SciELO - Scientific Electronic Library Online

 
vol.50 issue2Verification of runoff volume, peak discharge and sediment yield simulated using the ACRU model for bare fallow and sugarcane fieldsEffect of intercropping madumbe [Colocasia esculenta) and rice (Oryza sativa L.) on yield and land productivity under different irrigation water management techniques with effluent water author indexsubject indexarticles search
Home Pagealphabetic serial listing  

Services on Demand

Article

Indicators

Related links

  • On index processCited by Google
  • On index processSimilars in Google

Share


Water SA

On-line version ISSN 1816-7950
Print version ISSN 0378-4738

Water SA vol.50 n.2 Pretoria Apr. 2024

http://dx.doi.org/10.17159/wsa/2024.v50.i2.4078 

RESEARCH PAPER

 

Degradation of methyl orange by the Fenton-like reaction of pyrite-activated hydrogen peroxide forming the Fe(lll)/Fe(ll) cycle

 

 

Wenlong Bi; Ruojin Du; Hui Liu; Peng Fu; Zhenguo Li

College of Resources and Environment, Shanxi Agricultural University, Shanxi, 030801, PR China

Correspondence

 

 


ABSTRACT

In this study, the typical azo dye methyl orange (MO) was degraded by pyrite (FeS2) activated by hydrogen peroxide (H2O2). When [MO] = 0.1 mM, [FeSJ = 2.0 g/L and [H2O2] = 22 mM, 96.4% MO was removed in 120 min and the TOC removal rate was higher than 50%. HO* was the primary radical responsible for MO degradation, in addition, the acid condition promoted the degradation of MO in the FeS2/H2O2 system. MO in tap water and river water was not effectively degraded, whereas acidification could weaken the inhibitory effect on the FeS2/ H2O2 system to enable the degradation of MO in tap and river water. The OD600 indicated that the solution was environmentally friendly after the reaction, and three degradation pathways of MO were discussed. In summary, Fe(ll) could be dissolved from FeS2, which activated H2O2 to generate Fe(lll) and HO*. FeS2 could reduce Fe(lll) into Fe(ll), thus realizing the Fe(lll)/(ll) cycle and efficiently activating H2O2 to degrade MO.

Keywords: pyrite, hydrogen peroxide, methyl orange, advanced oxidation processes


 

 

INTRODUCTION

With the rapid development of printing and dyeing textile industries, the large amount of printing and dyeing wastewater produced has become one of the main water pollution sources in the world (Gu et al., 2021; Pinedo-Hernández et al., 2022). Printing and dyeing wastewater has a complicated composition, high concentration of pollutants, poor biochemical properties, and is difficult to treat (Ben Hafaiedh et al., 2020; Domingues et al., 2019; Shukla and Remya, 2021). Among the contaminants, methyl orange (MO) contains -N=N- and is a typical azo dye (Aghdasinia et al., 2016; Liu et al., 2022). Wastewater containing MO is discharged into the environment from industrial processes such as textiles and printing, and without effective treatment can have a serious impact on the environmental safety of surrounding water and soil, endangering human health (Giannakis et al., 2021). The accumulation of MO in water reduces the penetration of sunlight, inhibits the growth of aquatic organisms, and leads to an imbalance in the ecosystem (Song et al., 2022). Therefore, the study of efficient treatment methods for methyl orange wastewater has great significance.

Advanced oxidation processes (AOPs) are used to oxidize macromolecular refractory organics into environmentally friendly small molecules by generating free radicals with strong oxidizing properties through electricity, ultraviolet light, and ultrasound (Liu et al., 2022; Peng et al., 2019; Shao et al., 2020). AOPs, such as the Fenton reaction which involves reacting Fe2+ with H202 to produce hydroxy] radical (HO'), have a strong oxidation capacity, a fast reaction rate, high mineralization potential, and simple operation (Almeida et al., 2021; Gorozabel-Mendoza et al., 2021; Zhao et al., 2020). However the Fenton system has a high catalyst dosage and low pH after reaction (Jin et al., 2021). In addition, the high iron content of FeSO4 added to the Fenton system greatly increases the iron sludge treatment intensity and the disposal cost (Guo et al., 2020; Martins et al., 2017). Therefore, finding a low-cost and environmentally friendly catalyst to replace Fe(II) is critical for advanced oxidation processes.

To solve the iron-rich sludge problem in the Fenton system, a cheap and easily available iron-containing mineral needs to be found to replace ferrous ions. A promising candidate is pyrite (FeS2), which is the most abundant sulfide mineral on earth. FeS2 has been shown to be an efficient catalyst for the degradation of alachlor by Fenton's reagent (Kayan et al., 2020), and a nano-pyrite Fenton system was demonstrated to have improved the production of HO', resulting in the degradation of p-nitrophenol being three times faster than with the traditional Fenton system (Liu et al., 2020). Moreover, FeS2 is a low-cost, easily available, and environmentally friendly catalyst (Wang et al., 2021). The most noteworthy attribute of FeS2 is that the resource utilization of FeS2 can reduce the risk of FeS2 producing acidic mine wastewater (AMD) (Xia et al., 2023).

Therefore, in this study, FeS2 was used as a catalyst to explore the degradation and mechanism of MO in FeS2/H2O2 system. Furthermore, considering that printing and dyeing enterprises use a lot of tap water and river water, the degradation of MO in distilled water, tap water, and river water in FeS2/ H2O2 system was investigated.

 

MATERIALS AND METHODS

Chemicals

Methyl orange (MO, C14H14N3SO3Na, >98.8%) was purchased from Tianjin Tianli Chemical Reagent Co., Ltd. Hydrogen peroxide (H202, >30%) and concentrated sulfuric acid (H2SO4, 95.0%~98.0%) were purchased from Chengdu Kolon Chemical Reagent Co., Ltd. Ferrous sulfate (FeS04-7H2O, >99.0%), sodium hydroxide (NaOH, 96.0%), methanol (CH3OH, >99.5%), and tert-butyl alcohol (TBA, (CH3)3COH, >99.5%) were purchased from Tianjin Beichen Founder Reagent Factory. All the above reagents were analytically pure. All aqueous solutions were prepared with distilled water.

