Next Article in Journal
Selective Oligomerization of Isobutylene in Mixed C4 with Co/BETA-Loaded Molecular Sieve Catalysts
Previous Article in Journal
The Advancement of Supported Bimetallic Catalysts for the Elimination of Chlorinated Volatile Organic Compounds
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Application of Different Waveforms of Pulsed Current in the Classical Electro-Cocatalytic Process for Effective Removal of Sulfamethoxazole: Oxidation Mechanisms

1
Sichuan Provincial Key Laboratory of Universities on Environmental Science and Engineering, MOE Key Laboratory of Deep Earth Science and Engineering, Department of Environmental Science and Engineering, College of Architecture and Environment, Sichuan University, Chengdu 610065, China
2
Shanghai Key Laboratory of Molecular Catalysis and Innovation Materials, Collaborative Innovation Centre of Chemistry for Energy Materials, MOE Laboratory for Computational Physical Science, Shanghai Key Laboratory of Bioactive Small Molecules, Department of Chemistry, Fudan University, Shanghai 200433, China
*
Author to whom correspondence should be addressed.
Catalysts 2024, 14(8), 532; https://doi.org/10.3390/catal14080532
Submission received: 16 July 2024 / Revised: 9 August 2024 / Accepted: 12 August 2024 / Published: 16 August 2024
(This article belongs to the Section Industrial Catalysis)

Abstract

:
In this study, sulfamethoxazole (SMX) was applied as the model pollutant to assess the performance of pulsed current (PC) waveforms in the decontamination efficiency of the PC/peroxymonosulfate (PMS)/Fe(III) process and to investigate underlying oxidation mechanisms. Among the various waveforms tested, the sinusoidal wave (SIN), combined with the Dimensionally Stable Anode (DSA) electrode, demonstrated superior degradation performance, with the order being SIN > ramp > square > direct current (DC). The operational conditions for the SIN/PMS/Fe(III) system were optimized to an initial pH of 3, a voltage of 6 V, 0.6 mmol/L of Fe3+, 1.0 mmol/L of PMS, and a frequency of 1 kHz. The results of quenching and probe experiments confirmed the generation of abundant reactive radicals such as OH, SO4•−, O2•−, Fe(IV), and 1O2 in the SIN/PMS/Fe(III) process, which collectively enhanced the degradation of SMX. Additionally, results of high-resolution mass spectrometry analysis were employed to identify the SMX oxidation byproducts, and the toxicity of SMX byproducts was evaluated. Overall, the SIN/PMS/Fe(III) process exhibits effective degradation capacity with high energy efficiency, establishing itself as an effective strategy for the practical treatment of medical wastewater.

