic acid compared to single metal or activated carbon [11]. The micron−sized Fe0/Cu/O3 process enhanced p−nitrophenol mineralization more than twice the sum of ozone and Fe0/Cu alone [12]. The ratios of surface defective oxygen/lattice oxygen and Mn(III)/Mn(IV) of Mn−M bimetallic HZSM−5 (M: Fe, Cu, Ru, Ag) catalysts were higher than that of Mn−only catalyst, of which Ru−Mn/HZSM−5 showed the highest efficiency in degrading toluene [13,14].
Ozonation, radical oxidation, and nonradical reaction in the catalytic ozonation system mainly contributed to contaminant degradation [15,16,17]. In addition to HO°, reactive oxygen species (ROS), including singlet oxygen (1O2) and superoxide radical (O2°), were also involved in contaminant degradation during catalytic ozonation [18]. Oxygen vacancies and multi−valence Mn were the main Lewis acidic sites for Mn2O3/LaMnO3−δ perovskite composites during catalytic ozonation, and the generated HO°, 1O2, and O2° were possibly responsible for the degradation of 1H−benzotriazole [19]. As bimetallic nanoparticles and Fe− and Mn−based materials exhibited excellent performance on catalytic ozonation, however, Fe/Mn bimetallic material in catalytic ozonation has not been reported so far. Consequently, the performance of Fe/Mn bimetallic nanoparticles on catalyzing ozone and the corresponding reactive species in such systems needs further evaluation.
This study aims to investigate the performance of catalytic ozonation by synthesized Fe/Mn@γ−Al2O3 nanoparticles. A modified evaporation−induced self−assembly method was employed to fabricate Fe/Mn@γ−Al2O3 nanoparticles, and SEM, TEM, XRD, XPS, FTIR, and BET were utilized to characterize the morphology, crystal form, element distribution, and composition. DMP and 1−naphthol (1−NP), dye precursors and intermediates, were selected as model compounds. Their removal performance, reactive species, reaction mechanism, degradation pathway, and matrix factors were investigated in catalytic ozonation system in the presence of Fe/Mn@γ−Al2O3 nanoparticles.

2. Materials and Methods

2.1. Catalyst Preparation

Al2O3−based catalysts were prepared by a modified evaporation−induced self−assembly process as described previously for pure γ−Al2O3 [20]. Aluminum isopropoxide [Al(OiPr)3], Fe(NO3)3·9H2O, and Mn(NO3)2 were selected as the sources of Al, Fe, and Mn. [Al(OiPr)3] (8.4 g), glucose (7.2 g), and a required dosage of Fe(NO3)3·9H2O and Mn(NO3)2 were dissolved thoroughly in ultrapure water at 35 °C. The pH value of the mixture was adjusted to 5.5 using formic acid (10 wt.%), and the mixed solution was ultrasonicated for 4 h and heated at 105 °C. The final solid after grinding was calcined at 600 °C for 6 h, and a series of γ−Al2O3−based catalysts were obtained and termed as Fe@γ−Al2O3 Mn@γ−Al2O3 and Fe/Mn@γ−Al2O3.

2.2. Catalyst Characterization

The morphology and chemical composition of catalysts were characterized by transmission electron microscope (TEM, Talos L120C) and scanning electronic microscopy−energy dispersion spectroscopy (SEM−EDS, Merlin, Carl Zeiss AG, GER). The powder X−ray diffraction (XRD, Empyrean, NL) spectra of the obtained catalysts were collected with Cu−Kα radiation (λ = 1.5418 Å) at 40 kV and 40 mA operating from the 2θ angle of 10−90. The specific surface area and pore size distribution of the catalysts were calculated from the N2 adsorption/desorption isotherms at 77 K (ASAP2020, Micromeritics, Norcross, GA, USA). X−ray photoelectron spectroscopy (XPS, EscaLab Xi+, ThemoFisher Scientific, Leicestershire, UK) spectra were recorded by using Al−Kα radiation at 15 kV and 51 W, and the binding energies were calibrated with C1s at 284.8 eV. Fourier transform infrared spectra (FT−IR) were recorded using a CCR−1 (Thermo−Nicolet, Waltham, MA, USA).

2.3. Catalytic Ozonation Procedure

The catalytic ozonation experiments of target contaminants were performed in a 1.2 L column reactor at a temperature of 25 °C. Ozone was produced in situ from pure oxygen by a ZA−1G laboratory ozone generator (Guangzhou Zeao Ozone Equipment Co., Ltd., Guangzhou, China) and continuously bubbled into the reactor through the reactor’s porous plate at a bottom 200 mL min−1 flow rate (6 mg L−1). Catalyst powder (200 mg L−1) was mixed in the reactor under continuously magnetically stirring. Target compounds (DMP/1−NP, 50 mM) with/without quenchers (p−benzoquinone (p−BQ), nitrobenzene (NB), furfuryl alcohol (FFA)) or probes (NB, FFA) were added in the catalytic ozonation system. The same procedures were carried out for the control experiments, including ozone alone, catalyst without ozone, and catalyst with oxygen. All solutions were prepared with ultrapure water (≥18.2 MΩ×cm). Samples were taken at regular intervals, filtered by a Millipore filter (0.45 μm), and quenched by Na2S2O3, then DMP, 1−NP, NB, and FFA were analyzed. All experiments were conducted in duplicate.

2.4. Analytic Methods

DMP, 1−NP, NB, and FFA concentrations were quantified using high−performance liquid chromatography (UPLC, Agilent 1200) equipped with a PDA 2998 detector and a C−18 column (Agilent Eclipse, 5 μm, 4.6 mm × 250 mm). The mobile phase was acetonitrile/0.02% phosphoric acid water (60:40, v/v) with a flow rate of 0.1 mL min−1. The analytical wavelengths of DMP, 1−NP, NB, and FFA were set at 228 nm, 210 nm, 265 nm, and 204 nm at 30°C. The electron paramagnetic resonance (EPR) spectra were measured by Bruker EMX plus instrument (Karlsruhe, Germany). The compound 2,2,6,6−Tetramethylpiperidine (TEMP) was used to capture 1O2, and 5,5−dimethyl−1−pyrroline−N−oxide (DMPO) was applied to capture HO° and O2°. In detail, the samples for EPR tests were dispersed in a 100 mM DMPO/TEMP solution (2 mL) with O3 purging, and the solvent was H2O for trapping 1O2 and HO° and methanol for trapping O2°. The concentration of leached metals was determined by graphite furnace atomic absorption spectrometry (AAS, Pin AAcle 900T, Perkin Elmer, Waltham, MA, USA). The intermediates of DMP and 1−NP during catalytic ozonation were identified by using LC−MS (Thermo Fisher Scientific, Q Exactive Plus, USA).