Pyrite pretreatment

Natural pyrite (FeS2) was purchased from Hebei China. Pyrite was ground and passed through a 0.15 mm mesh sieve, placed in a beaker and soaked in 0.5 M H2SO4 solution for 1 h to remove impurities and then filtered. The pyrite was washed repeatedly with distilled water until the pH of the filtrate was 7, and then filtered. The filtered sample was dried in an oven at about 40°C for future use.

Experimental method

MO removal by pyrite alone or hydrogen peroxide alone

FeS2 (0.2-2.0 g/L) or H202 (30-50 mM) was added to MO (0.1 mM) and the solution placed on a magnetic stirrer; 3 mL samples were taken at 5,10,20, 30, 60, 90, and 120 min and then to each sample 0.5 mL of MeOH was added, which can quench free radicals to ensure sample stability. The concentration of MO was calculated by colorimetric measurement at 470 nm (iso-absorption point wavelength, which is the wavelength at the intersection of the acidic and alkaline absorption curves) with a 721 visible spectrophotometer.

Optimization of pyrite and hydrogen peroxide

The addition of FeS2 and H2O2 was optimised using the response surface method in Design-Expert software. The design of experiment (DOE) is given in Table 1 which was designed using response surface methodology (RSM) (Ebru et al., 2023). The initial concentration of MO was 0.1 mmol/L, and the software-designed doses of FeS2 and H2O2 were added sequentially before the start of the reaction. 3 mL of the sample was removed from the filtrate at 15, 30, 45 and 60 min by passing it through a 0.45 µm membrane and adding 0.5 mL of methanol (free radical scavenger), and the concentrations were measured using a 721 visible spectrophotometer. The pH and temperature of the solution was measured at the end of the reaction. The data were visualised and analysed in Design-Expert software to obtain the optimum dose of FeS2 and H2O2.

 

 

Effect of different factors

In pH-effect experiments, the pH of the MO (0.1 mM) solution was adjusted to 2, 4, 6, 8 and 10 with H2S04 and NaOH, respectively and then FeS2 and H2O2 were added after the above optimization. The reaction was started and timed at 5, 10, 20, 30, 60, 90 and 120 min. 3 mL of the sample was removed from the filtrate after passing through a 0.45 µm filter membrane and 0.5 mL of methanol was added. The concentrations were measured using a 721 visible spectrophotometer, and the pH and temperature of the solution were measured at the end of the reaction. In the free-radical-contribution experiment, 2 mL TBA was added to MO solution, after adding optimized FeS2 and H202 into the MO solution; the remaining experimental steps were as above. All water used in the above procedures was distilled water. The influence experiment for different water matrices was conducted using distilled water, tap water, and river water (the Uma River in Taigu) to prepare MO solutions for degradation experiments, and the remaining experimental steps were performed as above.

Biotoxicity experiments

Biotoxicity experiments used microbial communities obtained by isolation and activation from campus soil (the soil was taken to cultivate the microorganisms used in the experiment in the area where the vegetation on the campus was lush and the microorganisms were active). The composition and content of the medium is given in Table 2. The 15 mL reaction solution was added to 10 mL of the medium and 0.5 mL mixed microbial community solution, and placed in an oscillator (rotating speed 180 r/min, temperature 25°C) for oscillating culture. Samples were taken at 6, 9, 12, 24 and 48 h, and the OD600 value reflecting bacterial density was determined by spectrophotometer. According to the difference between the reaction solution and the blank treatment, the microbial toxicity of the treatment solution was determined.

 

 

Analysis methods

Determination of total organic carbon concentration was conducted using a total organic carbon analyser (TOC; Muti N/C-3100 model). A scanning electron microscope (SEM; JSM-7001F) and X-ray diffractometer (XRD; MiniFlex II) were used to characterize FeS2. FeS2 surface elements were analysed using Al-Ka monochromatic X-ray photoelectron spectroscopy (XPS; AXIS ULTRA DLD). The intermediate products of MO degradation were analysed by a liquid chromatography-mass spectrometer (LC-MS; Agilent 1290 11-6470).

 

RESULTS AND DISCUSSION

Removal of MO by FeS2 alone and H2O2 alone

The removal effect for MO by different concentrations of FeS2 was examined as shown in Fig. 1a. When the FeS2 concentration was in the range of 0.2-2.0 g/L, the adsorption capacity of FeS2 for MO was weak and fluctuated up and down in the range of 0-3.5%. With an increase in FeS2 concentration, adsorption of MO did not improve effectively. Figure 1b shows the removal effect for MO by different concentrations of H2O2. The removal rates of MO at 30, 40 and 50 mmol/L H2O2 after 150 min were 2.8%, 5.0% and 4.9%, respectively. As the concentration of H2O2 increased, removal of MO did not increase significantly, indicating that the oxidation of MO by H2O2 was weak.

The degradation of MO in FeS2/H2O2 with different concentrations of H2O2

As can be seen from Fig. 2, when 0.1 mmol/L MO and [2.0 g/L FeS2 were added, the residual of MO at 30, 40 and 50 mmol/L H2O2 after 150 min of reaction was 97.4%, 93.5% and 96.1%, respectively. The degradation rate of MO was significantly increased after the simultaneous addition of FeS2 and H2O2. This may be due to the continuous dissolution of Fe(II) from FeS2 and the continuous activation of H2O2 by Fe(II) to produce IIO' (Eq. 1) to degrade MO (Oral and Kantar, 2019). Moreover, the sulfide in FeS2 could reduce Fe(III) to Fe(II) (Eq. 2) (Liu et al., 2015), further promoting the Fenton-like reaction. With the increase in H2O2 concentration, the difference between the removal rates for methyl orange was only 3.9%. Excluding the interference of error lines and combined with the desire for green and efficient' a relatively low concentration of H2O2 (30 mmol/L) was finally chosen for the subsequent optimization experiments.

