Next Article in Journal
Multifunctional Loblolly Pine-Derived Superactivated Hydrochar: Effect of Hydrothermal Carbonization on Hydrogen and Electron Storage with Carbon Dioxide and Dye Removal
Next Article in Special Issue
Effect of Mn2+/Zn2+/Fe3+ Oxy(Hydroxide) Nanoparticles Doping onto Mg-Al-LDH on the Phosphate Removal Capacity from Simulated Wastewater
Previous Article in Journal
Development and Perspectives of Thermal Conductive Polymer Composites
Previous Article in Special Issue
Magnetic TiO2/CoFe2O4 Photocatalysts for Degradation of Organic Dyes and Pharmaceuticals without Oxidants
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Fabrication of Ternary Nanoparticles for Catalytic Ozonation to Treat Parabens: Mechanisms, Efficiency, and Effects on Ceratophyllum demersum L. and Eker Leiomyoma Tumor-3 Cells

by
Apiladda Pattanateeradetch
1,
Chainarong Sakulthaew
2,
Athaphon Angkaew
1,
Samak Sutjarit
2,
Thapanee Poompoung
2,
Yao-Tung Lin
3,
Clifford E. Harris
4,
Steve Comfort
5 and
Chanat Chokejaroenrat
1,*
1
Department of Environmental Technology and Management, Faculty of Environment, Kasetsart University, Bangkok 10900, Thailand
2
Department of Veterinary Technology, Faculty of Veterinary Technology, Kasetsart University, Bangkok 10900, Thailand
3
Department of Soil & Environmental Sciences, National Chung Hsing University, Taichung 402, Taiwan
4
Department of Chemistry and Biochemistry, Albion College, Albion, MI 49224, USA
5
School of Natural Resources, University of Nebraska-Lincoln, Lincoln, NE 68583, USA
*
Author to whom correspondence should be addressed.
Nanomaterials 2022, 12(20), 3573; https://doi.org/10.3390/nano12203573
Submission received: 26 August 2022 / Revised: 30 September 2022 / Accepted: 9 October 2022 / Published: 12 October 2022
(This article belongs to the Special Issue Advanced Nanocomposite Materials for Water and Wastewater Treatment)

Abstract

:
The use of parabens in personal care products can result in their leakage into water bodies, especially in public swimming pools with insufficient water treatment. We found that ferrite-based nanomaterials could catalytically enhance ozone efficiency through the production of reactive oxygen species. Our objective was to develop a catalytic ozonation system using ternary nanocomposites that could minimize the ozone supply while ensuring the treated water was acceptable for disposal into the environment. A ternary CuFe2O4/CuO/Fe2O3 nanocomposite (CF) delivered excellent degradation performance in catalytic ozonation systems for butylparaben (BP). By calcining with melamine, we obtained the CF/g-C3N4 (CFM) nanocomposite, which had excellent magnetic separation properties with slightly lower degradation efficiency than CF, due to possible self-agglomeration that reduced its electron capture ability. The presence of other constituent ions in synthetic wastewater and actual discharge water resulted in varying degradation rates due to the formation of secondary active radicals. 1O2 and O2 were the main dominant reactive species for BP degradation, which originated from the O3 adsorption that occurs on the CF≡Cu(I)–OH and CF≡Fe(III)–OH surface, and from the reaction with OH from indirect ozonation. Up to 50% of O3-treated water resulted in >80% ELT3 cell viability, the presence of well-adhered cells, and no effect on the young tip of Ceratophyllum demersum L. Overall, our results demonstrated that both materials could be potential catalysts for ozonation because of their excellent degrading performance and, consequently, their non-toxic by-products.

Graphical Abstract

1. Introduction

Parabens have been utilized widely as antimicrobial preservatives in many cosmetic and body care products for several decades. Swimming pools are a major sink of parabens from human activity, followed by municipal domestic sewage [1]. Although the use of chlorination and salinity in pool water should eliminate paraben residuals, Terasaki and Makino [2] found that parabens were, in turn, active ingredients for disinfection by-products (that is, chlorinated parabens). Given that human activity is a significant source of parabens, it is not surprising that urban areas are experiencing an increase in paraben-contaminated water. Up to 228 ng L−1 of parabens was detected in 39 different kinds of swimming pools from various places in China, such as hotel, residential, and commercial pools [3]. The improper discharge of pool water without any pretreatments has been reported and could consequently cause paraben contamination to any nearby natural receiving water and eventually to drinking water [4]. Some parabens were detected in 58 fish samples from 20 species collected from a fish market near Manila Bay, the Philippines [5]. While all parabens are categorized as endocrine disruption chemicals, butylparaben is more prone to induce vitellogenin production in male fish [6], and it has been reported to have higher estrogenic activity than other parabens [7]. This evidence confirms that a conventional water treatment system is insufficient; consequently, the development of vigorous treatment technology that results in non-toxic byproducts is needed.
Ozonation is an effective oxidant for rapidly transforming refractory pollutants due to its high oxidation potential (2.07 V). Despite such a high efficiency, using ozone alone (O3) can be overcome by other oxidation processes due to its oxidant solubility in water and the high energy requirement to provide sufficient O3 during utilization. Catalytic ozonation via a heterogeneous route has been used widely because it allows the recovery and reuse of materials, as well as the ability to enhance the mass transfer efficiency for treating target contaminants. During catalytic ozonation, utilization of O3 was reduced compared to the use of O3 alone [8]. The mechanisms of direct ozonation and indirect ozonation via the formation of OH have been well-established, but to a lesser extent than those of heterogeneous catalytic ozonation, especially with multiplex nanocomposites. This calls for these particular nanocomposites, which can strongly facilitate active radical generation (OH, O2•−, 1O2), to be fabricated by simple co-precipitation methods.
Several nanocomposites, such as transition metal ferrites (MFe2O4; M = Zn, Mn, Co, Ni, and Cu) and iron oxides (FexOy), have been shown to deliver outstanding performance in enhancing oxidation processes via electron transfer [9]. These materials have gained attention due to their natural abundance, low-cost fabrication methods, and chemical stability. The spinel ferrites exhibit crystalline stability and can support recyclability because they can provide magnetic separation properties and high catalytic activity [10]. Although FexOy is notable for providing higher magnetism, its activation ability towards PS activation is still unsatisfactory [11]. Yan et al. [12] found that catalytic O3 with α-Fe2O3 resulted in slower generation rates of active species compared to α-FeOOH and Fe3O4 due to the lack of surface Lewis acid sites on α-Fe2O3 that prevent the interaction of adsorbed O3 on the surface of Fe(III). Nonetheless, an Fe-based catalyst can serve as a suitable catalyst for O3 in degrading polychlorinated dibenzo-p-dioxins and polychlorinated dibenzofurans at low temperatures, providing that the catalyst has a high surface area of up to 300 m2 g−1 [13]. Several researchers have highlighted the benefit of using a multiplex catalyst in oxidation processes that can empower electron mobility by adding more than one element. For example, molybdenum oxide was used as a catalyst promoter in the CuO/Al2O3 synthesis process, resulting in the formation of Cu-Mo-Al complexes to enhance the ozonation ability of dimethyl sulfide [14]. Li et al. [15] replaced active sites of iron oxides that had a weaker catalytic activity with other metal ions with various valence states, such as Mn, to become a magnetic Fe3O4-MnO3 nanocomposite to enhance the activation of peroxymonosulfate. Complex nanocomposites of bimetal oxides deposited on Zn-Fe silicate were found to trigger chain reactions of O3 to form more hydroxyl radicals during catalytic ozonation [16].
While these complexes are not commonly used in a conventional water treatment system, this can be altered by introducing the right dopants into the nanocomposites and utilizing their unique properties. The construction of nanocomposites with a metal-free catalytic material, such as graphitic carbon nitride (g-C3N4), is also an interesting choice for providing superior magnetic separation properties [17]; however, its catalytic ozonation performance, especially with parabens, remains unknown. Given that several synthesis methods are available, co-precipitation is a very convenient approach and does not require any toxic solvent during synthesis, which is also feasible for large-scale production.
In the current study, several transition metals were selectively investigated in the search for better-mixed transition metal oxides for higher catalytic ozonation with a strong magnetic separation properties. The physicochemical characteristics of the material before and after ozonation were compared side-by-side to explain their mechanistic behavior. To ensure the system applicability, the effects from different water matrices and other environmental variables, which would have an impact on the paraben degradation rates, were also investigated. The reaction mechanism was determined through scavenging experiments using various chemical probes and changes in element composition using X-ray photoelectron spectroscopy (XPS). The transformation of the selected parabens during degradation was explained in detail. The toxicity of O3-treated was evaluated water on Ceratophyllum demersum L. As parabens have been known to possess endocrine-disruptive properties, we also evaluated their impact on Eker Leiomyoma Tumor-3 cells following exposure to the different proportions of O3-treated water.