3. Results

3.1. Characterization

A series of γ−Al2O3−based nanocatalysts were synthesized using the glucose template method [20]. SEM images (Figure 1A−D) show amorphous particle structure for all materials, and Fe/Mn@γ−Al2O3 was more compact compared to others. TEM images (Figure S1) displayed the nanoscale structure of Fe/Mn@γ−Al2O3 with a thickness of 5−10 nm. The element mapping of Fe/Mn@γ−Al2O3 indicated that four elements were distributed uniformly (Figure 1E,F), and the element composition of Fe/Mn was 1.3, which corresponded to the nominal loading molar ratio (1.5). Figure 1G shows the XRD patterns of γ−Al2O3−based nanomaterials, and five broad peaks (39.4°, 45.9°, 66.9°, 80.5°, 84.9° of 2θ) were observed and related to γ−Al2O3 phase [21]. The peaks of Fe/Mn oxide alloy (35° of 2θ), Fe2O3 (33° of 2θ), and MnO2 (21° of 2θ) were not observed, which might be affected by γ−Al2O3 peaks [22]. FT−IR spectroscopy at approximately 575, 748, and 819 cm−1 derived from the Al−O vibrations confirmed the formation of γ−Al2O3 (Figure S2) [23]. Additionally, XPS full spectra of γ−Al2O3−based nanomaterials identified O, Al, Fe, and Mn, indicating Fe and Mn oxides grew on the γ−Al2O3 (Figure S3).
The nitrogen sorption isotherm of Fe/Mn@γ−Al2O3 is shown in Figure 1H, exhibiting a type IV N2 sorption isotherm with an H3 hysteresis loop, which is a typical characteristic of a mesoporous material [24]. The BET surface area, total pore volume, and median mesopore diameter of Fe/Mn@γ−Al2O3 were 304.2 m2 g−1, 0.416 cm3 g−1, and 3.830 nm (Figure S4), which were similar to γ−Al2O3−based materials (Sasol product) [21]. This result indicates that the Fe−Mn catalyst was impregnated well into the γ−Al2O3 without pore blocking. Overall, the Fe/Mn@γ−Al2O3 catalyst was a mesoporous nanomaterial with uniform iron−manganese oxide growing on the surface.

3.2. Degradation of DMP and 1−NP

DMP and 1−NP, the precursor or intermediate of dyes, were selected as model compounds (Figure S5), and Figure 2 illustrates their degradation in the catalytic ozonation system. The observed first−order rate of 1−NP (4.13 × 10−3 s−1) under ozonation alone was higher in contrast to DMP (6.64 × 10−4 s−1) (Figure 2A,B). This result corresponds with the fact that the reactivity of DMP with O3 alone was relatively low (0.2 M−1 s−1) compared with 1−NP (~102 M−1 s−1) (Table 1). Note that the second−order rate of 1−NP with O3 was estimated roughly using the quench test (Figure S6), and the value was in the range of ~102–3 × 103 M−1 s−1 for naphthalene, 2−methylnaphthalene, and 1−chloronaphthalene [25,26,27]. Similarly, the observed first−order rate constants of 1−NP under each catalytic ozonation treatment (i.e., Fe@γ−Al2O3/O3, Mn@γ−Al2O3/O3, Fe/Mn@γ−Al2O3/O3) were higher than those of DMP. Specifically, the Fe/Mn@γ−Al2O3 catalytic was the most effective to catalyze ozone for removing DMP or 1−NP (2.04 × 10−3 s−1 for DMP, 2.25 × 10−2 s−1 for 1−NP), followed by Fe@γ−Al2O3, Mn@γ−Al2O3. Figure S7 shows that the catalytic activity decreased less than 10% after five cycles of use, and the concentration of leached−out Fe and Mn increased slightly during the cycling use (Table S1), implying that Fe/Mn@γ−Al2O3 was relatively stable during the ozonation reactions. Overall, the catalytic ozonation by Fe/Mn@γ−Al2O3 catalytic was feasible and efficient in removing O3−resistant compounds.
The quench and probe tests were conducted to investigate the reactive species involved in catalytic ozonation. As noted in previous studies, HO° played the dominant role in removing organic compounds [28,29,30,31]. Moreover, other ROS (e.g., 1O2, O2°) also participated in the chemical oxidation during catalytic ozonation [32]. Thus, two model compounds with different reactivity toward HO° and ROS (Table 1) were selected to distinguish their roles in the contaminant removal during catalytic ozonation.
NB, FFA, and p−BQ at a high level (100 mM) were chosen to selectively quench HO°, 1O2, and O2°, and their second−order rate constants are listed in Table S2. Figure 2C shows that NB can inhibit most DMP removal (> 99%), and the removal efficiency of DMP with the presence of FFA or p−BQ was almost the same as that with the presence of NB. This result indicated that HO° was the main contributor to DMP degradation under the Fe/Mn@γ−Al2O3/O3 system, whereas 1O2 or O2° was not involved in DMP degradation. Being different from DMP, the removal of 1−NP was not inhibited by NB completely (Figure 2C), and the removal rate constant decreased by 50% in contrast to that without NB. Both FFA and p−BQ almost inhibited the degradation of 1−NP completely. This result indicated that in addition to HO°, 1O2, or O2° also played a significant role in the degradation of 1−NP. EPR spectroscopy was employed to characterize ROS and demonstrated that HO°,1O2, and O2° were yielded under the Fe/Mn@γ−Al2O3/O3 system, but the signal of O2° was relatively low (detailed in Section 3.3).
The probe test was conducted to specify the contributions of diverse reactive species to DMP and 1−NP degradation under the Fe/Mn@γ−Al2O3/O3 system. NB and FFA were selected as the probes of HO° and 1O2, and the modeled steady−state concentrations of HO° and 1O2 ([HO°]SS, [1O2]SS) are listed in Table S3. The estimated observed first−order rate constants of DMP (3.0 × 10−3 s−1) or 1−NP (3.3 × 10−2 s−1) were calculated based on Equations (1)–(3), which were nearly the same as the experimental ones (2.1 × 10−3 s−1 or 2.2 × 10−2 s−1) (Figure 2D).
k = k O 3   + k HO °   + k O 1 2
k = k O 3 , DMP [ O 3 ] + k HO ° , DMP [ HO ° ] + k O 1 2 , DMP [ O 1 2 ]
k = k O 3 , 1 - NP [ O 3 ] + k HO ° , 1 - NP [ HO ° ] + k O 1 2 , 1 - NP [ O 1 2 ]
Specifically, [HO°]SS and [1O2]SS were 8.1 × 10−13 M and 1.8 × 10−9 M during the degradation of DMP in catalytic ozonation system. The estimated first−order rate constant of DMP with HO° was 3.0 × 10−3 s−1. Note that [1O2]SS was as high as [HO°]SS but can rarely contribute to the degradation of DMP (1.8 × 10−6 s−1) due to the relatively low reactivity ( k O 1 2 , DMP < 103 M−1 s−1). The DMP degradation during catalytic ozonation was consequent on the HO°−mediated oxidation. As for the 1−NP−involved system, [HO°]SS and [1O2]SS were 1.2 × 10−12 M and 2.0 × 10−9 M (Table S3), which were comparable with the DMP involved system. However, the degradation of 1−NP was not only contributed to by HO°, but 1O2 and O3 also participated in its removal. The estimated rate constant of each reactive species ( k O 3 , 1 - NP , k HO ° , 1 - NP , k O 1 2 , 1 - NP ) accounted for 7.51%, 46.85%, and 45.65% of the total 1−NP removal rate constant, respectively.
The above results suggested that the Fe/Mn@γ−Al2O3/O3 system can provide comparable levels of HO° (10−12 − 10−13 M) and 1O2 (10−9 M) that were involved in removing different pollutants. HO° (E0 = 2.7 V) is non−selective thus reacts with both DMP and 1−NP at considerably high second−order rate constants (109 − 1010 M−1 s−1) [37,38,39,40]. O3 and 1O2 are more selective to naphthalenes (e.g., naproxen and propranolol) than DMP. Therefore, the inherent O3 and the yielded 1O2 in the Fe/Mn@γ−Al2O3/O3 system made contributions to the degradation of 1−NP (> 40%).