 

 

 

Determination of the optimal concentration of FeS2 and H2O2

Response surface method optimization

By analysing the results of MO removal by FeS, alone, H,0, alone, and FeS,/H2O2 system (Figs 1 and 2), the concentration ranges of FeS2 and H2O2 were determined, and then 3D surface plots of MO removal rate after 150 min were fitted by Design-Expert software for the interaction between FeS2 and H2O2 concentrations (Abd Manan et al. 2019). As can be seen from Fig. 3, when the MO concentration was 0.1 mM, H2O2 concentration was 22 mM, and FeS2 concentration was 2.0 g/L, the MO removal rate was expected to reach 100% after 150 min.

 

 

The MO removal rate increased with the increase in FeS2 concentration. This was probably due to the degradation of MO with the increased FeS2 that could dissolve more Fe(II)-activated H2O2 generating HO- production.

When the MO concentration was 0.1 mM, FeS, concentration was 2.0 g/L, and with the increase in H2O2 concentration, the MO removal rate showed a trend of first increasing and then decreasing. The MO removal rate increased with the increase in H2O2 concentration in the range of 0-22 mM, which may be due to Fe(II) dissolved from FeS2, then reacting with H,0, to generate HO'. However, MO removal decreased once the H2O2 concentration was above 22 mM. On the one hand, H2O2 is a capture agent for HO', and excess H2O2 would not only react with Fc(II), but also with HO- to produce HO,' with low activity (Eq. 3) (Chang et al., 2020; Sun et al., 2019; Xiao et al., 2020), making the reaction rate decrease. On the other hand, excess HO' would react not only with the target substance but also with Fc(II), which made the degradation rate of MO decrease.

Verification of optimal conditions

To verify the above response surface methodology optimization results, the MO removal by different systems was plotted in Fig. 4a. MO removal by FeS2 or H2O2 alone was minimal. After 150 min of reaction, the MO removal by FeS2 alone was only 3.5%, and FeS2 had no obvious oxidative adsorption effect on MO.

The MO removal rate by H2O2 alone was only 4%, so the oxidation effect of H,0, on MO was weak.

When the MO concentration was 0.1 mM, H,0, concentration was 22 mM, and FeS, concentration was 2.0 g/L, the MO removal rate reached 96.3% at 120 min. The MO removal rate showed a trend of plateau, then a high rate, and finally a plateau again, which is consistent with a traditional Fenton-hke trend (Wang et al., 2013; Yan et al., 2017). The rapid degradation of MO at the stage of 30-90 min may be due to Fc(II) being dissolved from FeS2, which then reacted with H,0, to produce HO". FeS, also reduced Fe(III) to Fe(II), which continues to react with H2O2, resulting in rapid degradation of MO. The results for the UV scanning spectrum at different times also indicated rapid degradation of MO within 30-90 min (Fig. 4b). During the 90-150 min stage, the degradation of MO was slow. On the one hand, H2O2 was gradually consumed in FcS2/H2O2 system (Kokilavani et al., 2021), which made only a small amount of H2O2 react with Fe(II). On the other hand, intermediates of MO would compete with MO for IIO', slowing the MO degradation rate. To sum up, compared with FeS2 alone or H2O2 alone, the FeS2/H2O2 system could significantly remove MO. In addition, the degradation of MO tends to level off after 120 min. Because the concentration of MO at 150 min remained almost unchanged compared to that at 120 min, the final sampling time for subsequent experiments was 120 min.

Table 3 shows MO removal rates by other activated H2O2 methods. In Fc(II)/H2O2 system, although the MO is degraded in a short time, the amount of Fe(II) is large and the cost is high.

In UV/H2O2 system, an additional light source is needed and a large amount of H2O2 is required. Compared with other iron-hased/H2O2 systems, FeS2/H2O2 system could efficiently degrade 96.3% of MO with low doses of FeS, and H2O2. In addition, FeS2 is a mineral that exists in nature, is low cost and easy to obtain, and has good application value.

Contribution of different factors to FeS2/H2O2 system

Concentration of total Fe and Fe(ll)

Fe(II) plays a catalytic role in FeS,/H2O2 system, which could activate H2O2 to generate HO- and effectively degrade MO, and all iron ions in this system originate from FeS,. Therefore, we measured the total Fe and Fe(II) concentrations in FeS2/H2O2 system as shown in Fig. 5a. The total Fe concentration increased with time, but all of the concentrations recorded were lower than 3.5 mg/L, The Fe(II) concentration fluctuated from 0.33-0.64 mg/L. After 30 min of reaction, the total Fe and Fe(II) concentrations were 0.43 mg/L and 0.33 mg/L respectively, indicating that the slow degradation of MO in FeS2/H2O2 system before 30 min was due to the slow dissolution of Fe ions. During 30-120 min, the total Fe content increased significantly, reaching 3.12 mg/L at 120 min, while Fe(II) did not increase, reaching 0.64 mg/L at 120 min. This maybe due to the rapid reaction of dissolved Fe(II) with H2O2 to form Fe(III), FeS2 could reduce Fe(III) to Fe(II), and the reaction rate of Eq. 1 is faster than that of Eq. 2 (Diao and Chu 2021). Therefore, there was a large amount of Fe(III), and Fe(II)/Fe(III) in dynamic states, in the system, confirming that the degradation of MO was rapid after 30 min in FeS2/H2O2 system.

Validation of the interface effect of pyrite

H2O2 and Fe(II) were added to the MO solution to form Fe(II)/H2O2 homogeneous system to investigate the interface effect of FeS2. It can be seen from Fig. 5b that MO degraded slowly in Fe(II)/H2O2 system, and the degradation rate of MO was 77.4% after 120 min. H2O2 rapidly produced a large amount of HO' under the action of a high concentration catalyst in the early stage (Dulova et al., 2017). However, due to the limited reaction rate of HO' with organic matter, the free HO' not participating in the reaction aggregates, and free HO' reacted with itself to form H2O2 resulting in a part of the initially generated HO' being wasted (Liu et al., 2021). Furthermore, there were a large number of intermediate products of MO in the system, which might compete with MO for HO'. Therefore, the degradation trend of MO tended to level off. The pH after the reaction of the Fe(II)/H2O2 and FeS2/H2O2 systems was 4.00 and 3.25, respectively. FeS2 has the advantages of low cost and easy availability, so it has good application prospects.