1. Introduction

As a broad-spectrum antibiotic, sulfamethoxazole (SMX) plays a vital role in the medical therapy of both human and livestock illness, known for commonly being employed in the treatment of urinary tract infections and intestinal infections caused by Escherichia coli and Proteus [1], thus finding widespread application in medical settings. However, SMX has garnered attention due to its pervasive detection and high concentrations in various environmental matrices, including hospital wastewater, municipal wastewater, and groundwater, across different regions worldwide [2]. The discharge of SMX towards the environment causes a direct threat to non-target organisms and significantly amplifies the danger of disseminating drug-resistant bacteria and resistance genes [3], thereby adversely affecting human and environmental health. Consequently, wastewater from hospitals and pharmaceutical facilities containing SMX represents a formidable challenge to environmental ecology and human well-being. Addressing this challenge necessitates urgent scientific research aimed at developing innovative treatment methods to mitigate the environmental impact of SMX contamination.
Over the past few decades, researchers have been dedicated to developing technologies capable of efficiently removing resistant organic contamination from industrial wastewater. Among these technologies, advanced oxidation processes (AOPs) are expected as the most promising [4,5,6]. Typically, AOPs generate short-lived free radicals (OH) with powerful oxidizing capabilities under specific conditions (such as light, heat, electricity, or the presence of catalysts) for the oxidation of large, recalcitrant organic molecules into low-toxicity or non-toxic smaller molecules. Electrochemical Advanced Oxidation Processes (EAOPs), an important subset of AOPs, have gained significant reputation recently due to their environmental friendliness and high efficiency [7,8,9,10]. They have been diffusely utilized in the disposal of various organic pollutant-containing wastewaters. In EAOPs, the electro-Fenton technology leverages the interaction between H2O2 and Fe2+ to produce reactive radicals, which are employed for the removal of contaminations [11,12]. The choice of electrodes in the electro-Fenton method is crucial for the efficiency of organic pollutant oxidation. Dimensionally Stable Anodes (DSAs), such as Ti/RuO2–IrO2, are cost-effective; they involve depositing a subtle layer of metal oxide on a base metal (usually titanium), which reacts synergistically with catalysts or anions in water under the influence of an electric current. Graphite electrodes, used as cathodes, are low-cost, exhibit good electrical conductivity, and have demonstrated excellent properties in previous studies [13,14].
In recent years, pulse electrochemistry (PE) has garnered significant attention as an emerging electrocatalytic method, with many scientists seeking its advantages in terms of lower energy consumption and higher degradation efficiency [15,16]. Current research has demonstrated that PE systems have been used in conjunction with catalysts to determine inorganic metal pollutant concentrations [17], detect biological protein levels in biomedicine [18], and drive the periodic regeneration of catalysts [19]. Notably, studies on the degradation of dye wastewater or antibiotics wastewater using PE have found that the •OH radicals generated in pulse systems exceed those in direct current (DC) systems [20,21,22]. Although the mechanism of PE system degradation has not been thoroughly explained, experimental conclusions have provided a broad scope for further research. The waveform of the pulse current has been scarcely studied; thus, considering the impact of the current waveform on the reaction rate and degradation efficiency, we aim to identify the optimal current waveform and evaluate its advantages compared to DC and other pulse current waveforms, and to identify the reasons for its high efficiency and energy performance. Based on a literature report, the primary advantage of pulsed current lies in maintaining high degradation efficiency in the electro-Fenton process while simultaneously reducing energy consumption [20,23]. Moreover, pulsed current systems have been observed to generate new oxidative degradation pathways, potentially resulting in less toxic or non-toxic byproducts, thus being more environmentally friendly [24]. Under the unique parameter control of pulsed current, it is possible to achieve a degradation efficiency and energy utilization efficiency that direct current cannot attain.
Peroxymonosulfate (PMS), commonly used as an oxidant in the electro-Fenton method [25], is rapidly recognized as a selective, effective, and green option for aqua-purification [26]. The electrochemical activation of PMS can enhance PMS activation and decontamination efficiency through both radical oxidation (e.g., OH) and non-radical oxidation pathways [27]. Compared to other ions that can catalyze PMS, such as Co2+, Fe3+ has advantages including non-toxicity, availability, and low cost [28]. Additionally, the internal recycling reaction of Fe2+/Fe3+ within the system accelerates the Fenton reaction, resulting in a higher degradation efficiency [29].
Herein, this experiment focused on the following points: (1) investigating the removal efficiency and reaction rate of SMX, as the primary pollutant, and employing the SIN/PMS/Fe(III) system and comparing it with direct current (DC) and other waveform pulse currents; (2) determining the ideal experimental parameters for SMX removal within the SIN/PMS/Fe(III) process by varying factors such as pH, voltage, and Fe3+ concentration; (3) performing quenching and probe experiments to identify and quantify the active radicals produced within the SIN/PMS/Fe(III) process, and comparing these results with those from the DC system to clarify the SIN/PMS/Fe(III) system’s advantages; and (4) identifying the byproducts during degrading SMX employing the SIN/PMS/Fe(III) process, assessing their toxicity, and evaluating the practical applicability of the SIN/PMS/Fe(III) process within various water matrices under the influence of different coexisting ions.