2. Materials and Methods

Chemical vendors and manufacturers, analytical procedures, catalyst characteristics, and preparation of ELT3 cell culture are detailed in the Supplementary Materials (Sections S1–S4).

2.1. Catalyst Synthesis

All metal ferrite nanocomposites (MF) were synthesized through a co-precipitation method followed by calcination, as modified from our previous study [17]. A mixture of 0.2 M MSO4 (CuSO4, NiCl2·6H2O, CoCl2·6H2O, MnSO4·H2O), 0.4 M FeCl3·6H2O, and distilled water (DI) at a ratio of 1:1:2 (v/v) was heated under a constant stirring at 80 °C for 1 h. A heated solution of 8 M NaOH was used to slowly adjust the pH to 10.5 within 1 h and then the mixture was continuously stirred for 1 h. To obtain uniform particles, the mixture at room temperature was washed with DI water until a neutral pH was reached. The remaining particles were washed twice with ethanol prior to drying overnight in a hot-air oven at 95 °C. The final CuFe2O4/CuO/Fe2O3 nanocomposites (CF) were calcined at 550 °C for 3 h. For CF/g-C3N4 synthesis (CFM), melamine was mixed with the oven-dried particles at a ratio of 2:1 (w/w) before calcining at 550 °C. All materials were kept in a desiccator until use.

2.2. Typical Experimental Setup

All experiments were performed in a 250 mL beaker. Mixed parabens, consisting of methylparaben (MP), ethylparaben (EP), propylparaben (PP), and butylparaben (BP), at a concentration of 10 µM each were prepared in DI water and used altogether in one batch. As BP is frequently found in natural receiving water, BP degradation was selectively presented, unless stated otherwise. The as-obtained nanocomposites (CF or CFM) of 20 mg were added into the 100 mL mixed parabens solution, equivalent to 0.2 g L−1. Before purging with O3, each solution was stirred for 30 min to ensure no adsorption had occurred on the material surface. An ozone flux of 10 mg min−1 was continuously bubbled along with magnetic stirring. Samples were taken periodically at a designated time within 40 min, involving filtering with a 0.45 µm PTFE syringe filter before transferring to vials for further paraben analysis.

2.3. Operational and Influential Effect Experiments

The influential effect experiments were performed using the above protocol while each effect was individually varied. We selected chlorine concentration (1–4 mg L−1), hardness (90.28–361.10 mg L−1 as CaCO3), and bicarbonate (HCO3; 100–140 mg L−1) as variables that may be present in discharge water. Trichloroisocyanuric acid (TCCA) was used for chlorine variation content as it is the most-selected type of disinfectant for swimming pools and can provide up to 90% available chlorine in the form of hypochlorite (OCl).
Because real discharge pool water is usually mixed with naturally occurring other constituents, including a high loading of organic matter and several kinds of personal care products, the effect of water matrices was determine in four different kinds of water matrices (DC: discharge chlorine pool water, DS: discharge saline pool water, NW: natural receiving water, and MW: municipal wastewater). The DC and DS were collected from a public swimming pool located in Kasetsart University, Bangkok, Thailand. The NW was collected from a natural water basin (13.854640, 100.570256) that received treated pool water, while the MW was collected from an overcrowded residential area located in Chatujak district, Bangkok, Thailand. The basic water qualities were shown in the Supplementary Materials (Table S1). Unlike typical experiments, the BP was selected as the main target to represent all parabens; therefore, the initial concentration was raised by 5-fold.

2.4. Scavenging Experiments

To evaluate the possible degradation mechanism of BP in the catalytic ozonation system, we spiked a radical scavenger—tertbutyl alcohol (TBA), p- benzoquinone (pBQ), furfuryl alcohol (FFA), or dimethyl sulfoxide (DMSO)—in each experimental unit and maintained the final concentration ratio of BP to these scavengers at 1:5000, 1:500, 1:500, and 1:5000, respectively. These scavengers served as sources of the hydroxyl radical (OH), superoxide radical (O2•−), singlet oxygen (1O2), and electrons (e).

2.5. Recyclability Testing

Both CF and CFM were evaluated for their recyclability following the catalytic ozonation process achieved from the five-cycle experiments. The typical experiments were followed as discussed in Section 2.2, but the BP was selected as the main target contaminant. Freshly contaminated water was introduced into the O3 reactor for up to 8 h. Sampling was made every 30–60 min for residual BP in the solution. Catalysts were reused after manual (CF) or magnetic (CFM) separation from the solution using a neodymium magnet. After the fifth cycle, samples were oven-dried at 105 °C overnight prior to further crystallinity and surface analysis using FTIR, XRD, and XPS. The loss of material during each cycle was also monitored.

2.6. Toxicological Assays

Ceratophyllum demersum L. plants were purchased from a local market in Bangkok, Thailand. The top portion (6–13 cm) of each plant was selected with a dry weight of approximately 2.7–2.9 g. Each portion was grown in an 800 mL PVC bowl filled up with 500 mL of effluent water, which was O3-treated water at various proportions (0–100%). This water was mixed with water and used in matrices for growing the selected Ceratophyllum demersum L. The plants were collected for physical evaluation after 5 days of exposure to the effluents.
As parabens can interfere with the reproductive hormones, we used an in vitro bioassay based on the Eker Leiomyoma Tumor-3/Eker uterine leiomyoma cell line (ELT3) purchased from ATCC (Manassas, VA, USA). This cytotoxicity evaluation is rapid and reliable, and most importantly, negates the need to conduct whole-organism-level tests (Reference). Cell viability tests were performed for O3-treated water at varying amounts with DI water to ensure that there was no induced cytotoxicity in these cells. Details of culture methods are provided in Supplementary Materials (Section S4).

3. Results and Discussion

3.1. Nanocomposite Selection

Various types of treatments [(1) individual oxidant (O3, UV); (2) synthesized nanocomposite alone (MF; M(II) = Co(II), Cu(II), Mn(II), and Ni(II)); (3) O3/MF; (4) UV/MF; and (5) O3/UV/MF] were tested to compare the observed pseudo first-order rate constant (kobs) and %removal at 20 min (BP in Figure 1; MP, EP, and PP in Figures S1–S3). Overall, BP degradation efficiency was in the order of O3/MF > O3/UV/MF > O3 alone > UV alone > UV/MF > MF alone. Ninety-five percent removal of BP was observed within 40 min using O3 alone, while only up to 8% removal was observed when MnF was used, even in the dark, indicating that only slight adsorption occurred on the material surface (Figure 1A). Although the UV-assisted O3 and O3/UV/MF showed slightly better performance than the others, the enhancement of catalytic ozonation from the UV was unlikely. Some transition metal ferrite nanocomposites used in this study were identified as having a narrow bandgap, which is not an appropriate energy band for UV light. Aside from the UV radiation that only accounts for less than 5% of the solar spectrum, using any light may suffer from the catalysts or any suspended solids that can increase impeding light penetration. For example, under our UV/MF condition, the BP removal was reduced by 60%, and the kobs was decreased by 3-fold (Figure 1B).
Although CoF/O3 resulted in faster degradation performance than CuF/O3 in all parabens, the use of CoF resulted in evidence of agglomeration during utilization, which would easily hinder magnetic separation and may further obstruct the reusability of the material. Morphologically speaking, the spinel CuFe2O4 is more structurally stable and should provide a better magnetic response after simple modification than the CoFe2O4 [18]. The outstanding attribute of CuFe2O4 is its smaller pore volume expansion and its ability to donate more electrons, making it favorable for electron transfer [19]. Therefore, CuF was selected as the more suitable catalyst for enhancing the ozonation process.