3.3. Catalytic Ozonation Mechanism

HO° and 1O2 were identified and involved in degrading pollutants in the Fe/Mn@γ−Al2O3/O3 system. XPS and EPR spectra were utilized to explore their formation and catalytic mechanisms in such a system.
The XPS spectra of Fe 2p, Mn 2p, and O 1s for Fe/Mn@γ−Al2O3 before/after catalytic ozonation are shown in Figure 3A−C. As described in previous studies, the Fe 2p3/2 and Fe 2p1/2 peaks could be deconvoluted into three peaks, including Fe(II) (711.7 eV and 725.3 eV), Fe(III) (712.9 and 726. eV), and satellite (719.6 and 732.8 eV), respectively (Figure 3A) [41,42]. The Mn 2p3/2 and Mn 2p1/2 peaks could be deconvoluted into two peaks, including Mn(III) (642.2 eV and 653.9 eV) and Mn(IV) (643.8 eV and 654.8 eV) (Figure 3B) [22,43]. The contents of the high valence of both Fe and Mn (i.e., Fe(III) and Mn(IV)) increased slightly after ozonation (Table S4), indicating the conversion between Fe(II))/Mn(III) and Fe(III)/Mn(IV) in the Fe/Mn@γ−Al2O3/O3 system. The previous studies also found valence variation of doping metals during ozonation [44,45]. The formation of high−valence metals resulted in the transformation of oxygen vacancy (VO) to lattice oxygen (Olat) due to the strong storing/releasing oxygen nature of the metal redox couple [46]. Moreover, oxygen mobility by electron transfer caused the valence variation of doping metals [47,48,49].
Three characteristic peaks were identified in the O 1s spectrum of Fe/Mn@γ−Al2O3 and regarded as adsorbed oxygen (Oads, 530.3 eV), Olat, (532.4 eV), and oxygen of metal oxide (OMO, 531.2 eV), respectively (Figure 3C) [42,50]. After ozonation, the relative contents of Olat/Oads significantly increased from 1.32 to 1.64 (Table S4), whereas a subtle increase in that of Mn(IV) or Fe(II) was observed, indicating the prominent role of Oads in reacting with pollutants rather than high valence metal. Specifically, the increase in Olat contributed to the metal valence change and the formation of Mn(IV) or Fe(III). However, Oads declined slightly after ozonation, which might have resulted from the replenishment of VO [51,52]. Overall, the continuous catalytic activity of Fe/Mn@γ−Al2O3 originated from the interaction among VO, Olat, and the transformation of Fe(II)/(III)and Mn(III)/(IV) redox couple.
Solid EPR was applied to directly and accurately detect the unpaired electron information of Fe/Mn@γ−Al2O3 for identifying VO (Figure 3D). Note that the EPR signal of Fe/Mn@γ−Al2O3 at g = 2.095 was more intensive than others, demonstrating that the introduction of Fe and Mn led to more formation of VO on the Fe/Mn@γ−Al2O3 surface [45]. VO contained many unpaired electrons and preferred to serve as active sites for the adsorption of oxygen (e.g., molecule O2, molecule O3), thereby facilitating the formation of ROS (HO°, 1O2, O2°, etc.) [53].
The reactive species in the Fe/Mn@γ−Al2O3/O3 system was identified by EPR tests, as illustrated in Figure 3D. Generally, DMPO and TEMP are typical trapping chemicals for HO°, O2°, and 1O2. The characteristic signal of DMPO−HO° (AH = AN = 14.9 G), with a peak intensity ratio of 1:2:2:1, was obtained in both ozonation alone and Fe/Mn@γ−Al2O3/O3 systems, but no signal was observed in pure water. Moreover, the intensity of DMPO−HO° adduct elevated significantly under the Fe/Mn@γ−Al2O3/O3 system, implying Fe/Mn@γ−Al2O3 dramatically enhanced ozone decomposition to generate more HO°. TEMP was applied as the capture agent for 1O2. The typical signal of TEMP−1O2 (AN = 16.9 G) in the Fe/Mn@γ−Al2O3/O3 catalytic ozonation system was more intensive than that in the ozonation alone system. The results implied that 1O2 was generated in the ozone−based systems and enhanced in the Fe/Mn@γ−Al2O3 catalytic ozonation process. O2° in the ozonation alone and Fe/Mn@γ−Al2O3/O3 systems were also captured using DMPO to form DMPO−O2° (AH = 10.2 G,  AN = 12.9 G), but the intensity was very weak and not promoted in the catalytic system, indicating the neglectable role of 1O2 in the catalytic ozonation system.
As illustrated in Figure S2, FT−IR spectroscopy of Fe/Mn@γ−Al2O3 at ~3460 cm−1 and ~1634 cm−1 corresponded to the surface hydroxyl group and the chemisorbed water, respectively [24,54]. Metal oxides adsorbed water molecules further dissociated to hydroxyls, generating the surface hydroxyl groups at Lewis acid sites of metal oxide surfaces [55,56,57,58,59]. The surface hydroxyl group functioned as Brønsted acid for catalytic ozone decomposition and promoted HO° generation [55,56,57,58,59].
The catalytic ozonation mechanism of Fe/Mn@γ−Al2O3 was proposed and illustrated in Figure 3F. First, O3 was adsorbed at the surface hydroxyl group of metal oxide and interacted with Vo. The VO was replenished with O donated by O3 and produced an O2 molecule releasing into the environment and active oxygen (O2−) to form adsorbed oxygen (Oads) [60,61]. The O2− further reacted with O3 to generate O22− (peroxide) [50]. The yielded O2−/O22− (i.e., Oads) then reacted with water molecules to generate ROS, including HO°, 1O2, and O2° via charge−transfer interactions. Thereinto, O2° underwent self−quenching to form 1O2 or react with O3 to yield HO° (k = 1.6 × 109 M−1 s−1) [62]. VO also converted to Olat, leading to the metal valence increase. The interaction among VO, Fe(II/III)/Mn(III/IV), and Olat of Fe/Mn@γ−Al2O3 resulted in the circulation of metal valence, thus constantly catalyzing O3 [63].

3.4. Proposed Degradation Pathways of DMP and 1−NP

The intermediates of DMP and 1−NP during Fe/Mn@γ−Al2O3 catalytic ozonation were analyzed using LC−MS, and their identified intermediates are detailed in Figures S8 and S9. As previously noted, the degradation of DMP was dominated by HO°−induced reaction, and that of 1−NP resulted from O3, HO°, and 1O2 oxidation. Their degradation pathways are proposed and illustrated in Figure 4.
The degradation of DMP during Fe/Mn@γ−Al2O3 catalytic ozonation (Figure 4A) was initiated by HO° to attack the ester group to form P180 and further oxidized to generate 1,2−benzenedicarboxylicacid [64], or to undergo addition on the benzene ring to yield HO−adducts (P210, P226, P242) [65]. The formed 1,2−benzenedicarboxylicacid reacted with HO° to yield HO−adducts (P182) or transform to phthalic anhydride via dehydration. The adducts P210, P226, and P242 were oxidized by HO° to form P196, P182, and P198. 1,2−benzenedicarboxylicacid and its HO−adducts (P182 and P198) underwent a ring−opening reaction to generate maleic acid and phenols. Phthalic anhydride and maleic acid were decomposed to yield low molecular weight organic acids (e.g., acetic acid, oxalic acid) via ring−opening reactions, which were finally mineralized to CO2 and H2O [66].
Figure 4B shows the degradation pathway of 1−NP under the Fe/Mn@γ−Al2O3/O3 system. Ozone quickly reacted with the naphthalene group as electron−rich moieties to produce hydroxy−1,4−naphthoquinone (P174) that was oxidized to yield 1,4−naphthoquinone (P158) [67]. The initial attack of HO° on 1−naphthol generated transient naphthyloxy radical (1−NP°(−H)) via H−abstraction and transformed to 1,4−naphthoquinone (P158) [68]. 1O2 readily added to 1−NP to form P176, and P176 rearranged and dehydrated to yield 1,4−naphthoquinone (P158) [69]. The generated 1,4−naphthoquinone was further oxidized to 1,2−benzenedicarboxylicacid and phthalic anhydride by ozone and HO°, which were finally mineralized to CO2 and H2O [67].