2. Results and Discussion

2.1. Impact of System Parameters on Degradation Performance

In the preliminary investigation of the impacts of various current waveforms on the removal rate of SMX, four types of currents were selected: sinusoidal, square, ramp, and direct current. As depicted in Figure 1a,b, the removal efficiency is as follows: sinusoidal > square > ramp > square > direct current. The characterization reveals that the degradation efficacy of various alternating currents surpasses that of direct current. This can be attributed to the continuous conversion of anode–cathode polarity with the current, akin to the practical effect of the pulse duty cycle in experiments with square and ramp waveforms. The duty cycle denotes the percentage of time during the whole cycle that the pulse is applied. Generally, a higher duty cycle extends the discharge duration and enhances the production of active free radicals through the electrode surface, thereby accelerating the efficiency in degrading the pollutant, whereas extremely high duty cycles can induce concentration polarization, which diminishes the material transfer efficiency and subsequently hinders the degradation efficiency [30]. Consequently, a 50% duty cycle was chosen as the ideal condition to achieve both effective degradation and energy efficiency, and this condition was employed in subsequent experiments.
Sinusoidal current continuously reverses the polarity of the electrode, causing ions between the electrodes to undergo relative oscillations under the electric field in a very short period before reaching the electrode, thereby minimizing the likelihood of side reactions such as water electrolysis [20,23]. This minimizes the electrolysis of water reaction and allows more electrical energy to be utilized in catalyzing the PMS process, resulting in a higher degradation efficiency over a certain period compared to other waveform currents. Additionally, polarity switching helps prevent the accumulation of fouling layers on the electrodes, a common issue during extended operation in a fixed polarity mode that can hinder degradation processes [31]. Furthermore, pulsed current systems have been observed to generate novel oxidative degradation pathways, potentially leading to the formation of additional reactive oxygen species (ROS), which are essential for achieving efficient and environmentally friendly degradation [24].
Therefore, waveforms that can adjust generation time under low energy consumption may offer significant advantages. Furthermore, this article aims to enhance the general understanding in this field regarding the application of pulsed current.
During experimentation, optimization experiments were managed based on the SIN/PMS/Fe(III) system, considering initial pH, voltage, alternating current frequency, and iron ion concentration [32,33].
Figure 2a,d illustrate the removal rate of SMX by SIN/PMS/Fe(III) at different initial pH values. An initial pH of 3 was chosen in accordance with the research literature as the first-rank condition for the EAOP process [34], and it resulted in the complete degradation of pollutants under the optimal conditions in this study. As the initial pH increased, the removal rate of SMX decreased, but it is observed that the degradation increased again until the initial pH reached 9~11. This phenomenon may be related to the hydrolysis and precipitation of Fe(III) in an alkaline environment. Some scholars have also noted that at an initial pH of 3 (acidic conditions), the generated OH exhibits a higher potential (2.7 V), and simultaneously, the generated Fe(OH)2+ possesses a certain stability in the system [35]. When the initial pH increases to 9~11, PMS decomposes spontaneously to generate 1O2 to promote degradation [36], as shown in Equation (1), and 1O2 is also an important component of the active species for SMX degradation.
HSO 5 + SO 5 2 HSO 4 + SO 4 2 + O 2 1
Figure 2b,e illustrate the degradation efficiency of SIN/PMS/Fe(III) on SMX at different voltages. As sinusoidal current was employed, the voltage values represented the effective values of the sinusoidal voltage. When the voltages were 3 V, 4 V, 5 V, and 6 V, the efficiencies of the removal of SMX were 33.8, 64.3, 70.0, and 100, respectively. The voltage increased from 2 V to 6 V, achieving complete degradation under the conditions of 6 V. With increasing voltage, the oxidative performance on the anode also increased, simultaneously activating the catalytic performance of PMS to generate OH, O2•−, SO4•−, 1O2, and other active radicals for SMX degradation [37]. However, simultaneous observations of oxygen evolution reaction (OER) as well as hydrogen evolution reaction (HER) phenomena are noted, evidenced by bubble formation on the electrode plate, indicating that a portion of electrical energy was reluctantly utilized in the electrolysis of water processes, as depicted in Equations (2)–(4) [38,39].
H 2 O OH + H + + e O + 2 H + + e
O + H 2 O OOH O 2 + H + + e
2 H + + 2 e H 2
The EE/O value is typically used to indicate the energy required (kWh/m3) to decompose the target pollutant in 1 m3 of contaminated water matrix. The EE/O value was calculated based on the following equation for the degradation curves that fitted the pseudo-first-order kinetic model [40].
E E / O = φUIt Vlog ( C 0 C t )
where φ, U, and I denote the pulse duty cycle, applied voltage (kV), and current (A) of the power supply, and V, C0, and Ct refer to the wastewater volume, the initial pollutant concentration, and the specific concentration at reaction time t, respectively.
The relationship between different voltages and the corresponding EE/O values is illustrated in Figure S2. As shown, when the voltage was set to 6 V, the EE/O value decreased while the removal efficiency simultaneously reached 100%. Based on these results, the parameter of 6 V was obviously the optimal voltage condition because the highest energy efficiency (EE) and degradation performance were observed in such conditions, despite minor side reactions likely occurring.
Figure 2c,f present the degradation efficacy of SIN/PMS/Fe(III) on SMX under varying alternating current frequencies. Remarkably, when the frequencies were set at 0.1, 1, 10, 100, and 1000 kHz, the removal efficiency of SMX reached 100%. This outcome can be characterized as the notion that a pulse of higher frequency reduces the on-time within the pulse cycle, enhancing the material transfer process. This enhancement facilitates the utilization of OH radicals while reducing their recombination reactions. Consequently, the drawback of low pulse efficiency is mitigated by using a higher pulse frequency [41,42]. As a result, we selected 1 kHz as the optimal parameter.
In the experimental investigation of the impact of Fe(III) concentration on the removal efficiency of SMX, it increased continuously as the Fe(III) concentration increased from 0.1 mmol/L to 0.6 mmol/L, which is depicted in Figure 3a,b. This trend indicates that the decomposition of PMS catalyzes the production of more active species. Notably, SMX achieved complete removal while the Fe(III) concentration reached 0.6 mmol/L, and the following reactions occurred in the solution as shown in Equations (6)–(8) [43,44].
F e 3 + + e Fe 2 +
F e 2 + + HSO 5 F e 3 + + SO 4 + OH
F e 2 + + HSO 5 F e 3 + + SO 4 2 + OH
However, in Figure 3c,d, when the Fe(III) concentration exceeded 0.6 mmol/L, SMX degradation was hindered. At Fe(III) concentrations of 0.6, 0.7, 0.8, 0.9, and 1.0 mmol/L, the efficiency of the removal of SMX was 100%, 80.3%, 81.3%, 84.7%, and 83.9%, respectively. This is attributed to the excessive Fe(III) concentration leading to the formation of more Fe(OH)3 precipitated on the cathode. In addition, Fe(OH)3 at 1 mmol/L at pH 3 has already shown precipitation, and some OH- was generated in the surrounding area of the cathode; thus, on the cathode, the Fe(OH)3 precipitate was clearly deposited [45,46], thereby reducing the active sites for H2O2 generation [47], resulting in a decreased SMX removal efficiency.

