Acessibilidade / Reportar erro

Study of Fe3O4/PS System in Degrading BPA in Aqueous Solution

Abstract

The degradation of bisphenol A (BPA) by Fe3O4/persulfate system was investigated in aqueous solution. The influences of the initial concentrations of Fe3O4, persulfate (PS) and BPA, pH value, and initial reaction temperature on BPA removal were studied. The radical species was investigated by adding excessive dose of scavenger (methanol (MeOH) and tert-butanol (TBA)) into Fe3O4/PS system for the purpose of radical scavenging. The degradation products of BPA were detected by gas chromatography-mass spectrometry (GC-MS). The recyclability of Fe3O4 was also evaluated. The BPA removal rate of 80.7% was achieved under the following conditions: [BPA]0 = 1 mg L-1, [PS]0 = 0.2 mM, [Fe3O4]0 = 0.1 g L-1, T0 = 20 ± 1 ºC, pH0 = 6.8 ± 0.2. The results confirmed that the main free radicals in the reaction process were sulfate radicals, followed by hydroxyl radicals. Some intermediate products of BPA degradation, such as phenols, benzoquinones and benzoic acid were identified by GC-MS.

Keywords:
sulfate radicals; Fe3O4; bisphenol A; hydroxyl radicals; advanced oxidation processes


Introduction

Bisphenol A (2,2-bis(4-hydroxyphenyl)propane, BPA), a white solid, has been widely used as a ubiquitous intermediate in manufacturing polycarbonate plastics, epoxy resins and polysulfone.11 Vandenberg, L. N.; Hauser, R.; Marcus, M.; Olea, N.; Welshons, W. V.; Reprod. Toxicol. 2007, 24, 139.,22 Pahigian, J. M.; Zuo, Y.; Chemosphere 2018, 207, 469. Therefore, BPA is present in many daily supplies, such as toys, bottles, food and beverage packaging, water supply pipes, and the polymers used in dental treatment.33 Li, C.; Wang, Z.; Yang, Y. J.; Liu, J.; Mao, X.; Zhang, Y.; Chemosphere 2015, 125, 86.,44 Andaluri, G.; Manickavachagam, M.; Suri, R.; Environ. Monit. Assess. 2018, 190, 65. BPA is a well-known and studied endocrine disrupting compound (EDC), which can mimic hormones and cause reproductive damage, cancer and other adverse effects on the human body and the ecological environment.55 Huang, Y. Q.; Wong, C. K. C.; Zheng, J. S.; Bouwman, H.; Barra, R.; Wahlstrom, B.; Neretin, L.; Wong, M. H.; Environ. Int. 2012, 42, 91.,66 Arslan-Alaton, I.; Olmez-Hanci, T.; Dogan, M.; Ozturk, T.; Water Sci. Technol. 2017, 76, 2455. Previous studies77 Rykowska, I.; Wasiak, W.; Acta Chromatogr. 2006, 16, 7.,88 Wirasnita, R.; Hadibarata, T.; Yusoff, A. R. M.; Yusop, Z.; Water, Air, Soil Pollut. 2014, 225, 2148. have shown that BPA can interfere with the endocrine system of humans and animals even at concentrations below 1 μg m-3. Due to the discharge of domestic sewage and industrial wastewater, as well as the infiltration of landfill leachate, BPA has frequently been found in surface water and groundwater.99 Sharma, V. K.; Anquandah, G. A.; Yngard, R. A.; Kim, H.; Fekete, J.; Bouzek, K.; Ray, A. K.; Golovko, D.; J. Environ. Sci. Health, Part A: Toxic/Hazard. Subst. Environ. Eng. 2009, 44, 423.,1010 Careghini, A.; Mastorgio, A. F.; Saponaro, S.; Sezenna, E.; Environ. Sci. Pollut. 2015, 22, 5711. The study showed that the BPA concentration in some water environments can be as high as 100 μg L-1.1010 Careghini, A.; Mastorgio, A. F.; Saponaro, S.; Sezenna, E.; Environ. Sci. Pollut. 2015, 22, 5711. At present, the concentration of BPA in water environment and its harm to aquatic organisms have attracted the attention of researchers who are working on the research of refractory organics removal methods.