3.5. Effects of Water Matrix

The performance of the synthesized catalyst in the complex water matrix was evaluated to investigate its potential application, especially for dyeing wastewater effluent. Thus, the effects of ions, pH, and actual water samples on the degradation efficacy of DMP and 1−NP were investigated (Figure 5).
Carbonate, sulfate, chloride, and nitrate were widely detected in dyeing wastewater. Figure 5A,D show their effects on the performance of catalytic ozonation in the ultrapure water system. As seen, sulfate, chloride, and nitrate cannot distinctly inhibit the degradation of DMP and 1−NP (< 5%) because sulfate and nitrate barely consumed the reactive species (<10 M−1 s−1) involved in the Fe/Mn@γ−Al2O3/O3 system. Chloride reacts with HO° but cannot accelerate forward reaction at a low concentration (2 mM) (k+/− = 3.0 × 109 M−1 s−1/6.10 × 109 M−1 s−1) [70,71]. Being different from chloride, carbonate as HO° scavenger (k = 1.0 × 107 M−1 s−1) intensively scavenged [HO°], thus impeding their removal (> 50%). The formed CO3° (E0 = 1.63 V) was much less reactive than HO° (E0 = 2.7 V) [37,72].
The pH value affects the concentration and speciation of reactive species during catalytic ozonation and thus leads to the variation in pollutant removal efficiency. The removal rate increased from 1.83 × 10−3 s−1 to 2.46 × 10−3 s−1 for DMP and from 1.90 × 10−2 to 2.72 × 10−2 s−1 for 1−NP within pH 5−8 (Figure 5B,E). Their removal efficiencies exceeded 90% within 30 min, indicating that the synthesized Fe/Mn@γ−Al2O3 catalyst had a good ozone−catalyzing performance at a wide pH variation. Specifically, the alkali condition was more suitable for the contaminant degradation in the Fe/Mn@γ−Al2O3/O3 system as OH is mainly hydroxyl radical initiator in ozonation [73], corresponding with previous studies [74,75].
To further investigate the effect of the water matrix on the removal of 1−NP and DMP with the Fe/Mn@γ−Al2O3/O3 system, DMP/1−NP (50 μM) was spiked in printing and dyeing industrial park (PDIP) effluent, wastewater treatment plant (WWTP) effluent, and river water (RW). The water quality parameters of three water samples are listed in Table S5. The removal rate constants of 1−NP and DMP in all actual water samples decreased by 50–60% in contrast to ultrapure water (Figure 5C,F). This decrease resulted from the competition of reactive species by dissolved organic matter (DOM) in the actual water samples. In contrast, the Fe/Mn@γ−Al2O3/O3 system accelerated the DOC and UV254 decrease in the actual water samples compared with ozonation alone (Figure S11). Overall, the synthesized Fe/Mn@γ−Al2O3 can be potentially applied in the catalytic ozonation of actual water.

4. Conclusions

A novel Fe/Mn@γ−Al2O3 nanocatalyst was successfully synthesized and utilized for catalytic ozonation of DMP and 1−NP. Their removal rate constants could elevate 3−5 times with Fe/Mn@γ−Al2O3/O3 system compared to ozonation alone due to more yield of HO° and 1O2 involved in their degradation. The characterization of Fe/Mn@γ−Al2O3 suggested that the formation of HO° and 1O2 possibly resulted from the transformation of O3 from VO to Oads with the circulation of Mn(III)/(IV) and Fe(II)/(III). The degradation pathways of DMP and 1−NP were proposed based on the mechanisms and the identified intermediates. Additionally, the Fe/Mn@γ−Al2O3/O3 system could be applied in removing pollutants at pH 6−9, in the presence of different ions, and dealing with actual dye wastewater effluent samples. The performance and practicability of the Fe/Mn@γ−Al2O3/O3 system on the treatment of dye wastewater effluent should be further evaluated in a pilot−/full−scale experiment.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/w14010019/s1, Text S1: Steady−state concentrations of different reactive species, Text S2: Determination of the rate constant of DMP with HO°, Text S3: Factors influencing catalytic performance, Figure S1: TEM images of Fe/Mn@γ−Al2O3 (A) and a partial enlargement (B), Figure S2: FT−IR spectra of γ−Al2O3, Fe@γ−Al2O3, Mn@γ−Al2O3, and Fe/Mn@γ−Al2O3, Figure S3: The survey spectrum of XPS spectra of NiCo2O4 before catalytic ozonation, Figure S4: The corresponding pore size distribution in nitrogen adsorption and desorption isotherms, Figure S5: Influence of (A, C) Fe/Mn@γ−Al2O3 dose and (B, D) the ratio of iron to manganese in the catalyst on DMP and 1−NP decomposition. Conditions: [O3] = 0.04 mM, [catalyst] = 50−300 mg L−1, the molar ratio of Fe/Mn = 0.3−1.5, [DMP]/[1−NP] = 50 μM, 2 mM phosphate buffer, Figure S6: The degradation of 1−NP under ozonation in presence of FFA. Conditions: [O3] = 0.02 mM, [FFA] = 100 mM, [1−NP] = 50 μM, pH = 7.0 ± 0.1 with 2 mM phosphate buffer, Figure S7: DMP and 1−NP removal rate constants during the reuse of Fe/Mn@γ−Al2O3 in catalytic ozonation system, Figure S8: LC−MS total ion chromatograms for DMP degradation products, Figure S9: LC−MS total ion chromatograms for 1−NP degradation products, Figure S10: Growth kinetics of DMP−OH adducts at 320 nm with different concentrations of DMP (0.02−0.3 mM) determined by using a laser flash photolysis system. The inset is the plot of the first−order formation rate constants of DMP−OH adducts vs. DMP concentrations, Figure S11: The removal of UV254 and DOC during the catalytic ozonation in the real water samples. Conditions: [O3] = 0.02 mM, [catalyst] = 200 mg L−1, the molar ratio of Fe/Mn = 1.5, Table S1 Dissolution of metal ions in Fe/Mn@γ−Al2O3 catalytic ozonation, Table S2: The steady−state concentrations of different reactive species, Table S3: The reactivity of different probes/quenchers with reactive species, Table S4: XPS parameters of high−resolution Mn 2p3/2, Fe 2p3/2, and O 1s regions, Table S5: Parameters of different real water samples.