2.2. Effectiveness of Iron Species and PMS in SIN/PMS/Fe(III) Process

Simultaneously, as shown in Figure 4a, by measuring the concentrations of Fe2+ and total iron ions, it can be observed that the average contents of Fe(II) and total iron during the experiment are 4.6 and 251.5 μmol/L, respectively. As the conversion from Fe(II) to Fe(III) has been identified as the rate-limiting step within the process [48], it is demonstrated that sinusoidal waveform currents effectively utilize Fe(II) within the system for catalytic conversion. Additionally, some iron ions adsorbed onto the electrode plate from the outset of the reaction, resulting in the decrease in total iron [44]. Nonetheless, a significant portion of iron ions remained in a free form in the container, participating in catalytic cycles.
In the experiment, the 2,2′-azino-bis(3-ethylbenzothiazoline-6-sulfonate) (ABTS) method was employed in measuring the content of PMS [49,50] in Figure 4b. It was observed that although PMS acted as a catalyst, its content also decreased over time. This reduction is partly due to electrode adsorption, reducing the PMS content in the solution. Additionally, some of the PMS may react with SO42− in the solution, producing weaker oxidizing SO5•− radicals (as shown in Equations (9)–(11)), leading to the depletion of PMS [51,52].
SO 4 + HSO 5 SO 5 + SO 4 2 + H +
SO 4 + SO 4 S 2 O 8 2
SO 4 + OH HSO 5

2.3. Removal Efficiency of SMX in Various Processes

From Figure 5a,b, it is evident that in the absence of PMS and iron ions, and in the absence of electrical current, the concentration of SMX remained virtually unchanged. This observation aligns with the characteristic behavior of SMX as a persistent organic pollutant (POPs), while also indicating the weak adsorption of SMX onto DSA and graphite electrodes. Apart from the SIN/PMS/Fe(III) system, which achieved completely degraded SMX, the SMX removal efficiency in the remaining systems was only around 35%. Specifically, the system with only PMS and Fe(III) added demonstrated that the immediate oxidation of PMS and the adsorption of Fe(III) alone were insufficient in effectively degrading SMX [53]. Within the PMS/Fe(III) system, without the support of electrical current, Fe(III) did not exhibit a crucial catalytic effect on PMS, and their synergistic action also failed to efficiently remove SMX. The term “Switched” refers to experiments where DSA served as the cathode while the graphite electrode served as the anode, without adding any other substances, and only direct current was passed for SMX degradation. Continuous direct current electrolyzed water on the electrode, generating H2O2 at the anode for SMX degradation. However, the hydrolysis-generated radicals were also ineffective in degrading SMX. The system with only electrical current demonstrates that hydroxyl radicals are produced on the Ti/RuO2-IrO2 anode throughout immediate electron transfer or Equation (12) for SMX degradation [54].
M ( OH ) M O + H + + e