Since BPA is a refractory pollutant with two benzene rings, conventional biological processes in wastewater treatment plants are not efficient for complete removal of BPA.1111 Akbari, S.; Ghanbari, F.; Moradi, M.; Chem. Eng. J. 2019, 294, 298. There are several alternative approaches that have been used to deal with the BPA over the years, such as adsorption,1212 Dong, Y.; Wu, D.; Chen, X.; Lin, Y.; J. Colloid Interface Sci. 2010, 348, 585.,1313 Sui, Q.; Huang, J.; Liu, Y.; Chang, X.; Ji, G.; Deng, S.; Xie, T.; Yu, G.; J. Environ. Sci. 2011, 23, 177. chlorination with sodium hypochlorite,1414 Bourgin, M.; Bichon, E.; Antignac, J. P.; Monteau, F.; Leroy, G.; Barritaud, L.; Chachignon, M.; Ingrand, V.; Roche, P.; le Bizec, B.; Chemosphere 2013, 93, 2814. and the advanced oxidation processes (AOPs). The adsorption is restricted owing to the low hydrophobicity (logKow) of BPA that limits the adsorption efficiency,1515 Choi, K. J.; Kim, S. G.; Kim, C. W.; Park, J. K.; Korean J. Chem. Eng. 2006, 23, 399. and the chlorinated metabolites formed during chlorination with sodium hypochlorite cause some side effects.1616 Dupuis, A.; Migeot, V.; Cariot, A.; Albouy-Llaty, M.; Legube, B.; Rabouan, S.; Environ. Sci. Pollut. Res. Int. 2012, 19, 4193. In contrast, AOPs are popular due to their ability to decompose refractory organics into biodegradable and benign products by powerful free radicals such as hydroxyl radicals (•OH) or sulfate radicals (SO4•-).1717 Sharma, J.; Mishra, I. M.; Kumar, V.; J. Environ. Manage. 2016, 166, 12. In AOPs, ozone (O3),1818 Umar, M.; Roddick, F.; Fan, L.; Aziz, H. A.; Chemosphere 2013, 90, 2197. hydrogen peroxide (H2O2), peroxymonosulfate (PMS), and persulfate (PS) are commonly used oxidants, which can generate free radicals after being activated by certain ways.1111 Akbari, S.; Ghanbari, F.; Moradi, M.; Chem. Eng. J. 2019, 294, 298. For instance, in the UV/O3, UV/H2O2, and FeII/H2O2 processes, O3 and H2O2 can produce •OH due to UV or iron activation; in the UV/PS, Fe0/PS, FeII/PS, and Fe3O4/PS processes, PS can generate SO4•- also due to UV or iron activation.1111 Akbari, S.; Ghanbari, F.; Moradi, M.; Chem. Eng. J. 2019, 294, 298.,1818 Umar, M.; Roddick, F.; Fan, L.; Aziz, H. A.; Chemosphere 2013, 90, 2197. Although SO4•- and •OH have close oxidizing potentials (for SO4•-, E0 = 2.6-3.1 V; for •OH, E0 = 1.8-2.7 V), the half-life of SO4•- is longer than that of •OH, and the oxidant PS is more stable than O3 and H2O2. On the one hand, iron is a common material with benign property, which makes it as an activator to show economic and environmental advantages. On the other hand, from the standpoint of stability, ease of storage, and generation of sulfate radicals, PS is considered to be more suitable as an oxidant in AOPs. Therefore, the use of iron to activate PS to generate free radicals in degradation of refractory organics has become one of the research hotspots in AOPs. The mechanisms by which various forms of iron activate PS are attributed to the redox reactions between FeII and PS. Based on the existence form of iron, iron-based activator can be classified into homogeneous (such as FeII) and heterogeneous (such as Fe0 and Fe3O4).1919 Yang, D.; Zhou, X.; Liang, J.; Xu, Q.; Wang, H.; Yang, K.; Wang, B.; Wang, W.; J. Phys. D: Appl. Phys. 2021, 54, 244002. As a homogeneous activator, FeII is completely dissolved in the solution, which easily causes excessive FeII, thereby trapping the generated free radicals and reducing the degradation efficiency of organic matter.2020 Hong, Q.; Liu, C.; Wang, Z.; Li, R.; Liang, X.; Wang, Y.; Zhang, Y.; Song, Z.; Xiao, Z.; Cui, T.; Heng, B.; Xu, B.; Qi, F.; Ikhlaq, A.; Chem. Eng. J. 2021, 417, 129238.,2121 Moreira, F. C.; Rui, A.; Brillas, E.; Vilar, V. J. P.; Appl. Catal., B. 2015, 162, 34. On the contrary, as heterogeneous activates, Fe0 and Fe3O4 can release FeII into the water solution at a certain rate to avoid the extinction of free radicals caused by excessive FeII. Fe3O4 is a preferred activator with mixed valence oxide containing FeII and FeIII, and it is able to constantly produce FeII to react with PS and generate free radicals. In addition, Fe3O4 is the main component of magnetite in nature, so it is easy to obtain. Compared with FeII and Fe0, it is easy to recover Fe3O4 from wastewater based on its sub-magnetism. Fe3O4 can also reduce the cost of large volumes of waste effluent treatment and will not result in secondary pollution.2222 Jiang, S.; Zhu, J.; Wang, Z.; Ge, M.; Zhu, H.; Jiang, R.; Zong, E.; Guan, Y.; Ozone: Sci. Eng. 2018, 40, 457.