Author Contributions

Conceptualization, C.L. and X.L.; methodology, C.L.; writing—original draft preparation, C.L.; writing—review and editing, X.L. and Y.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Acknowledgments

This work was supported by the National Natural Science Fund of China (NO. 21477039, No. U1401235).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Coha, M.; Farinelli, G.; Tiraferri, A.; Minella, M.; Vione, D. Advanced oxidation processes in the removal of organic substances from produced water: Potential, configurations, and research needs. Chem. Eng. J. 2021, 414, 128668. [Google Scholar] [CrossRef]
  2. Pearce, C.I.; Lloyd, J.R.; Guthrie, J.T. The removal of colour from textile wastewater using whole bacterial cells: A review. Dye. Pigment. 2003, 58, 179–196. [Google Scholar] [CrossRef]
  3. McMullan, G.; Meehan, C.; Conneely, A.; Kirby, N.; Robinson, T.; Nigam, P.; Banat, I.; Marchant, R.; Smyth, W.F. Microbial decolourisation and degradation of textile dyes. Appl. Microbiol. Biotechnol. 2001, 56, 81–87. [Google Scholar] [CrossRef] [PubMed]
  4. Lara-Ramos, J.A.; Figueroa Angulo, M.A.; Machuca-Martínez, F.; Mueses, M.A. Sensitivity Analysis of the Catalytic Ozonation under Different Kinetic Modeling Approaches in the Diclofenac Degradation. Water 2021, 13, 3003. [Google Scholar] [CrossRef]
  5. Beltrán, F.J.; Rivas, J.; Montero−De−Espinosa, R. Iron type catalysts for the ozonation of oxalic acid in water. Water Res. 2005, 39, 3553–3564. [Google Scholar] [CrossRef] [PubMed]
  6. Sun, Q.; Yang, L. The adsorption of basic dyes from aqueous solution on modified peat–resin particle. Water Res. 2003, 37, 1535–1544. [Google Scholar] [CrossRef]
  7. Kumar, M.; Sridhari, T.; Bhavani, K.; Dutta, P. Trends in color removal from textile mill effluents. Colourage 1998, 45, 25–34. [Google Scholar]
  8. Zugic, B.; Wang, L.C.; Heine, C.; Zakharov, D.N.; Lechner, B.A.J.; Stach, E.A.; Biener, J.; Salmeron, M.; Madix, R.J.; Friend, C.M. Dynamic restructuring drives catalytic activity on nanoporous gold−silver alloy catalysts. Nat. Mater. 2017, 16, 558–564. [Google Scholar] [CrossRef]
  9. Ghuge, S.P.; Saroha, A.K. Catalytic ozonation of dye industry effluent using mesoporous bimetallic Ru−Cu/SBA−15 catalyst. Process. Saf. Environ. Prot. 2018, 118, 125–132. [Google Scholar] [CrossRef]
  10. Tang, Y.M.; Pan, Z.Q.; Li, L.S. pH−insusceptible cobalt−manganese immobilizing mesoporous siliceous MCM−41 catalyst for ozonation of dimethyl phthalate. J. Colloid Interf. Sci. 2017, 508, 196–202. [Google Scholar] [CrossRef]
  11. Lu, X.H.; Zhang, Q.Y.; Yang, W.Q.; Li, X.K.; Zeng, L.X.; Li, L.S. Catalytic ozonation of 2,4−dichlorophenoxyacetic acid over novel Fe−Ni/AC. Rsc. Adv. 2015, 5, 10537–10545. [Google Scholar] [CrossRef]
  12. Xiong, Z.K.; Cao, J.Y.; Lai, B.; Yang, P. Comparative study on degradation of p−nitrophenol in aqueous solution by mFe/Cu/O3 and mFe0/O3 processes. J. Ind. Eng. Chem. 2018, 59, 196–207. [Google Scholar] [CrossRef]
  13. Ma, J.; Sui, M.; Zhang, T.; Guan, C. Effect of pH on MnOx/GAC catalyzed ozonation for degradation of nitrobenzene. Water Res. 2005, 39, 779–786. [Google Scholar] [CrossRef]
  14. Einaga, H.; Futamura, S. Catalytic oxidation of benzene with ozone over alumina−supported manganese oxides. J. Catal. 2004, 227, 304–312. [Google Scholar] [CrossRef]
  15. Beltran, F.J.; Rivas, F.J.; Montero−De−Espinosa, R. Ozone−Enhanced Oxidation of Oxalic Acid in Water with Cobalt Catalysts. 1. Homogeneous Catalytic Ozonation. Ind. Eng. Chem. Res. 2003, 42, 3210–3217. [Google Scholar] [CrossRef]
  16. Pines, D.S.; Reckhow, D.A. Effect of Dissolved Cobalt(II) on the Ozonation of Oxalic Acid. Environ. Sci. Technol. 2002, 36, 4046–4051. [Google Scholar] [CrossRef] [PubMed]
  17. Ma, J.; Graham, N.J.D. Degradation of atrazine by manganese−catalysed ozonationInfluence of radical scavengers. Water Res. 2000, 34, 3822–3828. [Google Scholar] [CrossRef]
  18. Cao, H.; Xing, L.; Wu, G.; Xie, Y.; Shi, S.; Zhang, Y.; Minakata, D.; Crittenden, J.C. Promoting effect of nitration modification on activated carbon in the catalytic ozonation of oxalic acid. Appl. Catal. B Environ. 2014, 146, 169–176. [Google Scholar] [CrossRef]
  19. Wang, Y.X.; Chen, L.L.; Cao, H.B.; Chi, Z.X.; Chen, C.M.; Duan, X.G.; Xie, Y.B.; Qi, F.; Song, W.Y.; Liu, J.; et al. Role of oxygen vacancies and Mn sites in hierarchical Mn2O3/LaMnO3−delta perovskite composites for aqueous organic pollutants decon−tamination. Appl. Catal. B−Environ. 2019, 245, 546–554. [Google Scholar] [CrossRef]
  20. Xu, B.J.; Xiao, T.C.; Yan, Z.F.; Sun, X.; Sloan, J.; Gonzalez−Cortes, S.L.; Alshahrani, F.; Green, M.L.H. Synthesis of mesoporous alumina with highly thermal stability using glucose template in aqueous system. Micropor. Mesopor. Mat. 2006, 91, 293–295. [Google Scholar] [CrossRef]
  21. Cho, K.; Rana, B.S.; Cho, D.W.; Beum, H.T.; Kim, C.H.; Kim, J.N. Catalytic removal of naphthenic acids over Co−Mo/gamma−Al2O3 catalyst to reduce total acid number (TAN) of highly acidic crude oil. Appl. Catal. A−Gen. 2020, 606, 117835. [Google Scholar] [CrossRef]
  22. Teng, Y.; Wang, X.-D.; Liao, J.-F.; Li, W.-G.; Chen, H.-Y.; Dong, Y.-J.; Kuang, D.-B. Atomically Thin Defect−Rich Fe−Mn−O Hybrid Nanosheets as High Efficient Electrocatalyst for Water Oxidation. Adv. Funct. Mater. 2018, 28, 28. [Google Scholar] [CrossRef]
  23. Naskar, M.K. Hydrothermal Synthesis of Petal−Like Alumina Flakes. J. Am. Ceram. Soc. 2009, 92, 2392–2395. [Google Scholar] [CrossRef]
  24. Zhu, L.; Pu, S.; Liu, K.; Zhu, T.; Lu, F.; Li, J. Preparation and characterizations of porous γ−Al2O3 nanoparticles. Mater. Lett. 2012, 83, 73–75. [Google Scholar] [CrossRef]