2.4. Identification of Reactive Species and Mechanisms of SIN/PMS/Fe(III) Process

2.4.1. EPR Test

For the purpose of examining the positive species produced during various systems, EPR spin trapping experiments were utilized, which employed DMPO and TEMP as probes to detect SO4•−/OH and 1O2, respectively, which is illustrated in Figure 6a. These experiments aimed to elucidate the active radicals generated within specific processes [55]. Prior scholarly investigations have indicated the absence of OH and SO4•− generation in systems where only PMS is introduced [30]. Consequently, this serves as a critical baseline reference to discern the quantification of DMPO-OH or DMPO-SO4•− adducts generated within alternative systems [56]. Notably, the hierarchical arrangement of systems based on OH and SO4•− production, delineated from highest to lowest, is SIN/PMS/Fe(III) > DC/PMS/Fe(III) > SIN/PMS > PMS, suggesting the potential for sinusoidal waveform current systems to catalytically yield a higher abundance of active species.
In the context of EPR investigations targeting 1O2 with TEMP [57] in Figure 6b, a blank control value was established by initially employing TEMP prepared with ultrapure water. Significantly, the discernible production of the TEMP-1O2 adduct within the SIN/PMS/Fe(III) system underscores the presence of 1O2 within this specific experimental framework.

2.4.2. Quenching Experiment

In the quenching experiments, the quenching agents used were tert-butyl alcohol (TBA), ethanol, furfuryl alcohol (FFA), 1,4-benzoquinone (P-BQ), L-histidine, and superoxide dismutase (SOD); the reaction rate coefficients of various quenching agents for different reactive radicals are shown in Table S1; and the quenching efficiency is shown in Figure 7a.
TBA is an effective scavenger of OH radicals [58], while ethanol is also an efficient scavenger for both OH and SO4•− radicals [59]. Upon the addition of 100 mmol/L TBA, the removal rate of SMX decreased to 80.0%, whereas the addition of 100 mmol/L ethanol reduced the removal rate of SMX to 75.3%, reflecting the possible existence of OH and SO4•− radicals within the system. FFA, as a primary scavenger of 1O2, exhibits a high reaction rate [60]. Considering FFA’s strong reactivity with OH, its concentration was adjusted to one hundredth the concentration of TBA, 1 mmol/L, to eliminate the influence of OH [61]. From the graph, it can be observed that upon adding 1 mmol/L of FFA, the removal rate of SMX reduced to 45.9%, indicating the generation of 1O2 within this system. However, FFA may also react with PMS within the process, potentially interfering with the experimental results; thus, L-histidine was employed to scavenge 1O2 for further verification [62]. Upon adding 10 mmol/L of L-histidine, the removal rate of SMX reduced to 38.5%, also identifying the presence of 1O2 within the SIN/PMS/Fe(III) system. P-BQ exhibits high reaction rates for scavenging both OH and O2•− radicals, and upon adding 5 mmol/L of P-BQ, the removal rate of SMX reduced to 55.0%. Subsequently, the use of SOD, which exhibits high reactivity towards O2•− radicals, was employed for further validation [63]. Upon adding 500 U/mL of SOD, the degradation efficiency of SMX reduced to 63.6%, identifying the presence of O2•− radicals within the system.

2.4.3. Chemical Probe Experiments

Further evidence supporting the production of OH within the SIN/PMS/Fe(III) process was garnered through the coumarin (COU) fluorescence probing technique [64], utilizing COU, one kind of chemical probe. COU exhibits a notably high reaction constant with OH, yielding the principal byproduct 7-hydroxycoumarin (7-HC) [65], depicted in Equation (13).
COU + 2 OH 7 HC + H 2 O
Depicted in Figure 7b, experimental observations revealed a decrement in 7-HC concentration concomitant with an escalation in the concentration of the aforementioned byproducts over a 30 min reaction period, thereby pointing to the unequivocal production of OH during the SIN/PMS/Fe(III) process.
Furthermore, the detection of SO4•− generation within the SIN/PMS/Fe(III) process was approached indirectly through the p-benzoquinone (p-BQ) formation assessment, which occurred primarily when p-hydroxybenzoic acid (p-HBA) reacted with SO4•− [48]. The findings were derived from the probe experiments. Figure 7c suggests that the SIN/PMS/Fe(III) process produced SO4•− while the concentration was 6.67 µmol/L. Comparatively, earlier research indicated that the SO4•− concentration in the DC/PMS/Fe(III) process was 5.61 µmol/L [30]. This demonstrates that the SIN system produced 18 ± 1% more SO4•− than the DC system, highlighting one reason for the excellent degradation of the SIN/PMS/Fe(III) process.
Recent research has identified Fe(IV) as an available species in the Fe(II)/PMS process [66,67]. During this experiment, methyl phenyl sulfoxide (PMSO) was employed as a probe for Fe(IV), with PMSO2 detected as a product using HPLC. The production of Fe(IV) in the SIN/PMS/Fe(III) process occurred through the reaction between Fe(II) and PMS depicted in Equation (14).
HS O 5 + Fe 2 + Fe 4 O 2 + + SO 4 2 + H +
Figure 7c shows a rapid increase in PMSO2 concentration during the SIN/PMS/Fe(III) process. Additionally, the entire formation of PMSO2 in the SIN/PMS/Fe(III) process was 46.18 µmol/L. By comparison, previous studies reported a PMSO2 concentration of 44.06 µmol/L in a DC system [61], indicating that Fe(IV) oxidized PMSO to produce PMSO2 more efficiently and generated Fe(IV) in greater quantities in the SIN/PMS/Fe(III) process than that while using direct current.
In the experiment, Nitroblue Tetrazolium (NBT) was employed as the chemical probe to measure the superoxide radicals (O2•−) that may be generated by the reaction [68], depicted in Figure 7d. The absorbance of NBT ranged from 295 nm to 560 nm, and the formazan (MF) and diformazan (DF) formed by the reaction between NBT and O2•−, whose generation depicted in Equations (15) and (16) reduced the original absorbance of NBT.
NBT + 2 O 2 MF
NBT + 4 O 2 DF
Therefore, the full-spectrum scanning method was used for measurement. From the results, it can be observed that the absorbance of NBT decreased progressively from 0 min to 30 min, indicating that the production of O2•− accumulated over time in the experiment. Compared with previous studies [69,70], it was found that a significantly higher amount of O2•− is produced in the SIN/PMS/Fe(III) process, which may also be the answer to the great degradation efficiency of the SIN/PMS/Fe(III) system. In addition, O2•− is generally believed to have been produced in the following ways (Equations (17) and (18)):
H 2 O 2 + SO 4 2 HSO 4 + HO 2
HO 2 O 2 + H +