Some researchers used FeII and Fe0 to activate PS for the treatment of BPA,2222 Jiang, S.; Zhu, J.; Wang, Z.; Ge, M.; Zhu, H.; Jiang, R.; Zong, E.; Guan, Y.; Ozone: Sci. Eng. 2018, 40, 457.,2323 Gao, F.; Li, Y.; Xiang, B.; Ecotoxicol. Environ. Saf. 2018, 158, 239. and some researchers studied sulfamonomethoxine, ciprofloxacin hydrochloride degradation in Fe3O4/PS systems.2222 Jiang, S.; Zhu, J.; Wang, Z.; Ge, M.; Zhu, H.; Jiang, R.; Zong, E.; Guan, Y.; Ozone: Sci. Eng. 2018, 40, 457.,2424 Sepyani, F.; Soltani, R.; Jorfi, S.; Godini, H.; Safari, M.; J. Environ. Manage. 2018, 224, 315.,2525 Hu, Y.; Zhu, Q.; Yan, X.; Liao, C.; Jiang, G.; Environ. Res. 2019, 178, 108732. However, there are few studies towards BPA removal using Fe3O4/PS systems, especially the scavenging experiment of free radicals and the intermediate identification of BPA degradation products still need be further studied. In this paper, Fe3O4 magnetic particles have been used as activator to activate PS to produce free radicals for the removal of BPA. The performance of the Fe3O4/PS system on the degradation of BPA was systematically investigated, the possible effects of environmental factors on BPA removal, and BPA degradation mechanism were discussed in detail. Moreover, the role of active radical species was also explored using scavenging experiments, and the recycling performance of Fe3O4 was evaluated based on Fe3O4 reuse experiment.

Experimental

Chemicals and equipment

The main chemicals and equipment used in the experiments are present in Tables 1 and 2, respectively. The content of Fe3O4 in the iron oxide material is greater than 99.5%, the particle size of iron oxide is 20 nm, and the specific surface area of iron oxide is 51.46 m2 g-1.

Batch experiments

A 1000 mg L-1 of BPA aqueous stock solution was prepared for dilution into a series of concentrations in the batch experiments. The experiments were conducted in the 500-mL beakers placed on a magnetic heating stirrer with the operating speed of 200 rpm. Firstly, the BPA solution (500 mL) with the designed concentration was added to the beaker, then predetermined amount of PS and Fe3O4 was quickly plunge into the beaker to trigger the reaction. The reaction temperature was adjusted by the heater, and the pH was adjusted by adding 0.1 M NaOH or H2SO4 to the solution. The entire reaction time was set to 60 min. Aliquots (0.5 mL) were taken out of the beaker and put into the sample bottle. Then, 0.5 mL ethanol was immediately added into the sample bottle to stop the reaction. The samples were taken every 10 min from the beaker, and the residual concentration of BPA of the final samples were detected by high-performance liquid chromatography (HPLC). For the scavenging of free radicals generated in the reaction, the experiments were performed by adding methanol (MeOH) and tert-butanol (TBA) with preset concentrations in beakers.

Table 1
The main experimental chemicals
Table 2
The main equipment

Analytical methods

The BPA concentration in samples was examined by a HPLC equipped with a C18 column (4.6 mm × 150 mm, 4 μm) by using a fluorescence detector at 228 nm. The mobile phase was 70% methanol at a flow rate of 0.8 mL min-1. The excitation and emission wavelengths are 228 and 312 nm, respectively. The temperature of the column compartment was 25 ºC and the injection volume was 20 μL. The retention time of BPA under above conditions was 3.48 min.

The oxidation products of BPA were carried out by a gas chromatography-mass spectrometry (GC-MS). The chromatographic analysis was performed via injection in splitless mode (split ratio of 10:1, injection volume of 1 μL) at 260 ºC using a quartz capillary column (length: 30 m, inner diameter: 0.25 mm, film thickness: 0.25 μm) with a helium flow of 1.0 mL min-1. The qualitative analysis was performed with electron ionization (EI) at 70 eV using the full scan mode in the m/z range of 45-280. The ion source and quadrupole temperature were maintained at 230 and 150 ºC, respectively.

The removal rate η (%) of BPA is calculated using equation 1, where C0 is the initial concentration of BPA (mg L-1), Ct is the BPA concentration after t min treatment (mg L-1). The fitting of the kinetic equations in the experiment are based on the first-order reaction kinetic equation, as shown in equation 2, where k is the kinetic constant (min-1), b is the y intercept (no unit).