  25. Zhu, S.Y.; Zheng, X.S.; Li, D.T. Ozonation of naphthalene sulfonic acids in aqueous solutions. Part I: Elimination of COD, TOC and increase of their biodegradability. Water Res. 2002, 36, 1237–1243. [Google Scholar]
  26. Wang, P.; Matta, H.; Kuo, C.-H. Kinetics of ozonation of naphthalene and anthracene. J. Chin. I. Ch. E 1991, 22, 365–371. [Google Scholar]
  27. Hoigné, J.; Bader, H.; Haag, W.; Staehelin, J. Rate constants of reactions of ozone with organic and inorganic compounds in water—III. Inorganic compounds and radicals. Water Res. 1985, 19, 993–1004. [Google Scholar] [CrossRef]
  28. Huber, M.M.; Canonica, S.; Park, G.-Y.; von Gunten, U. Oxidation of Pharmaceuticals during Ozonation and Advanced Oxidation Processes. Environ. Sci. Technol. 2003, 37, 1016–1024. [Google Scholar] [CrossRef] [PubMed]
  29. Zhang, T.; Li, C.; Ma, J.; Tian, H.; Qiang, Z. Surface hydroxyl groups of synthetic α−FeOOH in promoting OH generation from aqueous ozone: Property and activity relationship. Appl. Catal. B Environ. 2008, 82, 131–137. [Google Scholar] [CrossRef]
  30. Ma, J.; Graham, N.J. Degradation of atrazine by manganese−catalysed ozonation: Influence of humic substances. Water Res. 1999, 33, 785–793. [Google Scholar] [CrossRef]
  31. Sui, M.; Liu, J.; Sheng, L. Mesoporous material supported manganese oxides (MnOx/MCM−41) catalytic ozonation of nitrobenzene in water. Appl. Catal. B Environ. 2011, 106, 195–203. [Google Scholar] [CrossRef]
  32. Yu, G.; Wang, Y.; Cao, H.; Zhao, H.; Xie, Y. Reactive Oxygen Species and Catalytic Active Sites in Heterogeneous Catalytic Ozonation for Water Purification. Environ. Sci. Technol. 2020, 54, 5931–5946. [Google Scholar] [CrossRef]
  33. Yao, C.C.D.; Haag, W.R. Rate constants for direct reactions of ozone with several drinking water contaminants. Water Res. 1991, 25, 761–773. [Google Scholar]
  34. Lei, Y.; Lu, J.; Zhu, M.; Xie, J.; Peng, S.; Zhu, C. Radical chemistry of diethyl phthalate oxidation via UV/peroxymonosulfate process: Roles of primary and secondary radicals. Chem. Eng. J. 2020, 379, 122339. [Google Scholar] [CrossRef]
  35. Kanodia, S.; Madhaven, V.; Schuler, R. Oxidation of naphthalene by radiolytically produced OH radicals. Int. J. Radiat. Appl. Instrumentation. Part. C. Radiat. Phys. Chem. 1988, 32, 661–664. [Google Scholar] [CrossRef]
  36. Darmanyan, A.P.; Moger, G. Interaction of singlet oxygen with nitroso− and phenol−type inhibitors. Magy. Kem. F. 1988, 94, 282–284. [Google Scholar]
  37. Zuo, Z.; Cai, Z.; Katsumura, Y.; Chitose, N.; Muroya, Y. Reinvestigation of the acid–base equilibrium of the (bi)carbonate radical and pH dependence of its reactivity with inorganic reactants. Radiat. Phys. Chem. 1999, 55, 15–23. [Google Scholar] [CrossRef]
  38. Buxton, G.V.; Greenstock, C.L.; Helman, W.P.; Ross, A.B. Critical Review of rate constants for reactions of hydrated electrons, hydrogen atoms and hydroxyl radicals (OH/O− in Aqueous Solution. J. Phys. Chem. Ref. Data 1988, 17, 513–886. [Google Scholar] [CrossRef] [Green Version]
  39. Zhao, L.; Sun, Z.; Ma, J. Novel Relationship between Hydroxyl Radical Initiation and Surface Group of Ceramic Honeycomb Supported Metals for the Catalytic Ozonation of Nitrobenzene in Aqueous Solution. Environ. Sci. Technol. 2009, 43, 4157–4163. [Google Scholar] [CrossRef]
  40. Ervens, B.; Gligorovski, S.; Herrmann, H. Temperature−dependent rate constants for hydroxyl radical reactions with organic compounds in aqueous solutions. Phys. Chem. Chem. Phys. 2003, 5, 1811–1824. [Google Scholar] [CrossRef]
  41. Liu, Z.; Chen, G.; Li, X.; Lu, X. Removal of rare earth elements by MnFe2O4 based mesoporous adsorbents: Synthesis, isotherms, kinetics, thermodynamics. J. Alloys Compd. 2021, 856, 158185. [Google Scholar] [CrossRef]
  42. Liu, Z.; Chen, G.; Hu, F.; Li, X. Synthesis of mesoporous magnetic MnFe2O4@CS−SiO2 microsphere and its adsorption performance of Zn2+ and MB studies. J. Environ. Manag. 2020, 263, 110377. [Google Scholar] [CrossRef]
  43. Xia, H.; Zhang, Z.; Liu, J.; Deng, Y.; Zhang, D.; Du, P.; Zhang, S.; Lu, X. Novel Fe−Mn−O nanosheets/wood carbon hybrid with tunable surface properties as a superior catalyst for Fenton−like oxidation. Appl. Catal. B Environ. 2019, 259, 118058. [Google Scholar] [CrossRef]
  44. Shen, T.; Su, W.; Yang, Q.; Ni, J.; Tong, S. Synergetic mechanism for basic and acid sites of MgMxOy (M = Fe, Mn) double oxides in catalytic ozonation of p−hydroxybenzoic acid and acetic acid. Appl. Catal. B Environ. 2020, 279, 119346. [Google Scholar] [CrossRef]
  45. Guan, S.; An, L.; Ashraf, S.; Zhang, L.; Liu, B.; Fan, Y.; Li, B. Oxygen vacancy excites Co3O4 nanocrystals embedded into carbon nitride for accelerated hydrogen generation. Appl. Catal. B Environ. 2020, 269, 118775. [Google Scholar] [CrossRef]
  46. Yan, H.; Lu, P.; Pan, Z.; Wang, X.; Zhang, Q.; Li, L. Ce/SBA−15 as a heterogeneous ozonation catalyst for efficient minerali−zation of dimethyl phthalate. J. Mol. Catal. A Chem. 2013, 377, 57–64. [Google Scholar] [CrossRef]
  47. Orge, C.A.; Órfão, J.J.; Pereira, M.F.; de Farias, A.M.D.; Fraga, M.A. Ceria and cerium-based mixed oxides as ozonation catalysts. Chem. Eng. J. 2012, 200-202, 499–505. [Google Scholar] [CrossRef]
  48. 4Rui, C.M.; Quinta−Ferreira, R.M. Catalytic ozonation of phenolic acids over a Mn–Ce–O catalyst. Appl. Catal. B Environ. Ment. 2009, 90, 268–277. [Google Scholar]
  49. Li, L.; Ye, W.; Zhang, Q.; Sun, F.; Li, X. Catalytic ozonation of dimethyl phthalate over cerium supported on activated car−bon. J. Hazard. Mater. 2009, 170, 411–416. [Google Scholar] [CrossRef]