2.5. Degradation Pathways of SMX and Toxicity Evaluation

Figure 8 illustrates the byproducts and pathways from the degradation of SMX employing the SIN/PMS/Fe(III) process. Through UHPLC-QTOF-MS analysis, eleven byproducts of SMX were identified and are detailed in Table S2, listing their molecular formulas, m/z values, and structures.
Furthermore, Figure 9 depicts the outcomes of toxicity assessment conducted employing the Toxicity Estimation Software Tool (T.E.S.T.) version 4.1 [71,72,73,74]. This implement utilizes a quantitative structure–activity relationship (QSAR) prediction matrix to evaluate the developmental toxicity, mutagenicity, acute toxicity (LD50), and bioaccumulation factor of SMX, as well as its byproducts. As shown in Figure 9a, the LD50 values for byproducts P284, P258, P266, P142, P99, P114, P255, P300, P503, P519, and P239 were 5384.89 mg/kg, 3158.52 mg/kg, 2663.45 mg/kg, 503.46 mg/kg, 383.09 mg/kg, 263.03 mg/kg, 33.88 mg/kg, 3256.46 mg/kg, 3215.08 mg/kg, 2764.26 mg/kg, and 6856.51 mg/kg, respectively. Byproducts exhibiting LD50 values lower than 1000 mg/kg were designated as possessing toxic properties. In Figure 9b, all byproducts exhibited lower bioaccumulation factors compared to SMX, except for P284, P284, and P239, indicating a potential decrease in bioaccumulation during degradation. Figure 9c illustrates that SMX is categorized as a “developmental toxicant” based on its developmental toxicity value of 0.85. Most intermediates also demonstrated developmental toxicity, with values greater than 0.5. The results for P142 and P99, which had developmental toxicity values less than 0.5, indicated that these intermediates were “developmental non-toxicants”. As depicted in Figure 9d, only P99 showed a mutagenicity value higher than 0.50, indicating that it was mutagenicity-positive. In contrast, both SMX and the other byproducts tested negative for mutagenicity. As shown in Figure 9, most byproducts had lower LD50 values than SMX, with only P284, P239, and another P284 showing a higher bioaccumulation potential. P142 and P99 exhibited developmental toxicity, and P99 was mutagenicity-positive. Although byproducts may show increased toxicity qualitatively, the general toxicity of the sample decreased due the low abundances of these byproducts. In addition, the decrease in total organic carbon (TOC) after the degradation process indicated that mineralization could be achieved in the current system with an extended reaction time to remove the toxicity of SMX byproducts.