(1)η=C0-CtC0×100
(2)ln(CtC0)=-kt+b

Results and Discussion

The effect of the initial Fe3O4 dose

As the activator of PS, Fe3O4 plays a very important role in the reaction of Fe3O4/PS system. We investigated the effects of six Fe3O4 doses used in activating PS on removing BPA. The experimental results are presented in Figure 1. Without adding Fe3O4, the BPA removal rate can reach 19.68% by PS alone within 60 min. We speculate that this was due to temperature or light energy, which promote PS to produce SO4•- (equation 3), thereby removing BPA. A previous research2626 Liang, C.; Wang, Z. S.; Bruell, C. J.; Chemosphere 2007, 66, 106. confirmed that PS can be catalyzed by ambient temperature to form SO4•- to degrade organic matter. As adding Fe3O4 to the system, the removal rate of BPA was significantly increased. The FeII on the surface of Fe3O4 can react with PS to generate SO4•- (equation 4), and the BPA was removed. As the dose of Fe3O4 increased from 0.1 to 0.2 g L-1, the BPA removal rate increased from 41.94 to 56.49% within 60 min. When the dose of Fe3O4 was increased to 0.3-0.5 g L-1, the removal rate of BPA did not change much, compared with the BPA removal rate of 0.2 g L-1 of Fe3O4. We speculate that there are three possible reasons for this BPA degradation curve. Firstly, due to the increase of Fe3O4 dosage, more FeII can be produced, and excessive FeII will extinguish SO4•- (equation 5),2727 Yan, J.; Lei, M.; Zhu, L.; Anjum, M. N.; Zou, J.; Tang, H.; J. Hazard. Mater. 2011, 186, 1398. which will reduce the BPA removal rate. Secondly, the excessive SO4•- formed can react between themselves and reduce the number of SO4•- (equation 6).2828 Ouyang, D.; Yan, J.; Qian, L.; Chen, Y.; Han, L.; Su, A.; Zhang, W.; Ni, H.; Chen, M.; Chemosphere 2017, 184, 609. Finally, the increase in the amount of Fe3O4 (0.3-0.5 g L-1) causes the consumption of PS to increase. At 40 min, PS is exhausted and the number of free radicals decreases, so the degradation rate of BPA increases slowly.

(3)S2O82-energy2SO4-
(4)FeII+S2O82-FeIII+SO4-+SO42-
(5)FeII+SO4-FeIII+SO42-
(6)SO4-+SO4-S2O82-

Figure 1
The effect of the initial Fe3O4 concentration on BPA removal. [BPA]0 = 5 mg L-1, [PS]0 = 0.2 mM, [Fe3O4]0 = 0-0.5 g L-1, T0 = 20 ± 1 ºC, pH0 = 6.8 ± 0.2.

The effect of the initial PS concentration

The initial concentration of PS is an important factor affecting the BPA removal by Fe3O4 activation. Figure 2 gives the effect of the PS initial concentration on BPA degradation by Fe3O4 activation. In the absence of PS, it was found that 10.3% of BPA was removed within 60 min, indicating that Fe3O4 could adsorb BPA as an adsorbent. Sun et al.2929 Sun, C.; Zhou, R.; E, J.; Sun, J.; Su, Y.; Ren, H.; RSC Adv. 2016, 6, 10633. also found that nano-Fe3O4 can adsorb 2,4-dichlorophenol, but the efficiency was very low compared with Fe3O4/PS catalytic degradation. With the increase in the PS concentration from 0.1 to 0.3 mM, the BPA removal rate was elevated from 34.3 to 65.2% within 60 min, which could be due to the larger amount of SO4•- produced by PS/Fe3O4 system (equation 4). However, when the initial concentration of PS increased from 0.3 to 0.5 mM, the BPA removal rate increased only by 1.4% within 60 min. We have identified three possible reasons for this result. Firstly, excessive PS (0.4-0.5 mM) slowed down the degradation rate of BPA due to the elimination of SO4•- by PS (equation 7). Secondly, more SO4•- could also react with each other and lose their oxidation performance (equation 6). Finally, the increase in PS concentration causes the continuous consumption of Fe3O4, and the amount of FeII released from Fe3O4 cannot meet the demand for high concentrations of PS. Therefore, the relatively insufficient amount of catalyst results in a slower reaction rate of catalyzing PS to SO4•-, and ultimately affects the degradation efficiency of BPA.

(7)S2O82-+SO4-S2O8-+SO42-

Figure 2
The effect of the initial PS concentration on BPA removal. [BPA]0 = 5 mg L-1, [PS]0 = 0-0.5 mM, [Fe3O4]0 = 0.1 g L-1, T0 = 20 ± 1 ºC, pH0 = 6.8 ± 0.2.