3.2. CF Nanocomposite Properties

3.2.1. Physical and Chemical Characteristics

From a catalytic activity standpoint, bare CuFe2O4 could be used as a backbone for mixed transition metal oxides [20]. Here, we characterized the nanocomposites and revealed that it was the ternary catalyst of CuFe2O4/CuO/Fe2O3, denoted as CF, which was later decorated with g-C3N4, denoted as CFM, while CF-a and CFM-a represented the used material in the O3-system to remove parabens. Using the crystallography open database (COD; No 9011012), diffraction peaks of 18.58°, 30.23°, 35.87°, 37.38°, 41.10°, 44.07°, 54.29°, and 64.21°, which are ascribed to the (101), (112), (211), (311), (222), (113), (400), (422), and (440) lattice planes of CuFe2O4, are more intensified with CFM than with the CF (Figure 2A). This finding indicated that the recrystallization was completed during the calcination processes. The small peaks observed at 35.87°, 38.98°, 48.96°, and 68.25°, corresponding to the reflection of the (110), (111), (202), and (113) planes of CuO, showed that both materials (CF-a versus CFM-a) lost CuO content after O3 treatment, even after undergoing the thermal process [21]. The hematite phase is indicated by the diffraction peaks of 33.40° and 35.6°, which are ascribed to the (104) and (110) planes on both materials. These peaks confirmed that as-obtained CF and CFM can be successfully fabricated using co-precipitation methods.
It appeared that pure-g-C3N4 had diffraction peaks of 27.4°, which also weakly appeared in the CFM nanocomposites, indicating that the calcination process may have restructured the material (Figure 2A). While Wang et al. [22] showed that losing g-C3N4 after calcination was possible, we believe that the calcination process causes the dispersion of CuFe2O4 particles and may overshadow the diffraction peaks of g-C3N4. Overall, the XRD diffraction peaks of CF-a showed slightly higher intensities than for CFM-a, indicating that ozonation may reduce material crystallinity. With a higher O3 flux, recyclability of CFM may be affected, and therefore using CF would probably be more tangible for practical operation.
The morphological features of CF showed its fluffy texture consisting of nanoparticles (15–60 nm), which are interconnected among the three transition metal oxides (Figure 2B,C). High-magnification of the CF TEM images revealed a mixture of the platelet-like hematite particles that are tightly attached to the spherical shape of CuFe2O4 and CuO (Figure 2F,G). Unlike CF, the CFM nanoparticles tended to agglomerate, resulting from recrystallization during the calcination stage with melamine. The majority of CFM particles were irregular platelets, with surrounding smaller cubes and spheres with an average size of 20.84 nm, which corresponded with the average size obtained from the XRD results (Figure 2D,E, Tables S2 and S3). In addition, CFM-a had a higher remaining content of CuFe2O4 than CF-a due to its agglomeration, as g-C3N4 can serve as a particle substrate. While both N2 adsorption/desorption isotherms confirmed their uniformly mesoporous material, changes following calcination were observed for the pore size (~37.5 to 32.7 nm) and the BET surface area (24.13 to 2.91 m2 g−1), indicating that there might be some compaction and distortion of the crystal structures from the use of the g-C3N4 substrate (Figure S4).
The measured lattice fringes of 0.250, 0.298, and 0.483 nm from the CFM HRTEM image corresponded to the CuFe2O4 planes (Figure 2I). In addition, these lattice spacings of CuO and Fe2O3 of 0.253 nm and 0.251 nm, which are consistent with the (002) and (110) plane of XRD results, also revealed that g-C3N4 and CF nanoparticles were unified with structural densification and uniform distribution (Figure 2G,I). The lack of the CuFe2O4 lattice plane in the CF HRTEM image was also in accordance with the small portion of CuFe2O4 revealed from its XRD results (Figure 2A), signifying the absence of CuFe2O4 during the CF synthesis and confirming that g-C3N4 enabled the magnetism of material via CuFe2O4 deposition (Figure 2G).
The formation of CuO on each nanoparticle was confirmed by the stretching vibration of the Cu-O bands on the FTIR spectra at 437.84, 516.92, and 590.22 cm−1 [23] (Figure 3A). The intensities of the non-magnetic CuO peaks showed that the surface functional groups were unchanged after use. By comparing the Raman spectra of the CF and CFM nanoparticles, CFM had six characteristics bands of 224, 291, 416, 540, 613, and 699 cm−1 that were well-defined as cubic CuFe2O4 nanoparticles [24], while those of CF showed a much weaker response (Figure 3B). This finding corresponded to our earlier discussion on the reappearance of CuFe2O4 in the CFM crystallization (Figure 2A and Figure 3A). Furthermore, some of the CF Raman bands were quite broad with a much weaker band than those of CFM, confirming it was composed of ternary nanocomposites.
All the CF and CFM samples displayed narrow hysteresis loops with the specific saturation magnetization (Ms) value of CFM (12.8 emu·g−1) being 4.5 times higher than that of CF (2.8 emu·g−1) (Figure 3C). This could be associated with the presence of CuFe2O4 in the nanocomposites in the XRD pattern at 18.58°, 37.38°, and 44.07°, and a stronger band at 570 cm−1 attributed to the Fe-O for CFM (Figure 2A). Here, the g-C3N4 increased the inter-particle interaction of these ternary catalyst particles enabling its adherence between microspheres [25]. Only slight changes in the magneticity of the used CFM were observed (~1.68 emu·g−1), indicating the material had strong magnetization after several uses and remained intact. The overall results revealed that CFM was more efficiently separated from the external magnetic source making it ideal as an environmental catalyst that could prevent secondary pollutants.

3.2.2. Element Compositions of CF Nanocomposites

The full-survey scan of the XPS spectra showed the existence of C 1s, O 1s, Fe 2p, and Cu 2p, with no evidence of impurities, while the fine scan revealed the elemental valence states and the bonding configuration of these elements (Figure S5 and Figure 4). Five major peaks at binding energies of 283.9 eV, 285.0 eV, 286.1 eV, 288.0 eV, and 288.9 eV in the C1s spectra were assigned to C–C, C–O, C–O–C, C=O, and O=C–O [16]. Additionally, the major peaks at binding energies of 530 eV, 531.7 eV, and 532.8 eV, were attributed to surface lattice oxygen (Olatt; O2−), surface adsorbed oxygen in the hydroxyl group (OOH), and the oxygen-containing functional group (Oads) (Figure 4). Although only a weak characteristic peak of g-C3N4 was noticeable in the XRD pattern of CFM-a, we identified no trace of N composition in the N XPS spectra (Figure 2A and Figure 4). We expected that the g-C3N4 structure might have been deformed to other NOx during the calcination process after facilitating the recrystallization process of the CFM.
All Fe spectrum displayed dominant doublet peaks located at binding energies of 710.2 eV (Fe 2p3/2) and 724.3 eV (Fe 2p1/2), which were attributed to Fe(II) and Fe(III) (Figure 4). Similar to the Cu spectrum, two major peaks were observed at binding energies of 934.6 eV (Cu 2p3/2) and 954.7 eV (Cu 2p1/2), which were attributed to Cu(II), and a small amount of Cu(I) at binding energies of 933.0 eV (Cu 2p3/2) and 953.2 eV (Cu 2p1/2), which were attributed to Cu(I). Both these results confirmed that the Cu(II) species made up the main proportion in the catalyst. After the deconvolutions of Fe 2p and Cu 2p, the changes in Fe and Cu were clear, indicating that electron transfer had occurred on the catalyst surface following the catalytic ozonation. Notably, there was evidence of the presence of the absorber copper hydroxide (Cu(OH)2) at binding energies of 936.1 eV and 956.3 eV, which have arisen from the use of NaOH in the co-precipitation process.