Free radical contribution

To investigate the reactive oxygen species that played a dominant role in the removal of MO from FeS,/H2O2 system, a certain amount of free radical scavenger was added to the reaction system. The reaction rate constants of tert-butanol (TBA) and HO' are (3.8-7.6) x 10s M/s, so TBA is commonly used as a HO' scavenger (Chen et al. 2021, Xu and Sheng 2021). In addition, to effectively quench HO' within the system, we added TBA (0.4 M) at 4 000 times the dose of MO (0.1 mM). The effect of TBA on the degradation of MO in FeS2/H2O2 system is shown in Fig. 5c. The addition of TBA significantly inhibited the degradation of MO, and the MO removal rate was only 40.2% in 120 min. Therefore, HO' was the dominant reaction species in the system, and its contribution to the degradation of MO was about 58.3%. The interfacial effect of FeS, also contributed as the added FeS2 particle interface reacted with H2O2.

Influence of pH on the degradation of MO in FeS2/H2O2 system

The degradation of MO in FeS2/H2O2 system under different pH levels was shown in Fig. 6. In FeS2/H2O2 system, MO could be effectively removed under neutral and acidic conditions, and the removal effect was better under acidic conditions (Jin et al., 2021). However, it was difficult to effectively remove MO at pH - 10.

When the initial pH was 2 and 4, the MO degradation showed a fast and then slow trend. The degradation rates in 0-60 min were fast, and the MO removal rate could reach 94.8% and 87.5% in 60 min. The degradation tends to reach equilibrium within 60-120 min, with an increase of 2.4% and 8.7% in MO removal at 120 min compared with 60 min. Under acidic conditions, the dissolution of Fe(II) in FeS2 could be promoted (Li et al. 2019). Therefore, in the early stage of the reaction, the Fe(II) would

immediately activate H2O2, which made MO degrade rapidly. The slow degradation of MO after 60 min may be due to the decrease of H2O2, and only a small amount of Fe(II) reacted with a small amount of H2O2, which made MO degrade slowly. When the initial pH was 6 and 8, MO removals were 94.0% and 93.0%, respectively, after 120 min. The degradation trends were all flat, then fast and finally flat again, similar to the degradation trends at unadjusted pH (Dos Santos et al., 2021). The rapid degradation of MO in 30-90 min was not only due to the massive dissolution of Fe(II) to produce HO', but also because FeS, would produce H+ during the reduction of Fe(III), resulting in more dissolution of Fe(II) after the reduction of pH. Under the alkaline condition, the MO removal rate was poor at the initial pH of 10, and was only 2.2% after 120 min. It may be that Fe(II) dissolved slowly in FeS, under the alkaline condition (Dong et al., 2019).

When the initial pH was 10, the pH after the reaction was 9.71 However, with all of the other starting pH values, the pH after the reaction was less than 4 (Table 4). During the reduction of Fe(III), FeS2 produces H+, which lowers the pH of the solution. Under a traditional Fenton system, the best degradation efficiency was achieved at pH 2.79. At an initial pH of 8.0, there was almost no degradation of MO (Youssef et al., 2016). Compared with the traditional Fenton system, the FeS2/H2O2 system has a wider range of pH applications.

Influence of different water matrices on MO degradation in FeS2/H2O2 system

In practice, most printing and dyeing textile industries would directly use tap water or river water for treatment and, after a series of processes, the discharged wastewater has a complex composition. Therefore, we investigated MO degradation in tap water and river water by FeS2/H2O2 system to verify the possibility of FeS, treatment toward disposal of actual printing and dyeing wastewater.

As shown in Fig. 7a, MO was not effectively degraded in river water and tap water, and its degradation effect was much lower than that in distilled water. By determining the physical and chemical properties of river water and tap water (Table A1, Appendix) it was shown that they were weakly alkaline in pH and all had high TOC content. Consequently, it is speculated that the alkaline water would react with HO' to form low levels of active free radicals (Eqs 4-5) (Jiang et al., 2021; Ma et al., 2018), and the organic matter in river water and tap water would compete with MO for HO-. Hence there was no obvious removal effect for MO.

Under the acidic condition, the dissolution of Fe (II) will be accelerated and this will promote MO degradation. Moreover, the acidification could remove CO32-, HCO3-, Br-, NO3-, and Cl- in water. The distribution of carbonation binding states in Fig. A1 (Appendix) shows that at pH < 4 there is only free CO2 in the water. Therefore, the effect of different water matrices at pH = 2 was further investigated (Fig. 7b). After 60 min of reaction, the MO removal rates in distilled water, tap water, and river water were 97.2%, 96.9%, and 97.6%, respectively, achieving the expected treatment effect. Moreover, we prepared MO wastewater with similar ions to the water taken from the Uma River. It was found

that the simulated wastewater was difficult to degrade in FeS2/H2O2 system without adjusted pH, whereas the simulated wastewater had good degradation at a pH of 2 (Fig. 8), which further confirmed that acidification could weaken the inhibitory effect on the FeS2/H2O2 system for degrading MO in tap and river water.

Characterization of pyrite before and after reaction

The elemental composition of FeS, and the chemical states of Fe and S elements before and after the reaction were analysed by XPS. As can be seen from Fig. 9a, FeS, before and after the reaction was mainly composed of S, C, O, Fe, and Na. It could also be detected by EDS spectra (Fig. A2, Appendix). The C and O may be surface oxides, and Fe and S decreased in intensity after use, which may be as a result of being dissolved from the mineral into water. A corresponding change in the relative intensity of FeS, before and after the reaction could also be observed by XRD (Fig. 9b).