The effect of the initial BPA concentration

The concentration of the substrate is also an important factor determining the degradation efficiency. It can be seen from the Figure 3 that when the initial BPA concentration increased from 1 to 5 mg L-1, after 60 min of reaction, the BPA removal rate dropped from 80.70 to 40.30%. When the initial concentration of BPA is 5 mg L-1, the kinetic constant for BPA removal is 0.00942 min-1 (ln(Ct/C0) = -0.00942t + 0.02804), and the reaction is the slowest. When the initial concentration of BPA was 1 mg L-1, the kinetic constant for BPA increased to 0.02466 min1 (ln(Ct/C0) = -0.02466t + 0.09686), which was 2.6 times faster than the degradation rate of 5 mg L-1 BPA. As the amount of PS and Fe3O4 remained unchanged, the number of SO4•- generated did not increase, and the increase of BPA concentration resulted to a relative shortage of SO4•-, which reduced the BPA removal rate. Therefore, in order to improve the degradation efficiency of BPA, the amount of oxidant and catalyst needs to be increased accordingly.

The effect of the initial pH

Figure 3
The effect of the initial BPA concentration on BPA removal. [BPA]0 = 1-5 mg L-1, [PS]0 = 0.2 mM, [Fe3O4]0 = 0.1 g L-1, T0 = 20 ± 1 ºC, pH0 = 6.8 ± 0.2.

It is clear that the advanced oxidation reaction in which PS is activated by transition metals is heavily dependent on solution pH. Therefore, we examined the BPA removal in Fe3O4/PS system at different pH of 3, 5, 7 and 9 (Figure 4). According to the BPA removal in descending order, the corresponding pH values were 5, 3, 7, 9, respectively. In other words, the pH value at the maximum BPA removal rate was 5, with the removal rate of 59.2%. Under acidic conditions of pH 3-5, the reaction proceed more efficiently due to the solubilization of the FeII on the surface of heterogeneous activator Fe3O4. Therefore, the homogeneous activation of FeII occurs more effectively for generating SO4•- towards removing BPA under acidic conditions. However, under neutral or alkaline conditions, FeII can hydrolyze to produce the precipitant Fe(OH)2 which forms a passivation layer on the surface of Fe3O4, which hinders the migration of FeII into water, and ultimately slows down the catalytic reaction. In addition, SO4•- can react with H2O and OH to produce •OH (equation 8) whose oxidizing power is less than SO4•-. These reasons resulted in the reduction of BPA oxidation performance under alkaline conditions. This finding was similar to the study of Liu et al.3030 Liu, Z.; Li, X.; Rao, Z.; Hu, F.; J. Environ. Manage. 2018, 208, 159. about treatment of landfill leachate biochemical effluent using Fe3O4/PS system.

(8)OH-+SO4-OH+SO42-

The effect of the reaction temperature

Temperature is a crucial parameter in the treatment of wastewater as it can affect the rate of chemical reactions. Some scholars have reported experiments that high temperature can activate PS.3131 Yang, S.; Wang, P.; Yang, X.; Shan, L.; Zhang, W.; Shao, X.; Niu, R.; J. Hazard. Mater. 2010, 179, 552.

32 Potakis, N.; Frontistis, Z.; Antonopoulou, M.; Konstantinou, I.; Mantzavinos, D.; J. Environ. Manage. 2017, 195, 125.
-3333 He, L.; Chen, H.; Wu, L.; Zhang, Z.; Ma, Y.; Zhu, J.; Liu, J.; Yan, X.; Li, H.; Yang, L.; Ecotoxicol. Environ. Saf. 2021, 208, 111522. We investigated the removal of BPA under four temperature conditions (20, 35, 50 and 70 ºC) in Fe3O4/PS and only PS system. The experimental results with or without Fe3O4 appearing in PS system at different temperatures are shown in Figure 5. In Figure 5, the curve marked with blank is the BPA removal without Fe3O4 (only PS). When only PS was added to the system, the removal rate of BPA increased significantly with the temperature increasing. The highest removal rate occurred at 70 ºC and the lowest removal rate occurred at 20 ºC. When the temperature was 70 ºC, the fastest kinetic constant for BPA removal was 0.00821 min-1 (ln(Ct/C0) = -0.00821t + 0.01496), which was 2 times faster than the 20 ºC rate of 0.00414 min-1 (ln(Ct/C0) = -0.00414t + 0.00818). Since heat can activate PS to produce SO4•- (equation 3), the increase in temperature was conducive to the degradation of BPA. When both PS and Fe3O4 were added to the system, the removal rate of BPA was significantly increased compared to the system with only PS. At 20 and 70 ºC, the removal rate of BPA was 40.46 and 87.10% after 60 min, respectively. This showed high temperature and Fe3O4 have the ability to synergistically catalyze the degradation of BPA by PS. However, the higher the temperature, the more energy is consumed. The actual situation of sewage treatment should be considered to determine the best temperature from the perspective of energy saving. In addition, as the temperature increased from 20 to 70 ºC, only adding Fe3O4 to the system, the BPA removal rate increased from 10.22 to 12.18% (data not shown). In other words, as the temperature increased, Fe3O4 alone cannot improve the efficiency of its adsorption of BPA.