3.3. Enhanced Ozone Degradation Performance

The catalytic ozonation performance of the CF and CFM nanoparticles were investigated by the removal of total parabens under varying catalyst doses (0.1–0.3 g L−1) (BP in Figure 5; MP, EP, and PP in Figure S6). Because initial CFM tests to enhance ozonation performance revealed insignificantly difference than that of CF in a single batch experiment (data not shown), the CFM was used to compare again in the application study (Section 3.5: Material recyclability). Compared to using O3 alone, the O3/CF increased the removal efficiency by approximately 20% (inset of Figure 5). Only slight adsorption of BP on the CF surface was observed (<6%). This slight adsorption was expected for ternary inorganic nanocomposites unlike other magnetic porous carbon-based material that were enriched in adsorption sites and were prone to adsorb organic contaminants [26]. Among all parabens tested, we found that BP had higher adsorption activity and faster degradation rates, which could have been due to the length of the ester chains and its possession of more unsaturated C=C bonds that are preferably attacked by the O3 molecule [27]. Here, by varying dosage, about 77.2%, 81.1%, and 91.4% of BP were degraded at 10 min of reaction (Figure 5) because the larger specific area created more reactive sites between O3 molecules and catalysts, which accelerated the generation of reactive species. Given that the self-scavenging effect has been observed in several studies using higher oxidant doses or excessive amounts of catalyst dosage, we did not observe that phenomenon. This was likely due to the short-lived O3 and the beneficial use of the ternary nanocomposites. Therefore, the CF of 0.3 g L−1 was selected as the optimum dosage for other successive experiments.

3.4. Influential Effects on Degradation Efficiency

In this part of the study, the BP degradation efficiency following the heterogeneous catalytic oxidation process was tested against the presence of common environmental factors or different water matrices that can only link to the paraben-contaminated water. Given that an alkaline pH can increase ozonation efficiency by several magnitudes from the presence of available OH, we did not select this variable due to its working condition being close to pool water and natural water environmental pH conditions [28]. Following treatment, we reported on the changes in kobs compared to the control experiment (no effect) (Figure 6). The existence of OCl caused negative effects on BP degradation by decreasing kobs by up to 30% due to the presence of secondary chlorine radicals (Figure 6A). Haag and Hoigné [29] found that O3 can directly react with OCl and yield Cl as a major product (Equation (1)). The generation of reactive chlorine radicals (Cl) and the dichlorine radical (Cl2•−) may occur during the ozonation in the presence of Cl with available OH from the indirect ozonation (Equations (2)–(4)) [30]. Although Yang et al. [31] reported that both chlorine radicals were prone to attacking phenol-containing compounds, the reaction between chlorine radicals and Cl may be faster than that with parabens. Once the secondary chlorine radicals (Cl2•−) were formed, they were likely self-scavenged into chloride and BP degradation then ceased (Equations (4) and (5)) [32].
O3 + OCl → 2O2 + Cl
OH + Cl → ClOH•−
ClOH•− + H+ → Cl + H2O
Cl + Cl ↔ Cl2•−
Cl2•− + Cl2•− → 2Cl + Cl2
Hardness is one of the dominant water parameters usually present in natural water. Here, we varied the amount of hardness and found that the presence of hardness negatively affected the catalytic ozonation of BP (Figure 6A). During catalytic ozonation, oxygenated compounds may occur, which initiate the reaction of abundant cations in very hard water, promoting coagulation and possibly suppressing BP degradation [33].
Because TCCA and other water constituents can contribute to the available alkalinity of pool water, the alkalinity is required to be maintained at 60–180 mg L−1 as CaCO3 [34]. Our results show that the remaining alkalinity could accelerate the BP degradation efficiency using ozonation, which was possibly due to the subsequent increase in pH (Figure 6A). In addition, our CF was prepared in NaOH medium which could also result in the strong intensity of the –OH group on the CF surface (Figure 3A). We believe that this sudden increase in BP degradation mainly resulted from the OH generation originating from the OH and O3 molecules to generate more of other secondary active radicals, such as O2•− and HO2 free radicals (HO2) (Equations (6)–(9)) [35].
O3 + OH → HO2 + O2
HO2 + H+ → H2O2
HO2 + O3 → O2•− + OH + O2
H2O2 + O3 → HO2 + OH + O2
In real-world applications, more than one type of paraben has been detected in the discharge water, so we used all four parabens as the starting substrate, but the plotted kobs represents BP only (Figure 6B). The results showed that the paraben degradation rates were higher than the control in all types of water matrices, indicating the presence of several constituents in the water matrices that may have contributed to the generation of secondary active radicals from the specific characteristics of the humic substance surface [36], some of which can facilitate the formation of reactive radicals. Although Haman et al. [37] reported that BP molecules were sensitive to residual chlorines, sudden decreases in BP concentration were possible. However, this was unlikely to occur, because chlorine residuals are susceptible to self-deterioration from maturation basins. The Koc of PP and BP was larger than those of MP and EP by two-fold, suggesting that both PP and BP could be expected to sorb onto suspended solids and the sediment surface [38]. Therefore, we believe that the decrease was also likely a result of the adsorption effect that occurred simultaneously.

3.5. Material Recyclability

Temporal changes in the BP concentrations showed that BP removal was completed in each cycle in 150–180 s for CF and in 300–360 s for CFM (Figure 7). Notably, the BP degradation in this five-cycle experiment was much faster than in the batch experiment (Figure 5 vs. Figure 7) for two reasons: (1) BP was not the sole contaminant in the batch experiment and (2) the size discrepancy of the contact tank, which may have required a larger O3 flux for large-scale application.
The %removal from the CF-O3 system remained the same in cycles 3–5, while that of the CFM-O3 system caused fluctuation in the %removal for both selected timelines. Either way, lower catalytic ozonation performance was observed for CFM. A mass-to-mass comparison revealed that CF provided much better performance than CFM even though manually removing CF may result in a slight loss of catalyst compared to using a neodymium magnet to remove CFM from the solution. During ozonation, the CFM aggregation may cause a reduction in reactive sites for O3 molecule transfer to generate reactive oxygen species, thereby causing lower degradation performance. Despite that the most beneficial aspect of using CFM was its retractability using a neodymium magnet during the manufacturing process, an additional calcination step was required. Therefore, CF would be more suitable for large-scale production as the additional calcination step is not needed.

3.6. Proposed Mechanisms

3.6.1. Enhanced Ozonation Mechanisms

In our experimental setup, two possible catalytic mechanisms emerge: (1) direct/indirect ozonation and (2) heterogeneous catalytic ozonation. To explain the catalytic mechanisms, we used a scavenging experiment to confirm the type of active species using several chemical probes and the XPS spectrum to elucidate the changes in element compositions (Figure 4 and Figure 8A). The results showed that the BP degradation rates were retarded in the following order TBA < DMSO << pBQ < FFA, indicating that BP degradation was mainly from 1O2 and O2•−, which corresponded to the lowest BP degradation rates, and OH was the least predominant radical species. High proportions of oxygen components in the O 1s spectrum of CF-a and CFM-a were observed, indicating strong hydroxylation on the catalyst surface during the ozonation process, which confirmed that these radicals resulted from initiation from the surface hydroxyl groups arising from the indirect ozonation. Electron transfer could also be observed from the Cu 2p and Fe 2p spectra (Cu(II)/Cu(I) and Fe(III)/Fe(II)), which can lead to the abstraction of hydrogen to produce reactive superoxide oxygens (Figure 4) [39]. This coincided with the decrease in Olatt (O2−) and the increase in Oads caused by the adsorption of O-containing molecules or some small amounts of parabens and their degradates that could be chemically adsorbed before degrading (Figure 4). This also resulted in an increase in oxygen vacancy (Ov) which later formed –OH2 on the surface (CF–OH2) (Equation (10)). Then, the O3 adsorption and decomposition occurred on the CF surface, causing a release of O3 free radicals (HO3), conversion of O2−, regeneration of Ov, and subsequently, the release a new generation of O2− (Equations (11)–(15); [37]) This was a benefit of the existence of several metal oxides in the nanocomposite that had both strong Lewis acid sites and oxygen vacancy sites [40].
Ov + H2O → CF–OH2
CF–OH2 + O3 → CF–OH + HO3
CF–OH + O3 → CF–HO2 + O2
CF–HO2 + O3 → CF–O2 + HO3
CF–O2 → O2 + Ov
O2 + 4e → 2O2
According to the above analysis on the regeneration of O2−, the electron transfer mechanisms of our ternary nanocomposite during catalytic ozonation are proposed (Figure 8B). Initially, O3 adsorption on the surface hydroxyl group occurred on the metal ions (CF≡Cu(I)–OH and CF≡Fe(III)–OH). Changes in the redox pairs of Fe(II)/Fe(III) and Cu(I)/Cu(II) were more obvious for CF/CF-a than for CFM/CFM-a, indicating that more electron capture occurred on the CF surface than for CFM and that it may have caused higher catalytic activity for CF in the ozonation process (Figure 4). The loss of an electron from Cu(I) on the CuO could result in the formation of Cu(II) along with the decomposition of O3, which later generated both OH and O2 and later self-reacted to form 1O2, as the major active species (Figure 8A) (Equations (16)–(19)) [41,42,43]. However, in addition, the OH was also responsible for this oxidation process (Equation (18)). This was confirmed by the changes in the BP degradation rates (Figure 8A) and the changes in the IR spectra of CF-a at 3400 cm−1 (Figure 3A).
CF≡Cu(I)–OH + O3 → CF≡Cu(I)–OH–O3
O2− + CF≡Cu(I)–OH–O3 → CF≡Cu(II) + HO2•− + O2
O3 + HO2•−OH + O2 + O2
OH + O2 → 1O2 + OH
On the other hand, O3 interaction with Fe(III) on the Fe2O3 surface would result in OHads and O3 (Equation (20)) [44]. Then, electron transfer occurred from Fe(III) to Fe(II), which corresponded to the increase in Fe(II) from the Fe 2p spectrum (Equation (21); Figure 4). Once these O species were formed with the available O3 or OH, they could react and later form both 1O2 and O2•− (Equations (19), (22)–(24)) [17,43].
CF≡Fe(III)–OH + O3 → CF≡Fe(III)–O3 + OHads
CF≡Fe(III)–O3 + OH → CF≡Fe(II) + HO2 + O2
HO2 → O2 + H+
O2 + HO2 → 1O2 + OOH
HO2 + HO2 → 1O2 + H2O2
Cu(II) + O2 → Cu(I) + O2
In CFM, the regenerated O2 could, in turn, reduce Cu(II) to Cu(I) via Equation (25); then, there was less O2 for the generation of 1O2 to begin the BP degradation, thus resulting in lower degradation efficiency [41].