The high-resolution Fe-2p spectra of FeS, is shown in Fig. 9c. The peaks of 707.3, 709.6, 710.8, 711.5, and 714.2 eV of Fe 2p3/2 orbit were FeS2, FeO, Fe2O3, FeOOH, and satellite peaks, respectively. Except for the satellite peak, the peak area ratios before reaction were 40.2%, 10.7%, 23.6%, and 25.5%, respectively. The FeS, peak area decreased by 8.9%, while the FeOOH peak area increased by 8.7% after the reaction.

As can be seen from Fig. 9d, the peaks at 162.6 and 163.9 eV correspond to the S 2p3/2 and S 2p1;2 orbitais of S,2- (Ye et al, 2021) The peak area of S22 decreases from 77.7% to 54.6% before and after the reaction. The peaks at 168.5 and 169.7 eV are the S 2p3/2 and S 2p1/2 orbitais of SO42 , and the peak area of SO42 increased by 23.1% before and after the reaction. This indicates that S,2 in FeS, was oxidized to S042- during the oxidation process, which was consistent with the reaction given in Eq. 2 (Gao et al., 2021) Since the conversion of elemental S was much larger than that of elemental Fe, it was speculated that sulfide in FeS, could promote the conversion of Fe(III) to Fe(II) (Yu et al., 2020).

Determination of biotoxicity and total organic carbon of MO degraded by FeS2/H2O2 system

The growth of microorganisms was observed by determining the OD600 value to assess the biotoxicity of the FeS2/H2O2 system's degradation of MO. As can be seen from Fig. 10a, during 0-12 h, the number of microorganisms in the FeS2/H2O2 system was more than that in distilled water and MO solution. However, the biotoxicity of FeS2/H2O2 system at 48 h was not significantly different from that of MO, further suggesting that the FeS2, H2O2 and intermediate degradation products of MO were environmentally friendly. Thus, there is an excellent application prospect.

The total organic carbon (TOC) concentration in the FeS,/H2O2 system was measured to observe the degradation of MO (Suave et al., 2018). The TOC removal rate increased with increasing reaction time (Fig. 10b), indicating that the total carbon content in MO wastewater gradually decreased with time (Hamad et al., 2016). In FeS2/H2O2 system, the TOC removal rate reached 58.8% after 2 h of reaction. In summary, the FeS,/H2O2 system was capable of effectively degrading MO.

Possible degradation pathways and mechanism of MO in FeS2/H2O2 system

To determine the degradation path of MO, LC-MS was used to analyse the degradation intermediates of MO. Figure A3 (Appendix) shows the liquid-phase mass spectra of MO at 120 min; 8 intermediates of MO were inferred, and their molecular structures are listed in Table A2 (Appendix). The degradation pathways are shown in Fig. 11a. In FeS2/H2O2 system, Na of MO generally existed in the form of ion, and MO degradation paths were roughly divided into two: one was demethylation and deaminization for PI, P5, and P6 products; P7 was the -N=N-breakage product (Zhang et al., 2020), and P3 and P8 were deaminization products. Pathway 2 was that MO first broke the N=N- to P2, further -NH-NH- broke to P3, P4, and eventually became other small molecules. Figure 11b shows the degradation mechanism for MO in FeS,/H2O2 system. Fe(II) dissolved from FeS2, and Fe(II) activated H2O2 to generate HO", while sulfide in FeS2, could reduce Fe(III) to Fe(II) and continuously activate H2O2 to generate HO" for MO degradation.

 

CONCLUSION

In this study, H2O2 activated by FeS2 was effective in degrading MO and HO' was the main radical. In addition, the removal rate of TOC was higher than 50%. In addition, the reaction solution was environmentally friendly and three possible pathways for the degradation of MO were analysed. Compared with the traditional Fenton system, the FeS2/H2O2 system has a wider range of pH applications. Furthermore, MO in tap water and river water could not be effectively degraded in FeS2/H2O2 system. However, acidity could promote MO degradation; hence MO in tap water and river water at pH = 2 was rapidly degraded within 60 min. The degradation mechanism was that Fe(II) dissolved from FeS2, Fe(II) activated H2O2 to generate HO', which could degrade MO. Meanwhile, FeS2 could reduce Fe(III) to Fe(II). Therefore, the Fe(III)/Fe(II) cycle formed could continuously activate H2O2 to produce HO'to degrade MO.

 

AUTHOR CONTRIBUTIONS

The manuscript was written by Hui Liu. Ruojin Du contributed to the preparation and review of this manuscript. Experiments were performed by Peng Fu, and data analyses were carried out by Zhenguo Li. The project was supervised by Wenlong Bi. All authors have read and agreed to the published version of the manuscript.

 

ACKNOWLEDGEMENTS

This work was supported by the Youth Foundation of Shanxi Agricultural University Science and Technology Innovation Enhancement Project (CXGC2023026). Thanks to the staff of the College of Resources and Environment, Shanxi Agricultural University, and we are grateful to the anonymous reviewers and the editors for their valuable suggestions and comments on revising and improving this paper.

 

CONFLICT OF INTEREST

The authors declare no conflict of interest.