Figure 4
The effect of the initial pH concentration on BPA removal. [BPA]0 = 5 mg L-1, [PS]0 = 0.2 mM, [Fe3O4]0 = 0.1 g L-1, T0 = 20 ± 1 ºC, pH0 = 3-9.
Figure 5
The effect of the reaction temperature on BPA removal. [BPA]0 = 5 mg L-1, [PS]0 = 0.2 mM, [Fe3O4]0 = 0 g L-1, 0.1 g L-1, T0 = 20-70 ºC, pH0 = 6.8 ± 0.2.

The scavenging of the free radicals

MeOH and TBA are commonly used chemical scavengers to distinguish SO4•- and •OH as they react with free radicals with different rate constants. MeOH is an effective scavenger for •OH (reaction rate constant, k = 9.7 × 108 M-1 s-1) and SO4•- (k = 1.1 × 107 M-1 s-1), while TBA is an effective scavenger for •OH (k = 6.0 × 108 M-1 s-1) and not for SO4•- (k = 9.1 × 105 M-1 s-1). As observed in Figure 6, without adding the MeOH or TBA, 48.5% of BPA degradation rate was acquired within 60 min. The MeOH and TBA of 80 mM added separately to the system resulted in BPA removal to drop to 25.9 and 45.4%, respectively. As the concentration of MeOH and TBA increased to 800 mM, the degradation rate of BPA further decreased to 12.8 and 39.9%. The addition of MeOH had a more obvious effect on the decrease of BPA degradation rate. Therefore, it was inferred that the SO4•- generated in the reaction was the main free radical, followed by the •OH.

Figure 6
The effect of MeOH and TBA on BPA removal. [BPA]0 = 5 mg L-1, [PS]0 = 0.2 mM, [Fe3O4]0 = 0.1 g L-1, T0 = 35 ± 1 ºC, pH0 = 6.8 ± 0.2.

BPA degradation products and pathways

GC-MS was used to qualitatively analyze the intermediate product of BPA degradation by nano-Fe3O4 activated PS. The reaction conditions are as follows: [BPA]0 = 5 mg L-1, [PS]0 = 0.2 mM, [Fe3O4]0 = 0.4 g L-1, T0 = 35 ± 1 ºC, pH0 = 5.0. The possible intermediate products of BPA degradation are shown in Table 3. The detected intermediate products were mainly aromatic compounds, including phenol, p-benzoquinone, p-hydroquinone, 4-(2-hydroxypropan-2-yl) phenol, 4-isopropenylphenol, p-hydroxybenzene propanoic acid and 4-hydroxybenzoic acid. The possible degradation pathways of BPA are shown in Figure 7. The free radicals generated from nano-Fe3O4/PS system attacked the C-C bond between isopropyl and benzene rings of bisphenol A, forming 4-(2-hydroxypropan-2-yl)phenol and phenol.3434 Du, J.; Bao, J.; Liu, Y.; Ling, H.; Zheng, H.; Kim, S. H.; Dionysiou, D. D.; J. Hazard. Mater. 2016, 320, 150. 4-(2-Hydroxypropan-2-yl) phenol was gradually oxidized into 4-isopropenylphenol, p-hydroxyphenylpropionic acid, and 4-hydroxybenzoic acid.3535 Torres-Palma, R. A.; Nieto, J. I.; Combet, E.; Pétrier, C.; Pulgarin, C.; Water Res. 2010, 44, 2245.,3636 Dong, Z.; Zhang, Q.; Chen, B. Y.; Hong, J.; Chem. Eng. J. 2019, 357, 337. Phenol was gradually oxidized into p-hydroquinone and p-benzoquinone. Then, all the aromatic compounds produced by two pathways were oxidized into small molecular compounds,1717 Sharma, J.; Mishra, I. M.; Kumar, V.; J. Environ. Manage. 2016, 166, 12.,3737 Li, X.; Wang, Z.; Zhang, B.; Rykov, A. I.; Ahmed, M. A.; Wang, J.; Appl. Catal., B 2016, 181, 788. some of which will be mineralized into CO2 and H2O under the conditions set in Fe3O4/PS system.

Figure 7
The possible degradation pathways of BPA.