3.6.2. Degradates of Paraben Catalytic Ozonation

Aliquot samples of BP following catalyzed-O3 treatment at 30, 90, and 150 min were analyzed using LC–MS to determine the degradation products. All initial concentrations were five times higher than the degradation efficiency experiment to warrant the emergence of degradation products. Chromatograms showed that the BP peak (tR 20.12 min) decreased with time and only a trace amount of BP was found at 150 min. The mass spectra of the various BP degradates showed the molecular ion peaks of m/z 209.3 for 1-hydroxy BPB (tR 24.32 min) (Figure 8C) and m/z 137.2 for 4-hydroxybenzoic acid (tR 16.73 min). The 1-hydroxy BPB was highest at 90 min of reaction and continued to decrease at 150 min (Figure 8D). 4-Hydroxybenzoic acid was detected at 30, 90, and 150 min.
There are two likely mechanisms for the degradation of BP based on literature precedent. Previous work [27,39,45] supports Mechanism 1, featuring oxidation by O3, OH, and/or O2, at the alkyl chain to give intermediate 1, and subsequent hydrolysis to the 4-hydroxy benzoic acid 2 (Figure 9). Asgari et al. [46] report that both O3 and OH can give Mechanism 2, where the aromatic ring is first oxidized to give intermediate 3, which could then hydrolyze to give 2,4-dihydroxybenzoic acid 4 (Figure 9). However, Asgari’s work showed continued attack by OH at the aromatic ring of 4 to give polyhydroxylated degradation products without the hydrolysis of the butyl ester group.
The present data only support Mechanism 1, in that m/z signals supporting both compound 1 and compound 2 are observed, and are consistent with previous work. While the first step of Mechanism 2 could be supported by assignment of the signal at m/z 209.3 to potential intermediate 3, this is not supported by other evidence. Compound 4 would be the obvious next product under the same conditions, but is not observed. No other polyhydroxylated compounds are observed. Thus, Mechanism 1 is best supported by the present evidence.

3.7. Toxicological Assessment

In general, treated wastewater is mixed with natural receiving water in varying amounts dependent upon the water treatment regime. To estimate the toxicological effects of these effluents, we used the changes in physical observation of Ceratophyllum demersum L. and the dose-dependent cytotoxicity in ELT3 cells exposed to the various components of O3-treated water (0–100%). The control, healthy Ceratophyllum demersum L., showed a uniform green color for the entire length of the stem; in contrast, at 5 days of sampling, the older parts of the plants were slightly detached, and the young tip was more resistant with 50% of treated water than with 100% of treated water (Figure 10A–E), perhaps because the tested water did not significantly harm the plants, supporting catalytic O3 as a clean treatment technology. The detachment of Ceratophyllum demersum L. leaves is the primary indicator of its death following exposure to toxic chemicals, but this incidence usually occurs after long-term exposure (>14 d) or a sudden high dose of chemicals. Chen et al. [47] observed negative changes when the plants were submerged in 40–80 mM of lead solution after 21 days. In our case, O3 and its reactive species have a very short half-life; therefore, the direct effect on the plants was minimal.
The question remaining is whether this O3-treated water is safe for other living organisms that are far more sensitive to Ceratophyllum demersum L. In an attempt to answer this question, we used ELT3 cells as representative cells. No samples exerted general cytotoxicity (<80% cell viability) at 0–20% of O3-treated water (Figure 10F). In addition, we found that the cytotoxicity potential of the O3-treated water was higher than for the untreated water by approximately 15% (Figure 10F), indicating a negative consequence from the O3 treatment. The viability discrepancy was more pronounced for a smaller component of O3-treated water, which corresponded to the attachment of cells seen using microscopic evaluation (Figure 10G–L). At 0–40% of the O3-treated mixture, the attachment of cells was clear, representing their better viability. This observation corresponded to the viability percentage, which implied that by decreasing O3-treated water from 60% to 20%, the viability reached 96%. Given that the bioassay results showed that at >40% O3-treated water, effluent samples can be toxic to tested cells, while, to a lesser extent, cell deterioration was unlikely.
Although parabens have been reported to cause cytotoxic effects on human peripheral lymphocytes, it was likely to occur within 1–2 days of exposure [48]. Cells were reported to undertake mitochondrial dysfunction following exposure to various concentrations of parabens [49]. Under our experimental conditions, we believe the 4-hydroxybenzoic acid and 4-hydroxybenzoic decreased viability of ELT3 cells in a dose-dependence manner. Given that the parent parabens were rapidly removed, the cause of algal morphological alteration and an increase in cell mortality were from the paraben degradates, with the toxicological effects being reduced by these three important factors: (1) a larger dose of O3 flux, (2) a longer contact time, and (3) a smaller proportion of treated water. A sequential release of treated water is suggested, and the cell deterioration mechanisms should be thoroughly investigated in the future study.

4. Conclusions

In the present study, the ternary CuFe2O4/CuO/Fe2O3 nanocomposites (CF), synthesized using co-precipitation methods, were used for the catalytic ozonation process to treat paraben-contaminated water. Constructing graphitic carbon nitride (g-C3N4) in the composites by the calcination of melamine (CFM) during synthesis resulted in improved material magnetism. The catalytic ozonation efficiency and recyclability performance of both materials were revealed. The O3/CF system could effectively remove all parabens from contaminated water within minutes, while CF displayed better performance for degrading parabens than CFM; however, the CFM had better reusability after five cycles. The presence of organic and inorganic constituents had both positive and negative effects on butylparaben degradation rates due to the presence of secondary active radicals. The main dominant reactive oxygen species in the system were 1O2 and O2, which are initiated from the OH generated through indirect ozonation. Through the scavenging experiment and changes in element composition from XPS, the catalytic ozonation of CuFe2O4/CuO/Fe2O3 nanocomposite catalytic ozonation was also proposed. The O3/CF system has proven that it is a clean technology and can be applied to conventional water treatment, as it shows no effects on both Ceratophyllum demersum L. and Eker Leiomyoma Tumor-3 cells at up to 40% composition with clean water. For larger-scale treatment in a real-world application, the O3:CF system is recommended, because it provides a promising catalytic ozonation process without the need for a large O3 supply and without toxic by-products.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano12203573/s1, Section S1. Chemical vendors and manufacturers; Section S2. Analytical procedures; Section S3. Catalyst characteristics; Section S4. Preparation of ELT3 cell culture; Figure S1. Degradation of methylparaben (MP) under different treatment systems; Figure S2. Degradation of ethylparaben (EP) under different treatment systems; Figure S3. Degradation of propylparaben (PP) under different treatment systems; Figure S4. Particle size distribution of: (A) CF, (B) CFM, and (C) N2 adsorption/desorption isotherms of both catalysts; Figure S5. Survey scan of XPS spectra of the CuFe2O4/CuO/Fe2O3 ternary nanocomposites before and after catalytic ozonation; Figure S6. Temporal changes in parabens at varying CF dosages in the O3/CF system Tables: Table S1. Basic water quality of real-word swimming pool water; Table S2. Average particle size obtained from XRD analysis; Table S3. Average particle size obtained from particle size distribution using TEM analysis.