 

REFERENCES

ABD MANAN TSB, KHAN T, SIVAPALAN S, JUSOH H, SAPARI N, SARWONO A, RAMLI RM, HARIMURTI S, BEDDU S, SADON SN, KAMAL NLM and MALAKAHMAD A (2019) Application of response surface methodology for the optimization of polycyclic aromatic hydrocarbons degradation from potable water using photo-Fenton oxidation process. Sci. Total Environ. 665 196-212. https://doi.org/10.1016/j.scitotenv.2019.02.060        [ Links ]

AGHDASINIA H, BAGHERI R, VAHID B and KHATAEE A (2016) Central composite design optimization of pilot plant fluidized-bed heterogeneous Fenton process for degradation of an azo dye. Environ. Technol. 37 (21) 2703-2712. https://doi.org/10.1080/09593330.2016.1159734        [ Links ]

ALMEIDA LNB, JOSU TG, FIDELIS MZ, ABREU E, BECHLIN MA, DOS SANTOS OAA and LENZI GG (2021) Process comparison for caffeine degradation: Fenton, photo-Fenton, UV/H202 and UV/Fe3+. Water Aír Soil Pollut. 232 (4). https://doi.org/10.1007/s11270-021-05115-1        [ Links ]

GIANNAKIS S, LIN K Y A and GHANBARI F. (2021) A review of the recent advances on the treatment of industrial wastewaters by sulfate radical-based advanced oxidation processes(SR-AOPs). Chem. Eng. J. 406 127083. https://doi.org/10.1016/jxej.2020.127083        [ Links ]

BEN HAFAIEDH N, FOURCADE F, BELLAKHAL N and AMRANE A (2020) Iron oxide nanoparticles as heterogeneous electro-Fenton catalysts for the removal of AR18 azo dye. Environ. Technol. 41 (16) 2146-2153. https://doi.org/10.1080/09593330.2018.1557258        [ Links ]

CHANG Y, CHO Y-C and LIN Y-P (2020) Degradation of PFOS by a Mn02/H2O2 process. Environ. Sci. Water Res. Technol. 6 (12) 3476-3487. https://doi.org/10.1039/d0ew00739k        [ Links ]

CHEN Z, FU M, YUAN C, HU X, BAI J, PAN R, LU P and TANG M (2021) Study on the degradation of tetracycline in wastewater by micro-nano bubbles activated hydrogen peroxide. Environ. Technol 43 (23) 3580-3590. https://doi.org/10.1080/09593330.2021.1928292        [ Links ]

DIAO ZH and CHU W (2021) FeS2 assisted degradation of atrazine by bentonite-supported nZVI coupling with hydrogen peroxide process in water: Performance and mechanism. Sci. Total Environ. 754 142155. https://doi.org/10.1016/j.scitotenv.2020.142155        [ Links ]

DOMINGUES FS, FREITAS T, DE ALMEIDA CA, DE SOUZA RP, AMBROSIO E, PALACIO SM and GARCIA JC (2019) Hydrogen peroxide-assisted photocatalytic degradation of textile wastewater using titanium dioxide and zinc oxide. Environ. Technol. 40 (10) 1223-1232. https://doi.org/10.1080/09593330.2017.1418913        [ Links ]

DONG H, WANG B, LI L, WANG Y, NING Q, TIAN R, LI R, CHEN J and XIE Q (2019) Activation of persulfate and hydrogen peroxide by using sulfide-modified nanoscale zero-valent iron for oxidative degradation of sulfamethazine: A comparative study. Sep. Purif. Technol. 218 113-119. https://doi.Org/10.1016/j.seppur.2019.02.052        [ Links ]

DOS SANTOS NO, TEIXEIRA LA, ZHOU Q, BURKE G and CAMPOS L (2021) Fenton pre-oxidation of natural organic matter in drinking water treatment through the application of iron nails. Environ. Technol. 43(17)2590-2603.https://doi.org/10.1080/09593330.2021.1890838        [ Links ]

DULOVA N, KATTEL E and TRAPIDO M (2017) Degradation of naproxen by ferrous ion-activated hydrogen peroxide, persulfate and combined hydrogen peroxide/persulfate processes: The effect of citric acid addition. Chem. Eng. J. 318 254-263. https://doi.org/10.1016/j.cej.2016.07.006        [ Links ]

SHIYANG XIA, ZILING SONG, XIAOLIANG ZHAO and JIYANG LI (2023) Review of the recent advances in the prevention, treatment, and resource recovery of acid mine wastewater discharged in coal mines.J. Water Process Eng. 52 2214-7144. https://doi.org/10.1016/j.jwpe.2023.103555        [ Links ]

EBRU T, EMMANUEL LKECHUKWU U, EY P S and ZEYNI A (2023) Application of biopolymer in turbidity removal and sludge settling behaviour of travertine-processing wastewater: Performance optimization using response surface methodology (RSM). Water SA 49 (1 January), https://doi.org/10.17159/wsa/2023.v49.il.3952        [ Links ]

GAO Y, LI Z, FU Z, ZHANG H, WANG G and ZHOU H (2021) Highly selective capacitive deionization of copper ions in FeS2@N, S co-doped carbon electrode from wastewater. Sep. Purif. Technol. 262 118336. https://doi.org/10.1016/j.seppur.2021.118336        [ Links ]

GOROZABEL-MENDOZA ML, FILHO OAE, ZAMBRANO-INTRIAGO LA, BAQUERIZO-CRESPO RJ, GILER-MOLINA JM and RODRIGUEZ-D AZ JM (2021) Degradation of Blue 1 and Yellow 6 Dyes in binary mixture using photo-Fenton/sunlight system: optimization by factorial designs. Water Aír Soil Pollut. 232 (12). https://doi.org/10.1007/s11270-021-05461-0        [ Links ]

GU C, HU J, ZHANG M, DING J, GONG T, WANG Z, ZHU J and GAN M (2021) Development of a hydroxyl group-mediated biosynthetic schwertmannite as a persulfate activator for efficient degradation of RhB and Cr(VI) removal, J. Hazardous Mater. 419 126496. https://doi.Org/10.1016/j.jhazmat.2021.126496        [ Links ]

GUO S, WANG Q, LUO C, YAO J, QIU Z and LI Q (2020) Hydroxyl radical-based and sulfate radical-based photocatalytic advanced oxidation processes for treatment of refractory organic matter in semi-aerobic aged refuse biofilter effluent arising from treating landfill leachate. Chemosphere 243 125390. https://doi.org/10.1016/j.chemosphere.2019.125390        [ Links ]

HAJI S, BENSTAALI B and AL-BASTAKI N (2011) Degradation of methyl orange by UV/H2O2 advanced oxidation process. Chem. Eng. J. 168 (1) 134-139. https://doi.org/10.1016/jxej.2010.12.050        [ Links ]