Reusability of Fe3O4

In order to investigate the recyclability of Fe3O4, a magnet was used to separate Fe3O4 from the solution. After separation, the Fe3O4 was rinsed with deionized water, and dried indoors with natural ventilation for use in the next experiment. It can be seen from Figure 8 that the removal rate of BPA dropped from 68.0 to 26.9% after five cycles of Fe3O4. In the absence of Fe3O4, PS alone could remove about 20% of BPA, so it is believed that after five cycles of Fe3O4, the catalytic activity of Fe3O4 is already very low. It is speculated that the decrease in the activity of Fe3O4 is due to the fact that most of the FeII becomes FeIII that cannot effectively activate PS.3838 Leng, Y.; Guo, W.; Shi, X.; Li, Y.; Wang, A.; Hao, F.; Xing, L.; Chem. Eng. J. 2014, 240, 338.

Table 3
The possible intermediates of BPA degradation
Figure 8
Reuse of Fe3O4 and BPA removal rate. [BPA]0 = 5 mg L-1, [PS]0 = 0.2 mM, [Fe3O4]0 = 0.2 g L-1, T0 = 35 ± 1 ºC, pH0 = 5 ± 0.2.

Conclusions

Fe3O4/PS performed well as a sulfate radical-based AOP process in degradation of endocrine disruptor BPA. The removal rate of BPA was related to Fe3O4 dose, PS concentration, BPA concentration, pH and temperature. The best BPA removal rate of 80.7% could be acquired under the selected conditions of [BPA]0 = 1 mg L-1, [PS]0 = 0.2 mM, [Fe3O4]0 = 0.1 g L-1, T0 = 20 ± 1 ºC, pH0 = 6.8 ± 0.2. The oxidation of BPA by Fe3O4/PS mainly relies on sulfate radicals generated during the reaction, and hydroxyl radicals play a minor role. The possible intermediates of BPA degradation were determined as phenol, p-benzoquinone, p-hydroquinone, 4-(2-hydroxypropan-2-yl) phenol, 4-isopropenylphenol, p-hydroxybenzene propanoic acid and 4-hydroxybenzoic acid. The performance of Fe3O4 was significantly reduced after five times of reuse.

Acknowledgments

The authors wish to acknowledge the financial support from the National Natural Science Foundation of China (grants No. 42077160) and Jilin Province Education Department of China (grants No. JJKH20200288KJ).