Author Contributions

Conceptualization, C.S. and C.C.; Data curation, A.P.; Formal analysis, A.A.; Funding acquisition, C.C.; Investigation, A.P.; Methodology, A.P., T.P. and C.C.; Project administration, C.C.; Resources, C.S.; Supervision, C.C.; Validation, C.S., A.A., C.E.H. and C.C.; Visualization, A.P. and C.E.H.; Writing—original draft, A.P., C.S. and C.C.; Writing—review and editing, C.S., S.S., Y.-T.L., C.E.H., S.C. and C.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research is funded by Kasetsart University through the Graduate School Fellowship Program and the Office of the Ministry of Higher Education, Science, Research and Innovation; and the Thailand Science Research and Innovation through Kasetsart University Reinventing University Program 2021.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The authors confirm that the data supporting the findings of this study are available within the article.

Acknowledgments

This research is funded by Kasetsart University through the Graduate School Fellowship Program and the Office of the Ministry of Higher Education, Science, Research and Innovation; and the Thailand Science Research and Innovation through Kasetsart University Reinventing University Program 2021. The authors also gratefully acknowledge the Faculty of Environment and the Faculty of Veterinary Technology, Kasetsart University, for providing facility support and access to analytical instruments. The Chulabhorn Walailak Swimming Pool, Kasetsart University provided pool water samples.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Reichert, G.; Mizukawa, A.; Antonelli, J.; de Almeida Brehm Goulart, F.; Filippe, T.C.; de Azevedo, J.C.R. Determination of Parabens, Triclosan, and Lipid Regulators in a Subtropical Urban River: Effects of Urban Occupation. Water Air Soil Pollut. 2020, 231, 133. [Google Scholar] [CrossRef]
  2. Terasaki, M.; Makino, M. Determination of chlorinated by-products of parabens in swimming pool water. Int. J. Environ. Anal. Chem. 2008, 88, 911–922. [Google Scholar] [CrossRef]
  3. Li, W.; Shi, Y.; Gao, L.; Liu, J.; Cai, Y. Occurrence and human exposure of parabens and their chlorinated derivatives in swimming pools. Environ. Sci. Pollut. Res. 2015, 22, 17987–17997. [Google Scholar] [CrossRef] [PubMed]
  4. Dominguez, J.R.; Gonzalez, T.; Cuerda-Correa, E.M.; Muñoz-Peña, M.J. Combating paraben pollution in surface waters with a variety of photocatalyzed systems: Looking for the most efficient technology. Open Chem. 2019, 17, 1317–1327. [Google Scholar] [CrossRef]
  5. Ramaswamy, B.R.; Kim, J.-W.; Isobe, T.; Chang, K.-H.; Amano, A.; Miller, T.W.; Siringan, F.P.; Tanabe, S. Determination of preservative and antimicrobial compounds in fish from Manila Bay, Philippines using ultra high performance liquid chromatography tandem mass spectrometry, and assessment of human dietary exposure. J. Hazard. Mater. 2011, 192, 1739–1745. [Google Scholar] [CrossRef] [PubMed]
  6. Bjerregaard, P.; Hansen, P.R.; Larsen, K.J.; Erratico, C.; Korsgaard, B.; Holbech, H. Vitellogenin as a biomarker for estrogenic effects in brown trout, Salmo trutta: Laboratory and field investigations. Environ. Toxicol. Chem. 2008, 27, 2387–2396. [Google Scholar] [CrossRef] [PubMed]
  7. Mizuno, H.; Hirai, H.; Kawai, S.; Nishida, T. Removal of estrogenic activity of iso-butylparaben and n-butylparaben by laccase in the presence of 1-hydroxybenzotriazole. Biodegradation 2009, 20, 533–539. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Zhang, M.; Yin, D.; Guo, J.; Wu, H.; Gong, M.; Feng, X. Ternary catalyst Mn-Fe-Ce/Al2O3 for the ozonation of phenol pollutant: Performance and mechanism. Environ. Sci. Pollut. Res. 2021, 28, 32921–32932. [Google Scholar] [CrossRef]
  9. Yao, Y.; Lu, F.; Zhu, Y.; Wei, F.; Liu, X.; Lian, C.; Wang, S. Magnetic core–shell CuFe2O4@C3N4 hybrids for visible light photocatalysis of Orange II. J. Hazard. Mater. 2015, 297, 224–233. [Google Scholar] [CrossRef]
  10. Du, Y.; Ma, W.; Liu, P.; Zou, B.; Ma, J. Magnetic CoFe2O4 nanoparticles supported on titanate nanotubes (CoFe2O4/TNTs) as a novel heterogeneous catalyst for peroxymonosulfate activation and degradation of organic pollutants. J. Hazard. Mater. 2016, 308, 58–66. [Google Scholar] [CrossRef] [PubMed]
  11. Wang, Z.; Ma, H.; Zhang, C.; Feng, J.; Pu, S.; Ren, Y.; Wang, Y. Enhanced catalytic ozonation treatment of dibutyl phthalate enabled by porous magnetic Ag-doped ferrospinel MnFe2O4 materials: Performance and mechanism. Chem. Eng. J. 2018, 354, 42–52. [Google Scholar] [CrossRef]
  12. Yan, L.; Bing, J.; Wu, H. The behavior of ozone on different iron oxides surface sites in water. Sci. Rep. 2019, 9, 14752. [Google Scholar] [CrossRef] [Green Version]
  13. Wang, H.C.; Chang, S.H.; Hung, P.C.; Hwang, J.F.; Chang, M.B. Catalytic oxidation of gaseous PCDD/Fs with ozone over iron oxide catalysts. Chemosphere 2008, 71, 388–397. [Google Scholar] [CrossRef]
  14. Devulapelli, V.G.; Sahle-Demessie, E. Catalytic oxidation of dimethyl sulfide with ozone: Effects of promoter and physico-chemical properties of metal oxide catalysts. Appl. Catal. A Gen. 2008, 348, 86–93. [Google Scholar] [CrossRef]
  15. Li, R.; Cai, M.; Xie, Z.; Zhang, Q.; Zeng, Y.; Liu, H.; Liu, G.; Lv, W. Construction of heterostructured CuFe2O4/g-C3N4 nanocomposite as an efficient visible light photocatalyst with peroxydisulfate for the organic oxidation. Appl. Catal. B Environ. 2019, 244, 974–982. [Google Scholar] [CrossRef]
  16. Liu, B.; Song, W.; Wu, H.; Liu, Z.; Teng, Y.; Sun, Y.; Xu, Y.; Zheng, H. Degradation of norfloxacin with peroxymonosulfate activated by nanoconfinement Co3O4@CNT nanocomposite. Chem. Eng. J. 2020, 398, 125498. [Google Scholar] [CrossRef]
  17. Angkaew, A.; Chokejaroenrat, C.; Sakulthaew, C.; Mao, J.; Watcharatharapong, T.; Watcharenwong, A.; Imman, S.; Suriyachai, N.; Kreetachat, T. Two facile synthesis routes for magnetic recoverable MnFe2O4/g-C3N4 nanocomposites to enhance visible light photo-Fenton activity for methylene blue degradation. J. Environ. Chem. Eng. 2021, 9, 105621. [Google Scholar] [CrossRef]
  18. Xian, G.; Kong, S.; Li, Q.; Zhang, G.; Zhou, N.; Du, H.; Niu, L. Synthesis of Spinel Ferrite MFe2O4 (M = Co, Cu, Mn, and Zn) for Persulfate Activation to Remove Aqueous Organics: Effects of M-Site Metal and Synthetic Method. Front. Chem. 2020, 8, 177. [Google Scholar] [CrossRef] [PubMed]
  19. Duan, X.; Su, C.; Miao, J.; Zhong, Y.; Shao, Z.; Wang, S.; Sun, H. Insights into perovskite-catalyzed peroxymonosulfate activation: Maneuverable cobalt sites for promoted evolution of sulfate radicals. Appl. Catal. B Environ. 2018, 220, 626–634. [Google Scholar] [CrossRef]