HAJI S, KHALAF M, SHUKRALLAH M, ABDULLAH J and AHMED S (2014) A kinetic comparative study of azo dye decolorization by catalytic wet peroxide oxidation using Fe-Y zeolite/H2O2 and photooxidation using UV/H2O2. Reaction Kinetics, Mech. Catal. 114 (2) 795-815. https://doi.org/10.1007/slll44-014-0810-3        [ Links ]

HAMAD D, DHIB R and MEHRVAR M (2016) Effects of hydrogen peroxide feeding strategies on the photochemical degradation of polyvinyl alcohol. Environ. Technol. 37 (21) 2731-2742. https://doi.org/10.1080/09593330.2016.1160959        [ Links ]

HANG Y, SUN D, WANG D, TONG L, ZHANG Z, CAO N and GONG Z (2021) Rapid degradation of tetracycline in aqueous solution by Fe/Cu catalysis enhanced by H2O2 activation. Environ. Technol. 43 (24) 1-9. https://doi.org/10.1080/09593330.2021.1933610        [ Links ]

TIN X, SU J and YANG Q (2021) A comparison study of Fenton-like and Fenton reactions in dichloromethane removal. Environ. Technol. 9: 1-11. https://doi.org/10.1080/09593330.2021.1954096        [ Links ]

KAYAN I, OZ NA and KANTAR C (2020) Coupling pyrite-Fenton process with aerobic biodégradation for the treatment of 2-chlorophenol. Water Aír Soil Pollut. 231 (9). https://doi.org/10.1007/S11270-020-04836-Z        [ Links ]

KOKILAVANI S, SYED A, RAJU LL, AL-RASHED S, ELGORBAN AM, THOMAS AM and KHAN SS (2021) Synthesis of novel heterostructured FeS2/Ag2Mo04 nanocomposite: Characterization, efficient antibacterial and enhanced visible light driven photocatalytic activity. Surf. Interf. 23 101003. https://doi.org/10.1016/j.surfin.2021.101003        [ Links ]

LI L, LAI C, HUANG F, CHENG M, ZENG G, HUANG D, LI B, LIU S, ZHANG M, QIN L, LI M, HE J, ZHANG Y and CHEN L (2019) Degradation of naphthalene with magnetic bio-char activate hydrogen peroxide: Synergism of bio-char and Fe-Mn binary oxides. Water Res. 160 238-248. https://doi.Org/10.1016/j.watres.2019.05.081        [ Links ]

LIU H, FU P, LIU F, HOU Q, TONG Z and BI W (2022) Degradation of ciprofloxacin by persulfate activated with pyrite: mechanism, acidification and tailwater reuse. RSC Adv. 12 (46) 29991-30000. https://doi.org/10.1039/d2ra05412d        [ Links ]

LIU H, LIU F, ZHANG J, ZHOU J, BI W, QIN J, HOU Q, NI Y, XU S and YANG C (2022) Degradation of methyl orange by pyrite activated persulfate oxidation: mechanism, pathway and influences of water substrates. Water Sci. Technol. 85 (10) 2912-2927. https://doi.org/10.2166/wst.2022.134        [ Links ]

LIU T, CHEN N, DENG Y, CHEN F and FENG C (2020) Degradation of p-nitrophenol by nano-pyrite catalyzed Fenton reaction with enhanced peroxide utilization. RSC Adv. 10 (27) 15901-15912. https://doi.org/10.1039/d0ra01177k        [ Links ]

LIU W, WANG Y, AI Z and ZHANG L (2015) Hydrothermal synthesis of FeS2 as a high-efficiency fenton reagent to degrade alachlor via superoxide-mediated Fe(II)/Fe(III) Cycle. ACS Appl. Mater. Interfaces 7 (51) 28534-28544. https://doi.org/10.1021/acsami.5b09919        [ Links ]

LIU Y, ZHAO Y and WANG J (2021) Fenton/Fenton-like processes with in-situ production of hydrogen peroxide/hydroxyl radical for degradation of emerging contaminants: Advances and prospects. J. Hazardous Mater. 404 (Pt B) 124191. https://doi.org/10.1016/j.jhazmat.2020.124191        [ Links ]

MA X CHENG Y, GE Y, WU H, LI Q, GAO N and DENG J (2018) Ultrasound-enhanced nanosized zero-valent copper activation of hydrogen peroxide for the degradation of norfloxacin. Ultrasonics Sonochem. 40 (Pt A) 763-772. https://doi.Org/10.1016/j.ultsonch.2017.08.025        [ Links ]

MARTINS A, TEIXEIRA LAC, DA FONSECA FV and YOKOYAMA L (2017) Evaluation of the mercaptobenzothiazole degradation by combined adsorption process and Fenton reaction using iron mining residue. Environ. Technol. 38 (16) 2032-2039. https://doi.org/10.1080/09593330.2016.1244571        [ Links ]

ORAL O and KANTAR C (2019) Diclofenac removal by pyrite-Fenton process: Performance in batch and fixed-bed continuous flow systems. Sci. Total Environ. 664 817-823. https://doi.org/10.1016/j.scitotenv.2019.02.084        [ Links ]

PENG J, ZHANG Y, ZHANG C, MIAO D, LI J, LIU H, WANG L and GAO S (2019) Removal of triclosan in a Fenton-like system mediated by graphene oxide: Reaction kinetics and ecotoxicity evaluation. Sci. Total Environ. 673 726-733. https://doi.org/10.1016/j.scitotenv.2019.03.354        [ Links ]

PINEDO-HERNANDEZ S, SÁNCHEZ-MENDIETA V, SOLACHE-RlOS M and GUTIRREZ-SEGURA E (2022) Degradation of brilliant blue by heterogeneous Fenton and UV-fenton processes. Water Air Soil Pollut. 233 (7). https://doi.org/10.1007/s11270-022-05722-6        [ Links ]