  50. He, C.; Wang, Y.; Li, Z.; Huang, Y.; Liao, Y.; Xia, D.; Lee, S.-C. Facet Engineered α−MnO2 for Efficient Catalytic Ozonation of Odor CH3SH: Oxygen Vacancy−Induced Active Centers and Catalytic Mechanism. Environ. Sci. Technol. 2020, 54, 12771–12783. [Google Scholar] [CrossRef]
  51. He, D.; Wan, G.; Hao, H.; Chen, D.; Lu, J.; Zhang, L.; Liu, F.; Zhong, L.; He, S.; Luo, Y. Microwave−assisted rapid synthesis of CeO2 nanoparticles and its desulfuri−zation processes for CH3SH catalytic decomposition. Chem. Eng. J. Lausanne 2016, 289, 161–169. [Google Scholar] [CrossRef]
  52. Zhang, B.; Wang, L.; Zhang, Y.; Ding, Y.; Bi, Y. Ultrathin FeOOH Nanolayers with Abundant Oxygen Vacancies on BiVO4 Photoanodes for Efficient Water Oxidation. Angew. Chem. Int. Ed. 2018, 57, 2248–2252. [Google Scholar] [CrossRef] [PubMed]
  53. Gui, L.; Wang, Z.; Zhang, K.; He, B.; Liu, Y.; Zhou, W.; Xu, J.; Wang, Q.; Zhao, L. Oxygen vacancies−rich Ce0.9Gd0.1O2−δ decorated Pr0.5Ba0.5CoO3−δ bifunctional catalyst for efficient and long−lasting rechargeable Zn−air batteries. Appl. Catal. B Environ. 2020, 266, 118656. [Google Scholar] [CrossRef]
  54. Yan, S.; Zhang, X.; Shi, Y.; Zhang, H. Natural Fe−bearing manganese ore facilitating bioelectro−activation of peroxymonosulfate for bisphenol A oxidation. Chem. Eng. J. 2018, 354, 1120–1131. [Google Scholar] [CrossRef]
  55. Ikhlaq, A.; Kasprzyk−Hordern, B. Catalytic ozonation of chlorinated VOCs on ZSM−5 zeolites and alumina: Formation of chlorides. Appl. Catal. B Environ. 2017, 200, 274–282. [Google Scholar] [CrossRef] [Green Version]
  56. Ikhlaq, A.; Brown, D.; Kasprzyk−Hordern, B. Catalytic ozonation for the removal of organic contaminants in water on ZSM−5 zeolites. Appl. Catal. B Environ. 2014, 154–155, 110–122. [Google Scholar] [CrossRef]
  57. Ikhlaq, A.; Brown, D.; Kasprzyk−Hordern, B. Mechanisms of catalytic ozonation: An investigation into superoxide ion radical and hydrogen peroxide formation during catalytic ozonation on alumina and zeolites in water. Appl. Catal. B Environ. 2013, 129, 437–449. [Google Scholar] [CrossRef]
  58. Bing, J.; Hu, C.; Nie, Y.; Yang, M.; Qu, J. Mechanism of Catalytic Ozonation in Fe2O3/Al2O3@SBA−15 Aqueous Suspension for Destruction of Ibuprofen. Environ. Sci. Technol. 2015, 49, 1690–1697. [Google Scholar] [CrossRef] [PubMed]
  59. Yang, L.; Hu, C.; Nie, Y.; Qu, J. Catalytic Ozonation of Selected Pharmaceuticals over Mesoporous Alumina—Supported Manganese Oxide. Environ. Sci. Technol. 2009, 43, 2525–2529. [Google Scholar] [CrossRef] [PubMed]
  60. Huang, J.; Zhong, S.; Dai, Y.; Liu, C.-C.; Zhang, H.J. Effect of MnO2 Phase Structure on the Oxidative Reactivity toward Bisphenol A Degradation. Environ. Sci. Technol. 2018, 52, 11309–11318. [Google Scholar] [CrossRef] [PubMed]
  61. Roman, P.; Veltman, R.; Bijmans, M.F.M.; Keesman, K.J.; Janssen, A.J.H. Effect of Methanethiol Concentration on Sulfur Production in Biological Desulfurization Systems under Haloalkaline Conditions. Environ. Sci. Technol. 2015, 49, 9212–9221. [Google Scholar] [CrossRef]
  62. Buehler, R.E.; Staehelin, J.; Hoigne, J. Ozone decomposition in water studied by pulse radiolysis. 1. Perhydroxyl (HO2)/hyperoxide (O2−) and HO3/O3− as intermediates. J. Phys. Chem. 1984, 88, 2560–2564. [Google Scholar] [CrossRef]
  63. Long, L.; Zhao, J.; Yang, L.; Fu, M.; Wu, J.; Huang, B.; Ye, D. Room Temperature Catalytic Ozonation of Toluene over MnO2/Al2O3. Chin. J. Catal. 2011, 32, 904–916. [Google Scholar] [CrossRef]
  64. Liu, Y.; Feng, Y.; Zhang, Y.; Mao, S.; Wu, D.; Chu, H. Highly efficient degradation of dimethyl phthalate from Cu(II) and dimethyl phthalate wastewater by EDTA enhanced ozonation: Performance, intermediates and mechanism. J. Hazard. Mater. 2019, 366, 378–385. [Google Scholar] [CrossRef]
  65. An, T.; Gao, Y.; Li, G.; Kamat, P.V.; Peller, J.; Joyce, M.V. Kinetics and Mechanism of •OH Mediated Degradation of Dimethyl Phthalate in Aqueous Solution: Experimental and Theoretical Studies. Environ. Sci. Technol. 2014, 48, 641–648. [Google Scholar] [CrossRef] [PubMed]
  66. Zhang, Q.; Li, D.; Liu, Y.; Wang, H.; Zhang, C.; Huang, H.; He, Y.; Chen, X.; Du, Z.; Zheng, X. Potential anticancer activity of curcumin analogs containing sulfone on human cancer cells. Arch. Biol. Sci. 2016, 68, 125–133. [Google Scholar] [CrossRef]
  67. Wang, X.; Chen, C.; Li, J.; Wang, X. Ozone degradation of 1−naphthol on multiwalled carbon nanotubes/iron oxides and recycling of the adsorbent. Chem. Eng. J. 2015, 262, 1303–1310. [Google Scholar] [CrossRef]
  68. Acero, J.L.; Haderlein, S.B.; Schmidt, T.C.; Suter, M.J.-F.; von Gunten, U. MTBE Oxidation by Conventional Ozonation and the Combination Ozone/Hydrogen Peroxide: Efficiency of the Processes and Bromate Formation. Environ. Sci. Technol. 2001, 35, 4252–4259. [Google Scholar] [CrossRef]
  69. Madhavan, D.; Pitchumani, K. Photoreactions in clay media: Singlet oxygen oxidation of electron−rich substrates mediated by clay−bound dyes. J. Photochem. Photobiol. A Chem. 2002, 153, 205–210. [Google Scholar] [CrossRef]
  70. Kim, K.-J.; Hamill, W.H. Pulse radiolysis of concentrated aqueous solutions of chloride, iodide, and persulfate ions. J. Phys. Chem. 1976, 80, 2325–2330. [Google Scholar] [CrossRef]
  71. Jayson, G.G.; Parsons, B.J.; Swallow, A.J. Some simple, highly reactive, inorganic chlorine derivatives in aqueous solution. Their formation using pulses of radiation and their role in the mechanism of the Fricke dosimeter. J. Chem. Soc. Faraday Trans. 1 Phys. Chem. Condens. Phases 1973, 69, 1597–1607. [Google Scholar] [CrossRef]
  72. Buxton, G.V.; Wood, N.D.; Dyster, S. Ionisation constants of °OH and HO°2 in aqueous solution up to 200 °C. A pulse radiolysis study. J. Chem. Soc. Faraday Trans. 1 Phys. Chem. Condens. Phases 1988, 84, 1113–1121. [Google Scholar] [CrossRef]