  20. Ma, Q.; Zhang, H.; Zhang, X.; Li, B.; Guo, R.; Cheng, Q.; Cheng, X. Synthesis of magnetic CuO/MnFe2O4 nanocompisite and its high activity for degradation of levofloxacin by activation of persulfate. Chem. Eng. J. 2019, 360, 848–860. [Google Scholar] [CrossRef]
  21. Dhanda, R.; Kidwai, M. Magnetically separable CuFe2O4/reduced graphene oxide nanocomposites: As a highly active catalyst for solvent free oxidative coupling of amines to imines. RSC Adv. 2016, 6, 53430–53437. [Google Scholar] [CrossRef]
  22. Wang, L.; Ma, X.; Huang, G.; Lian, R.; Huang, J.; She, H.; Wang, Q. Construction of ternary CuO/CuFe2O4/g-C3N4 composite and its enhanced photocatalytic degradation of tetracycline hydrochloride with persulfate under simulated sunlight. J. Environ. Sci. 2022, 112, 59–70. [Google Scholar] [CrossRef]
  23. Zhao, W.; Zhang, S.; Ding, J.; Deng, Z.; Guo, L.; Zhong, Q. Enhanced catalytic ozonation for NOx removal with CuFe2O4 nanoparticles and mechanism analysis. J. Mol. Catal. A Chem. 2016, 424, 153–161. [Google Scholar] [CrossRef]
  24. Yadav, R.S.; Kuřitka, I.; Vilcakova, J.; Havlica, J.; Masilko, J.; Kalina, L.; Tkacz, J.; Hajdúchová, M.; Enev, V. Structural, dielectric, electrical and magnetic properties of CuFe2O4 nanoparticles synthesized by honey mediated sol–gel combustion method and annealing effect. J. Mater. Sci. Mater. Electron. 2017, 28, 6245–6261. [Google Scholar] [CrossRef]
  25. Zhang, R.; Wang, Y.; Zhang, Z.; Cao, J. Highly Sensitive Acetone Gas Sensor Based on g-C3N4 Decorated MgFe2O4 Porous Microspheres Composites. Sensors 2018, 18, 2211. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Feng, Z.; Zhai, X.; Sun, T. Sustainable and efficient removal of paraben, oxytetracycline and metronidazole using magnetic porous biochar composite prepared by one step pyrolysis. Sep. Purif. Technol. 2022, 293, 121120. [Google Scholar] [CrossRef]
  27. Tay, K.S.; Rahman, N.A.; Abas, M.R.B. Ozonation of parabens in aqueous solution: Kinetics and mechanism of degradation. Chemosphere 2010, 81, 1446–1453. [Google Scholar] [CrossRef] [PubMed]
  28. Altmann, J.; Ruhl, A.S.; Zietzschmann, F.; Jekel, M. Direct comparison of ozonation and adsorption onto powdered activated carbon for micropollutant removal in advanced wastewater treatment. Water Res. 2014, 55, 185–193. [Google Scholar] [CrossRef]
  29. Haag, W.R.; Hoigné, J. Ozonation of water containing chlorine or chloramines. Reaction products and kinetics. Water Res. 1983, 17, 1397–1402. [Google Scholar] [CrossRef]
  30. Hoigné, J.; Bader, H. The role of hydroxyl radical reactions in ozonation processes in aqueous solutions. Water Res. 1976, 10, 377–386. [Google Scholar] [CrossRef]
  31. Yang, L.; Zhang, Y.; Liu, X.; Jiang, X.; Zhang, Z.; Zhang, T.; Zhang, L. The investigation of synergistic and competitive interaction between dye Congo red and methyl blue on magnetic MnFe2O4. Chem. Eng. J. 2014, 246, 88–96. [Google Scholar] [CrossRef]
  32. Ji, Y.; Dong, C.; Kong, D.; Lu, J.; Zhou, Q. Heat-activated persulfate oxidation of atrazine: Implications for remediation of groundwater contaminated by herbicides. Chem. Eng. J. 2015, 263, 45–54. [Google Scholar] [CrossRef]
  33. Sadrnourmohamadi, M.; Gorczyca, B. Effects of ozone as a stand-alone and coagulation-aid treatment on the reduction of trihalomethanes precursors from high DOC and hardness water. Water Res. 2015, 73, 171–180. [Google Scholar] [CrossRef]
  34. APSP. ANSI/APSP/ICC-11 2019 American National Standard for Water Quality in Public Pools and Spas. Available online: https://issuu.com/thephta/docs/apsp-11_2019 (accessed on 25 March 2022).
  35. Kwon, B.G.; Lee, J.H. A Kinetic Method for HO2/O2•− Determination in Advanced Oxidation Processes. Anal. Chem. 2004, 76, 6359–6364. [Google Scholar] [CrossRef]
  36. Sharma, J.; Mishra, I.M.; Kumar, V. Degradation and mineralization of Bisphenol A (BPA) in aqueous solution using advanced oxidation processes: UV/H2O2 and UV/S2O82− oxidation systems. J. Environ. Manag. 2015, 156, 266–275. [Google Scholar] [CrossRef] [PubMed]
  37. Haman, C.; Dauchy, X.; Rosin, C.; Munoz, J.-F. Occurrence, fate and behavior of parabens in aquatic environments: A review. Water Res. 2015, 68, 1–11. [Google Scholar] [CrossRef] [PubMed]
  38. Gasperi, J.; Geara, D.; Lorgeoux, C.; Bressy, A.; Zedek, S.; Rocher, V.; El Samrani, A.; Chebbo, G.; Moilleron, R. First assessment of triclosan, triclocarban and paraben mass loads at a very large regional scale: Case of Paris conurbation (France). Sci. Total Environ. 2014, 493, 854–861. [Google Scholar] [CrossRef] [Green Version]
  39. Daghrir, R.; Dimboukou-Mpira, A.; Seyhi, B.; Drogui, P. Photosonochemical degradation of butyl-paraben: Optimization, toxicity and kinetic studies. Sci. Total Environ. 2014, 490, 223–234. [Google Scholar] [CrossRef] [Green Version]
  40. Luo, X.; Su, T.; Xie, X.; Qin, Z.; Ji, H. The Adsorption of Ozone on the Solid Catalyst Surface and the Catalytic Reaction Mechanism for Organic Components. ChemistrySelect 2020, 5, 15092–15116. [Google Scholar] [CrossRef]
  41. Liu, Y.; Wu, D.; Peng, S.; Feng, Y.; Liu, Z. Enhanced mineralization of dimethyl phthalate by heterogeneous ozonation over nanostructured Cu-Fe-O surfaces: Synergistic effect and radical chain reactions. Sep. Purif. Technol. 2019, 209, 588–597. [Google Scholar] [CrossRef]
  42. Wang, X.; Min, J.; Li, S.; Zhu, X.; Cao, X.; Yuan, S.; Zuo, X.; Deng, X. Sono-assisted synthesis of CuO nanorods–graphene oxide as a synergistic activator of persulfate for bisphenol A removal. J. Environ. Chem. Eng. 2018, 6, 4078–4083. [Google Scholar] [CrossRef]
  43. Soltani, T.; Lee, B.-K. Improving heterogeneous photo-Fenton catalytic degradation of toluene under visible light irradiation through Ba-doping in BiFeO3 nanoparticles. J. Mol. Catal. A Chem. 2016, 425, 199–207. [Google Scholar] [CrossRef]
  44. 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]
  45. Chuang, L.C.; Luo, C.H. Photocatalytic degradation of parabens in aquatic environment: Kinetics and degradation pathway. Kinet. Catal. 2015, 56, 412–418. [Google Scholar] [CrossRef]
  46. Asgari, E.; Esrafili, A.; Rostami, R.; Farzadkia, M. O3, O3/UV and O3/UV/ZnO for abatement of parabens in aqueous solutions: Effect of operational parameters and mineralization/biodegradability improvement. Process Saf. Environ. Prot. 2019, 125, 238–250. [Google Scholar] [CrossRef]
  47. Chen, M.; Zhang, L.-L.; Li, J.; He, X.-J.; Cai, J.-C. Bioaccumulation and tolerance characteristics of a submerged plant (Ceratophyllum demersum L.) exposed to toxic metal lead. Ecotoxicol. Environ. Saf. 2015, 122, 313–321. [Google Scholar] [CrossRef] [PubMed]