2.6. Practical Application of the SIN/PMS/Fe(III) System

Anions in natural water matrices such as Cl, NO3, and H2PO4 are widely found, and their presence may impact the removal of various contaminants. Hence, the impacts of varying concentrations of Cl, NO3, H2PO4, and humic acid (HA) on SMX removal employing the SIN/PMS/Fe(III) process were explored.
Figure 10a depicts the influences of Cl on the SMX removal. The removal rate of SMX notably rose with higher concentrations of Cl. This appearance can be attributed to the oxidation of Cl to generate Cl2, which subsequently reacts with water to produce free chlorine species (HOCl, ClO) [75,76,77,78] through the following reactions (Equations (19) and (20)), thereby accelerating the degradation of SMX.
2 Cl Cl 2 + 2 e
Cl 2 + H 2 O H O C l + H + + Cl
On the other hand, Cl can be generated from Cl reacting with OH and SO4•− through the following reactions (Equations (21)–(23)). Subsequently, Cl2•− can be further generated from Cl reacting with Cl (Equation (24)) [79,80], which has been shown to boost the removal of other organic contamination, including Cl [77].
Cl + OH ClOH
ClOH + H + Cl + H 2 O
Cl + SO 4 Cl + SO 4 2
Cl + Cl Cl 2
Figure 10b depicts the impact of NO3 on SMX degradation. At concentrations of 1 mmol/L and 5 mmol/L, NO3 revealed a minimal influence on degradation efficiency. This is likely due to the excess NO3 ions potentially scavenging a fraction of both OH and SO4•− radicals, thereby marginally reducing the reaction rate of the system [81]. Figure 10c depicts the impact of H2PO4 on SMX decontamination. Within 30 min each, the removal rate of SMX reduced from 100.0% to 59.6% and 50.3% while the concentration of H2PO4 rose from 0 to 1.0 and 5.0 mmol/L, respectively. These phenomena showed that H2PO4 may not facilitate the activation of PMS, as observed in preceding experiments [82,83]. Moreover, the energy barrier for PMS activation by H2PO4 appears to be too high to overcome, indicating its ineffectiveness in activating PMS. Figure 10d illustrates the influence of HA on SMX removal. Despite HA exhibiting a prohibitive effect on SMX degradation, the degradation efficiency reached 100.0% after 30 min in the experiment. This slight prohibitive effect is attributed to the carboxyl and hydroxyl groups within HA molecules potentially quenching some active species in the system, thereby reducing SMX degradation [81,84,85].
Figure 10e demonstrates how real-world water matrices impact SMX removal within the SIN/PMS/Fe(III) process. The data reveal a modest reduction in the efficiency of SMX removal when tap water and lake water are utilized as matrices in comparison with ultrapure water. This decline is attributed to the presence of diverse anions and humic acids (HAs) in natural water samples, which hinder the degradation process [60].
Moreover, Figure 10f shows the outcomes of the cyclic testing for SMX degradation. Under consistent parameters in experiments, the reusability of the SIN/PMS/Fe(III) process was intuitionistic to evaluate. Remarkably, with six consecutive cycles, this system has maintained the full degradation efficiency of SMX within a 30 min reaction period, underscoring its exceptional reusability. In conclusion, these findings affirm the robustness of the SIN/PMS/Fe(III) process for effectively treating SMX in real-world water matrices, ensuring a high sustained degradation efficiency and system stability.

3. Material and Methods

3.1. Chemicals and Reagents

All the chemicals utilized in this study are listed in Text S1 of the Supplementary Materials.

3.2. Degradation Experiment

The degradation experiment of SMX was operated in an open container of a glass cylindrical vessel, with a diameter of 75 mm and height of 80 mm, and it had three openings on its cover. Two of these openings were used for placing electrodes, while the third was utilized for adding and sampling. The anode used for degradation was DSA with dimensions of 3.0 cm × 3.0 cm × 0.1 cm, and the cathode was a graphite electrode with dimensions of 3.0 cm × 3.0 cm × 0.1 cm. After each experiment, ultrasonic oscillation rinsing was performed. The power supply used for this experiment was a dual-channel function/arbitrary waveform generator (DG1032Z, RIGOL, Shanghai, China), capable of providing current with different waveforms in the 0.1–100 kHz frequency range and 0–10 V voltage range. To ensure stable operation and meet the desired current requirements, a power amplifier (PA1011, RIGOL, Shanghai, China) was also employed for experiments.
The degradation process primarily goes as follows: Firstly, the pollutant SMX, PMS, Na2SO4, Fe2(SO4)3, and ultrapure water are added into the reaction vessel, resulting in a final volume of 150 mL, with a concentration of SMX of 8 μmol/L, PMS of 0.001 mol/L, Fe3+ of 0.6 mol/L, and Na2SO4 of 0.02 mol/L. Secondly, the dropwise addition of 100 mmol/L of H2SO4 or NaOH is performed to coordinate the pH to stabilize at 3. The reaction time is set to 30 min, and samples of 1 mL of the solution while reacting are gathered at 0, 5, 10, 15, 20, 25, and 30 min for the detection of SMX concentration. Thirdly, the samples are placed in sample vials with a volume of 2 mL, and 40 μL of Na2S2O3 is added as a terminating agent in advance. A control group is set up for error analysis in this experiment, and all data displayed in the figures and tables of the paper are covered as the mean and standard deviation derived from replicate samples.