References

  • 1
    Vandenberg, L. N.; Hauser, R.; Marcus, M.; Olea, N.; Welshons, W. V.; Reprod. Toxicol. 2007, 24, 139.
  • 2
    Pahigian, J. M.; Zuo, Y.; Chemosphere 2018, 207, 469.
  • 3
    Li, C.; Wang, Z.; Yang, Y. J.; Liu, J.; Mao, X.; Zhang, Y.; Chemosphere 2015, 125, 86.
  • 4
    Andaluri, G.; Manickavachagam, M.; Suri, R.; Environ. Monit. Assess 2018, 190, 65.
  • 5
    Huang, Y. Q.; Wong, C. K. C.; Zheng, J. S.; Bouwman, H.; Barra, R.; Wahlstrom, B.; Neretin, L.; Wong, M. H.; Environ. Int 2012, 42, 91.
  • 6
    Arslan-Alaton, I.; Olmez-Hanci, T.; Dogan, M.; Ozturk, T.; Water Sci. Technol 2017, 76, 2455.
  • 7
    Rykowska, I.; Wasiak, W.; Acta Chromatogr 2006, 16, 7.
  • 8
    Wirasnita, R.; Hadibarata, T.; Yusoff, A. R. M.; Yusop, Z.; Water, Air, Soil Pollut. 2014, 225, 2148.
  • 9
    Sharma, V. K.; Anquandah, G. A.; Yngard, R. A.; Kim, H.; Fekete, J.; Bouzek, K.; Ray, A. K.; Golovko, D.; J. Environ. Sci. Health, Part A: Toxic/Hazard. Subst. Environ. Eng. 2009, 44, 423.
  • 10
    Careghini, A.; Mastorgio, A. F.; Saponaro, S.; Sezenna, E.; Environ. Sci. Pollut 2015, 22, 5711.
  • 11
    Akbari, S.; Ghanbari, F.; Moradi, M.; Chem. Eng. J 2019, 294, 298.
  • 12
    Dong, Y.; Wu, D.; Chen, X.; Lin, Y.; J. Colloid Interface Sci 2010, 348, 585.
  • 13
    Sui, Q.; Huang, J.; Liu, Y.; Chang, X.; Ji, G.; Deng, S.; Xie, T.; Yu, G.; J. Environ. Sci. 2011, 23, 177.
  • 14
    Bourgin, M.; Bichon, E.; Antignac, J. P.; Monteau, F.; Leroy, G.; Barritaud, L.; Chachignon, M.; Ingrand, V.; Roche, P.; le Bizec, B.; Chemosphere 2013, 93, 2814.
  • 15
    Choi, K. J.; Kim, S. G.; Kim, C. W.; Park, J. K.; Korean J. Chem. Eng 2006, 23, 399.
  • 16
    Dupuis, A.; Migeot, V.; Cariot, A.; Albouy-Llaty, M.; Legube, B.; Rabouan, S.; Environ. Sci. Pollut. Res. Int 2012, 19, 4193.
  • 17
    Sharma, J.; Mishra, I. M.; Kumar, V.; J. Environ. Manage 2016, 166, 12.
  • 18
    Umar, M.; Roddick, F.; Fan, L.; Aziz, H. A.; Chemosphere 2013, 90, 2197.
  • 19
    Yang, D.; Zhou, X.; Liang, J.; Xu, Q.; Wang, H.; Yang, K.; Wang, B.; Wang, W.; J. Phys. D: Appl. Phys. 2021, 54, 244002.
  • 20
    Hong, Q.; Liu, C.; Wang, Z.; Li, R.; Liang, X.; Wang, Y.; Zhang, Y.; Song, Z.; Xiao, Z.; Cui, T.; Heng, B.; Xu, B.; Qi, F.; Ikhlaq, A.; Chem. Eng. J. 2021, 417, 129238.
  • 21
    Moreira, F. C.; Rui, A.; Brillas, E.; Vilar, V. J. P.; Appl. Catal., B. 2015, 162, 34.
  • 22
    Jiang, S.; Zhu, J.; Wang, Z.; Ge, M.; Zhu, H.; Jiang, R.; Zong, E.; Guan, Y.; Ozone: Sci. Eng. 2018, 40, 457.
  • 23
    Gao, F.; Li, Y.; Xiang, B.; Ecotoxicol. Environ. Saf. 2018, 158, 239.
  • 24
    Sepyani, F.; Soltani, R.; Jorfi, S.; Godini, H.; Safari, M.; J. Environ. Manage. 2018, 224, 315.
  • 25
    Hu, Y.; Zhu, Q.; Yan, X.; Liao, C.; Jiang, G.; Environ. Res 2019, 178, 108732.
  • 26
    Liang, C.; Wang, Z. S.; Bruell, C. J.; Chemosphere 2007, 66, 106.
  • 27
    Yan, J.; Lei, M.; Zhu, L.; Anjum, M. N.; Zou, J.; Tang, H.; J. Hazard. Mater. 2011, 186, 1398.
  • 28
    Ouyang, D.; Yan, J.; Qian, L.; Chen, Y.; Han, L.; Su, A.; Zhang, W.; Ni, H.; Chen, M.; Chemosphere 2017, 184, 609.
  • 29
    Sun, C.; Zhou, R.; E, J.; Sun, J.; Su, Y.; Ren, H.; RSC Adv 2016, 6, 10633.
  • 30
    Liu, Z.; Li, X.; Rao, Z.; Hu, F.; J. Environ. Manage 2018, 208, 159.
  • 31
    Yang, S.; Wang, P.; Yang, X.; Shan, L.; Zhang, W.; Shao, X.; Niu, R.; J. Hazard. Mater 2010, 179, 552.
  • 32
    Potakis, N.; Frontistis, Z.; Antonopoulou, M.; Konstantinou, I.; Mantzavinos, D.; J. Environ. Manage 2017, 195, 125.
  • 33
    He, L.; Chen, H.; Wu, L.; Zhang, Z.; Ma, Y.; Zhu, J.; Liu, J.; Yan, X.; Li, H.; Yang, L.; Ecotoxicol. Environ. Saf. 2021, 208, 111522.
  • 34
    Du, J.; Bao, J.; Liu, Y.; Ling, H.; Zheng, H.; Kim, S. H.; Dionysiou, D. D.; J. Hazard. Mater 2016, 320, 150.
  • 35
    Torres-Palma, R. A.; Nieto, J. I.; Combet, E.; Pétrier, C.; Pulgarin, C.; Water Res 2010, 44, 2245.
  • 36
    Dong, Z.; Zhang, Q.; Chen, B. Y.; Hong, J.; Chem. Eng. J. 2019, 357, 337.
  • 37
    Li, X.; Wang, Z.; Zhang, B.; Rykov, A. I.; Ahmed, M. A.; Wang, J.; Appl. Catal., B 2016, 181, 788.
  • 38
    Leng, Y.; Guo, W.; Shi, X.; Li, Y.; Wang, A.; Hao, F.; Xing, L.; Chem. Eng. J. 2014, 240, 338.

Publication Dates

  • Publication in this collection
    26 Nov 2021
  • Date of issue
    Dec 2021

History

  • Received
    05 May 2021
  • Accepted
    19 Aug 2021
Sociedade Brasileira de Química Instituto de Química - UNICAMP, Caixa Postal 6154, 13083-970 Campinas SP - Brazil, Tel./FAX.: +55 19 3521-3151 - São Paulo - SP - Brazil
E-mail: office@jbcs.sbq.org.br