  48. Güzel Bayülken, D.; Ayaz Tüylü, B. In vitro genotoxic and cytotoxic effects of some paraben esters on human peripheral lymphocytes. Drug Chem. Toxicol. 2019, 42, 386–393. [Google Scholar] [CrossRef] [PubMed]
  49. Nakagawa, Y.; Moldéus, P. Mechanism of p-Hydroxybenzoate Ester-induced Mitochondrial Dysfunction and Cytotoxicity in Isolated Rat Hepatocytes. Biochem. Pharmacol. 1998, 55, 1907–1914. [Google Scholar] [CrossRef]
Figure 1. Degradation of butylparaben (BP) under different treatment systems: (A) degradation kinetics, and (B) observed rate constant (kobs) and %removal at 20 min (Catalyst: 0.2 g L−1, [BPo]: 10 µM, pH 7, 25 °C). Error bars indicate ± standard deviation.
Figure 1. Degradation of butylparaben (BP) under different treatment systems: (A) degradation kinetics, and (B) observed rate constant (kobs) and %removal at 20 min (Catalyst: 0.2 g L−1, [BPo]: 10 µM, pH 7, 25 °C). Error bars indicate ± standard deviation.
Nanomaterials 12 03573 g001
Figure 2. CuFe2O4/CuO/Fe2O3 ternary nanocomposites: (A) XRD patterns of catalysts before and after catalytic ozonation, (B,C) CF SEM images, (D,E) CFM SEM images, (F,G) CF TEM images, (H,I) CFM TEM images.
Figure 2. CuFe2O4/CuO/Fe2O3 ternary nanocomposites: (A) XRD patterns of catalysts before and after catalytic ozonation, (B,C) CF SEM images, (D,E) CFM SEM images, (F,G) CF TEM images, (H,I) CFM TEM images.
Nanomaterials 12 03573 g002
Figure 3. CuFe2O4/CuO/Fe2O3 ternary nanocomposites: (A) FTIR spectra before and after catalytic ozonation, (B) Raman spectra, and (C) vibrating sample magnetometer hysteresis loops before and after catalytic ozonation.
Figure 3. CuFe2O4/CuO/Fe2O3 ternary nanocomposites: (A) FTIR spectra before and after catalytic ozonation, (B) Raman spectra, and (C) vibrating sample magnetometer hysteresis loops before and after catalytic ozonation.
Nanomaterials 12 03573 g003
Figure 4. XPS spectra of CuFe2O4/CuO/Fe2O3 ternary nanocomposites before (CF, CFM) and after (CF-a, CFM-a) catalytic ozonation (C 1s, O 1s, Fe 2p, and Cu 2p).
Figure 4. XPS spectra of CuFe2O4/CuO/Fe2O3 ternary nanocomposites before (CF, CFM) and after (CF-a, CFM-a) catalytic ozonation (C 1s, O 1s, Fe 2p, and Cu 2p).
Nanomaterials 12 03573 g004
Figure 5. (A) Temporal changes of BP at different CF dosages in the O3/CF system (Catalyst: 0.2 g L−1, [each parabeno]: 10 µM, pH 7, 25 °C). (B) Paraben removal efficiency at 20 min under different treatments (O3, CF, and O3/CF), and Error bars indicate ± standard deviation.
Figure 5. (A) Temporal changes of BP at different CF dosages in the O3/CF system (Catalyst: 0.2 g L−1, [each parabeno]: 10 µM, pH 7, 25 °C). (B) Paraben removal efficiency at 20 min under different treatments (O3, CF, and O3/CF), and Error bars indicate ± standard deviation.
Nanomaterials 12 03573 g005
Figure 6. (A) Comparison of observed rate constant (kobs) with different influential effects, (B) Temporal changes in BP at different water matrices (catalyst: 0.2 g L−1, [BPo]: 10 µM, pH 7, 25 °C). Error bars indicate ± standard deviation.
Figure 6. (A) Comparison of observed rate constant (kobs) with different influential effects, (B) Temporal changes in BP at different water matrices (catalyst: 0.2 g L−1, [BPo]: 10 µM, pH 7, 25 °C). Error bars indicate ± standard deviation.
Nanomaterials 12 03573 g006
Figure 7. Degradation of BP following five cycles of pumping BP-contaminated water through the catalytic ozonation system using CF or CFM (each catalyst: 0.2 g L−1, [BPo]: 10 µM, pH 7, 25 °C).
Figure 7. Degradation of BP following five cycles of pumping BP-contaminated water through the catalytic ozonation system using CF or CFM (each catalyst: 0.2 g L−1, [BPo]: 10 µM, pH 7, 25 °C).
Nanomaterials 12 03573 g007
Figure 8. Following catalytic ozonation using O3/CF system: (A) effects of four scavenging agents on BP removal (catalyst: 0.2 g L−1, [BPo]: 10 µM, pH 7, 25 °C), and (B) proposed catalytic mechanism of O3/CF for BP degradation; (C,D) mass spectra of BP major transformation products from LC–MS at (C) 90 min and (D) 150 min.
Figure 8. Following catalytic ozonation using O3/CF system: (A) effects of four scavenging agents on BP removal (catalyst: 0.2 g L−1, [BPo]: 10 µM, pH 7, 25 °C), and (B) proposed catalytic mechanism of O3/CF for BP degradation; (C,D) mass spectra of BP major transformation products from LC–MS at (C) 90 min and (D) 150 min.
Nanomaterials 12 03573 g008
Figure 9. Proposed reaction pathways for the catalytic ozonation of butyl paraben.
Figure 9. Proposed reaction pathways for the catalytic ozonation of butyl paraben.
Nanomaterials 12 03573 g009
Figure 10. After exposure of samples to different percentages of O3-treated water: (AE) visual observation of Ceratophyllum demersum L., (F) cytotoxicity assay of Eker Leiomyoma Tumor-3 (ELT3) cell line, and (GL) optical microscopy images of ELT3 cells.
Figure 10. After exposure of samples to different percentages of O3-treated water: (AE) visual observation of Ceratophyllum demersum L., (F) cytotoxicity assay of Eker Leiomyoma Tumor-3 (ELT3) cell line, and (GL) optical microscopy images of ELT3 cells.
Nanomaterials 12 03573 g010
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Pattanateeradetch, A.; Sakulthaew, C.; Angkaew, A.; Sutjarit, S.; Poompoung, T.; Lin, Y.-T.; Harris, C.E.; Comfort, S.; Chokejaroenrat, C. Fabrication of Ternary Nanoparticles for Catalytic Ozonation to Treat Parabens: Mechanisms, Efficiency, and Effects on Ceratophyllum demersum L. and Eker Leiomyoma Tumor-3 Cells. Nanomaterials 2022, 12, 3573. https://doi.org/10.3390/nano12203573

AMA Style

Pattanateeradetch A, Sakulthaew C, Angkaew A, Sutjarit S, Poompoung T, Lin Y-T, Harris CE, Comfort S, Chokejaroenrat C. Fabrication of Ternary Nanoparticles for Catalytic Ozonation to Treat Parabens: Mechanisms, Efficiency, and Effects on Ceratophyllum demersum L. and Eker Leiomyoma Tumor-3 Cells. Nanomaterials. 2022; 12(20):3573. https://doi.org/10.3390/nano12203573

Chicago/Turabian Style

Pattanateeradetch, Apiladda, Chainarong Sakulthaew, Athaphon Angkaew, Samak Sutjarit, Thapanee Poompoung, Yao-Tung Lin, Clifford E. Harris, Steve Comfort, and Chanat Chokejaroenrat. 2022. "Fabrication of Ternary Nanoparticles for Catalytic Ozonation to Treat Parabens: Mechanisms, Efficiency, and Effects on Ceratophyllum demersum L. and Eker Leiomyoma Tumor-3 Cells" Nanomaterials 12, no. 20: 3573. https://doi.org/10.3390/nano12203573

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