  73. Kasprzyk−Hordern, B.; Ziółek, M.; Nawrocki, J. Catalytic ozonation and methods of enhancing molecular ozone reactions in water treatment. Appl. Catal. B Environ. 2003, 46, 639–669. [Google Scholar] [CrossRef]
  74. Huang, Y.; Luo, M.; Li, S.; Xia, D.; Tang, Z.; Hu, S.; Ye, S.; Sun, M.; He, C.; Shu, D. Efficient catalytic activity and bromate minimization over lattice oxygen−rich MnOOH nanorods in catalytic ozonation of bromide−containing organic pollutants: Lattice oxygen−directed redox cycle and bromate reduction. J. Hazard. Mater. 2021, 410, 124545. [Google Scholar] [CrossRef]
  75. Shahmahdi, N.; Dehghanzadeh, R.; Aslani, H.; Shokouhi, S.B. Performance evaluation of waste iron shavings (Fe0) for catalytic ozonation in removal of sulfamethoxazole from municipal wastewater treatment plant effluent in a batch mode pilot plant. Chem. Eng. J. 2020, 383, 123093. [Google Scholar] [CrossRef]
Figure 1. The SEM of γ−Al2O3 (A), Fe@γ−Al2O3 (B), Mn@γ−Al2O3 (C), Fe/Mn@γ−Al2O3 (D); the EDS spectra of the synthesized catalysts (E); the elemental mapping of the Fe/Mn@γ−Al2O3 (F); the powder XRD patterns (G); the N2 adsorption−desorption isotherms (H).
Figure 1. The SEM of γ−Al2O3 (A), Fe@γ−Al2O3 (B), Mn@γ−Al2O3 (C), Fe/Mn@γ−Al2O3 (D); the EDS spectra of the synthesized catalysts (E); the elemental mapping of the Fe/Mn@γ−Al2O3 (F); the powder XRD patterns (G); the N2 adsorption−desorption isotherms (H).
Water 14 00019 g001
Figure 2. The degradation of DMP (A) and 1−NP (B) during the catalytic ozonation. The degradation of DMP or 1−NP during the Fe/Mn@γ−Al2O3/O3 system in presence of different quenchers (NB/FFA/p−BQ, 100 mM) (C) or different probes (NB/FFA, 5 μM) (D). Conditions: [O3] = 0.02 mM, [catalyst] = 200 mg L−1, the molar ratio of Fe/Mn = 1.5, [DMP]/[1−NP] = 50 μM, pH = 7.0 ± 0.1 with 2 mM phosphate buffer. The probe test is detailed in Text S1.
Figure 2. The degradation of DMP (A) and 1−NP (B) during the catalytic ozonation. The degradation of DMP or 1−NP during the Fe/Mn@γ−Al2O3/O3 system in presence of different quenchers (NB/FFA/p−BQ, 100 mM) (C) or different probes (NB/FFA, 5 μM) (D). Conditions: [O3] = 0.02 mM, [catalyst] = 200 mg L−1, the molar ratio of Fe/Mn = 1.5, [DMP]/[1−NP] = 50 μM, pH = 7.0 ± 0.1 with 2 mM phosphate buffer. The probe test is detailed in Text S1.
Water 14 00019 g002
Figure 3. XPS spectra of Fe 2p (A), Mn 2s (B), and O 1s (C) for Fe/Mn@γ−Al2O3 before/after catalytic ozonation; oxygen vacancy of γ−Al2O3 and Fe/Mn@γ−Al2O3 (D); the EPR spectra of HO° and O2° using DMPO in water and methanol respectively, and the EPR spectra of 1O2 using TEMP from single ozonation and catalytic ozonation processes (E); the proposed catalytic ozonation mechanism of Fe/Mn@γ−Al2O3 (F).
Figure 3. XPS spectra of Fe 2p (A), Mn 2s (B), and O 1s (C) for Fe/Mn@γ−Al2O3 before/after catalytic ozonation; oxygen vacancy of γ−Al2O3 and Fe/Mn@γ−Al2O3 (D); the EPR spectra of HO° and O2° using DMPO in water and methanol respectively, and the EPR spectra of 1O2 using TEMP from single ozonation and catalytic ozonation processes (E); the proposed catalytic ozonation mechanism of Fe/Mn@γ−Al2O3 (F).
Water 14 00019 g003
Figure 4. Degradation pathways of DMP (A) and 1−NP (B) during Fe/Mn@γ−Al2O3 catalytic ozonation.
Figure 4. Degradation pathways of DMP (A) and 1−NP (B) during Fe/Mn@γ−Al2O3 catalytic ozonation.
Water 14 00019 g004
Figure 5. The degradation of DMP and 1−NP in presence of different ions (CO32−, SO42−, Cl, NO3, 2 mM) (A,D), at pH of 6−9 (B,E), in different real water samples (ultrapure water, PDIP effluent, WWTP effluent, RW) (C,F). Conditions: [O3] = 0.04 mM, [catalyst] = 200 mg L−1, the molar ratio of Fe/Mn = 1.5, [DMP]/[1−NP] = 50 μM, 2 mM phosphate buffer. Details are described in Text S3.
Figure 5. The degradation of DMP and 1−NP in presence of different ions (CO32−, SO42−, Cl, NO3, 2 mM) (A,D), at pH of 6−9 (B,E), in different real water samples (ultrapure water, PDIP effluent, WWTP effluent, RW) (C,F). Conditions: [O3] = 0.04 mM, [catalyst] = 200 mg L−1, the molar ratio of Fe/Mn = 1.5, [DMP]/[1−NP] = 50 μM, 2 mM phosphate buffer. Details are described in Text S3.
Water 14 00019 g005
Table 1. The estimated and experimental observed removal rate constants of DMP/1−NP.
Table 1. The estimated and experimental observed removal rate constants of DMP/1−NP.
CompoundRate Constant (M−1 s−1)Exp. k (s−1)Est. k * (s−1)
O3HO°1O2
DMP0.2 [33]3.7 × 109 #<103 [34]3.0 × 10−32.1 × 10−3
1−NP~102 $1.3 × 1010 [35]7.6 × 106 [36]3.3 × 10−22.2 × 10−2
* The observed rate constant of DMP and 1−NP (Est. k) were estimated as described in Text S1. # The second−order rate constant of HO° with DMP was determined using laser flash photolysis technology and competitive kinetics detailed in Text S2 and Figure S10. $ The second−order rate constant of O3 and 1−NP was estimated roughly based on the quench test and other naphthalenes.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liang, C.; Luo, X.; Hu, Y. Enhanced Ozone Oxidation by a Novel Fe/Mn@γ−Al2O3 Nanocatalyst: The Role of Hydroxyl Radical and Singlet Oxygen. Water 2022, 14, 19. https://doi.org/10.3390/w14010019

AMA Style

Liang C, Luo X, Hu Y. Enhanced Ozone Oxidation by a Novel Fe/Mn@γ−Al2O3 Nanocatalyst: The Role of Hydroxyl Radical and Singlet Oxygen. Water. 2022; 14(1):19. https://doi.org/10.3390/w14010019

Chicago/Turabian Style

Liang, Chen, Xinhao Luo, and Yongyou Hu. 2022. "Enhanced Ozone Oxidation by a Novel Fe/Mn@γ−Al2O3 Nanocatalyst: The Role of Hydroxyl Radical and Singlet Oxygen" Water 14, no. 1: 19. https://doi.org/10.3390/w14010019

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