SHAO P, YU S, DUAN X, YANG L, SHI H, DING L, TIAN J, YANG L, LUO X and WANG S (2020) Potential difference driving electron transfer via defective carbon nanotubes toward selective oxidation of organic micropollutants. Environ. Sci. Technol. 54 (13) 84648472. https://doi.org/10.1021/acs.est.0c02645        [ Links ]

SHUKLA N and REMYA N (2021) Microwave photo-oxidation with diverse oxidants for Congo red degradation: effect of oxidants, degradation pathway and economic analysis. Environ. Technol. 42 (10) 1482-1492. https://doi.org/10.1080/09593330.2019.1670737        [ Links ]

SONG W, LI J, ZHANG X, FENG J, DU X, WANG Q, FU C, QIU W, WANG Z and GAO X (2022) A feasible approach for azo-dye methyl orange degradation in siderite/H2O2 assisted by persulfate: Optimization using response surface methodology and pathway, J. Environ. Manage. 308 114397. https://doi.org/10.1016/j.jenvman.2021.114397        [ Links ]

SUAVE J, JOSE HJ and MOREIRA R (2018) Photocatalytic degradation of polyvinylpyrrolidone in aqueous solution using Ti02/H2O2/UV system. Environ. Technol. 39 (11) 1404-1412. https://doi.org/10.1080/09593330.2017.1330365        [ Links ]

SUN Y, CHO DW, GRAHAM NJD, HOU D, YIP ACK, KHAN E, SONG H, LI Y and TSANG DCW (2019) Degradation of antibiotics by modified vacuum-UV based processes: Mechanistic consequences of H2O2 and K2S208 in the presence of halide ions. Sci. Total Environ. 664 312-321. https://doi.org/10.1016/j.scitotenv.2019.02.006        [ Links ]

TANG J and WANG J (2017) Fe-based metal organic framework/ graphene oxide composite as an efficient catalyst for Fenton-like degradation of methyl orange. RSC Adv. 7 (80) 50829-50837. https://doi.org/10.1039/c7ral0145g        [ Links ]

WANG C, SUN R, HUANG R and CAO Y (2021) A novel strategy for enhancing heterogeneous Fenton degradation of dye wastewater using natural pyrite: Kinetics and mechanism. Chemosphere 272 129883. https://doi.org/10.1016/jxhemosphere.2021.129883        [ Links ]

WANG WM, SONG J and HAN X (2013) Schwertmannite as a new Fenton-like catalyst in the oxidation of phenol by H202. J. Hazardous Mater. 262 412-419. https://doi.org/10.1016/j.jhazmat.2013.08.076        [ Links ]

XIAO R, HE L, LUO Z, SPINNEY R, WEI Z, DIONYSIOU DD and ZHAO F (2020) An experimental and theoretical study on the degradation of clonidine by hydroxyl and sulfate radicals. Sci. Total Environ. 710 136333. https://doi.Org/10.1016/j.scitotenv.2019.136333        [ Links ]

XU H and SHENG Y (2021) New insights into the degradation of chloramphenicol and fluoroquinolone antibiotics by peroxymonosulfate activated with FeS: Performance and mechanism. Chem. Eng. J. 414 128823. https://doi.org/10.1016/jxej.2021.128823        [ Links ]

YAN J, QIAN L, GAO W, CHEN Y, OUYANG D and CHEN M (2017) Enhanced Fenton-like degradation of trichloroethylene by hydrogen peroxide activated with nanoscale zero valent iron loaded on biochar. Sci. Rep. 7 43051. https://doi.org/10.1038/srep43051        [ Links ]

YE J, ZANG Y, WANG Q, ZHANG Y, SUN D, ZHANG L, WANG G, ZHENG X and ZHU J (2021) Nitrogen doped FeS2 nanoparticles for efficient and stable hydrogen evolution reaction, J. Energ. Chem. 56 283-289. https://doi.Org/10.1016/j.jechem.2020.08.014        [ Links ]

YOUSSEF NA, SHABAN SA, IBRAHIM FA and MAHMOUD AS (2016) Degradation of methyl orange using Fenton catalytic reaction. Egyptian J. Petrol. 25 (3) 317-321. https://doi.org/10.1016/j.ejpe.2015.07.017        [ Links ]

YU F, WANG Y, MA H and ZHOU M (2020) Hydrothermal synthesis of FeS2 as a highly efficient heterogeneous electro-Fenton catalyst to degrade diclofenac via molecular oxygen effects for Fe(II)/Fe(III) cycle. Sep. Purif. Technol. 248 117022. https://doi.org/10.1016/j.seppur.2020.117022        [ Links ]

ZHANG X, QIN Y, ZHANG W, ZHANG Y and YUAN GE (2020) Oxidative degradation of Orange G in aqueous solution by persulfate activated with pyrite. Water Sci. Technol. 82 (1) 185-193. https://doi.org/10.2166/wst.2020.352        [ Links ]

ZHAO Z, PAN S, YE Y, ZHANG X and PAN B (2020) FeS2/H2O2 mediated water decontamination from p-arsanilic acid via coupling oxidation, adsorption and coagulation: Performance and mechanism. Chem. Eng. J. 381 122667. https://doi.org/10.1016/j.cej.2019.122667        [ Links ]

ZHIYONG Y, WENHUA W, LIN S, LIQIN L, ZHIYIN W, XUANFENG J, CHAONAN D and RUIYING Q (2013) Acceleration comparison between Fe2+/H2O2 and Co2+/oxone for decolouration of azo dyes in homogeneous systems. Chem. Eng. J. 234 475-483. https://doi.org/10.1016/jxej.2013.08.013        [ Links ]

 

 

Correspondence:
Wenlong Bi
Email: jiayangbihu@126.com

Received: 14 May 2023
Accepted: 22 March 2024

 

 

APPENDIX

 

 

 


Table A2 - Click to enlarge

 

 

 

 

 

 


Figure A3 - Click to enlarge

Creative Commons License All the contents of this journal, except where otherwise noted, is licensed under a Creative Commons Attribution License