3.3. Analytical Methods

Equipped with an Eclipse XDB C-18 column (5 µm, 250 × 4.6 mm, Agilent, Santa Clara, CA, USA), high-powered liquid chromatography (HPLC, Agilent 1260, Agilent, Santa Clara, CA, USA) was utilized to measure the concentrations of SMX and the rest of the model compounds as well. The specifics of the method are shown in Text S2 and Table S1.
Ultra-performance liquid chromatography quadrupole time-of-flight mass spectrometry (UHPLC-QTOF-MS, 6545 QTOF system, Agilent, Santa Clara, CA, USA) is capable of identifying the byproducts from SMX degradation, and an electrospray ionization (ESI) source was operated in positive mode. More method specifics are shown in Text S3 as well. UV-vis spectroscopy (UV-L5S, Shanghai, China) was utilized to measure iron species and PMS concentrations with more method details, respectively, outlined in Text S4 and Text S5.
Electron Paramagnetic Resonance (EPR) spectroscopy spectra were obtained using an EPR200-Plus instrument (CIQTEK, Hefei, China), and the methodology is outlined in Text S6. Reactive oxygen species (ROS) concentrations were quantified using chemical probe experiments, as detailed in Text S7.

3.4. Theoretical Calculation

The Gaussian 09 package and Multiwfn software were employed to represent density functional theory (DFT) calculations, which underpin the analyses presented throughout this study. The 6-31g(d,p) basis set was employed to optimize the geometry of the condensed Fukui index and the Highest Occupied Molecular Orbital (HOMO) of SMX and its byproducts, with detailed methodologies provided in Text S8. The T.E.S.T. was used to estimate acute toxicity, developmental toxicity, mutagenicity, and bioaccumulation factors of SMX and byproducts from its degradation, leveraging foreseen QSAR models.

4. Conclusions

In this work, the SIN/PMS/Fe(III) system has been established with an exceptionally high removal efficiency of SMX, reaching a complete degradation rate (100%). Through optimization of the system’s experimental parameters, it was determined that effective SMX degradation could be accomplished under the following conditions: pH of 3, voltage of 6 V, and concentrations of PMS and Fe3+ of 1.0 mmol/L and 0.6 mmol/L, respectively. In addition, the concentrations of Fe3+ and PMS did not fully deplete after the reaction, indicating the system’s potential for sustained reactivity. Results of EPR, quenching, and probe experiments further demonstrated that the SIN/PMS/Fe(III) process generated higher levels of OH, SO4•−, 1O2, Fe(IV), and O2•− radicals in comparison with the DC system, which likely accounts for its high efficiency. Toxicity assessments of byproducts from SMX degradation employing the T.E.S.T. with the QSAR model indicated slightly increased toxicity of the byproducts. This study also investigated, within the SIN/PMS/Fe(III) system, the impacts of various anions and water matrices on SMX removal, revealing negligible effects from these factors and emphasizing the system’s resilience. This study reveals the significance of new knowledge that the choice of current waveforms of pulsed current may lead to elevated energy efficiency and degradation performance. Overall, these results underscore the SIN/PMS/Fe(III) process’s high efficiency and successful capabilities of degradation, indicating its potential as a dependable approach for effectively treating authentic wastewater.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal14080532/s1, Text S1, chemicals and reagents; Text S2, HPLC measurement method; Table S1, parameters of HPLC methods; Text S3, UHPLC-QTOF-MS analysis method; Text S4, UV-vis method for determination of iron species; Text S5, UV-vis method for determination of PMS; Text S6, EPR analysis of reactive radicals; Text S7, chemical probe methods; Text S8, DFT calculation process; Figure S1 shows the relative abundances of Fe(II) species at pH 1.0–12.0; Figure S2 shows the values of EE/O at different voltage in SIN/PMS/Fe(III) system; Figure S3 show the different waveforms of current applied in the PE/PMS/Fe(III) system; Table S2 shows the reaction rate coefficients of various quenching agents for different reactive radicals; Table S3 shows the identified SMX byproducts in the SIN/Fe(III)/PMS system based on the UHPLC-QTOF-MS analysis.

Author Contributions

Conceptualization, R.H.; Methodology, J.F., I.Y.Z. and R.H.; Formal analysis, J.F., Y.W. and I.Y.Z.; Investigation, J.F., I.Y.Z. and R.H.; Resources, J.F., J.W. and R.H.; Data curation, J.W. and Y.W.; Writing—original draft, J.F.; Writing—review & editing, R.H.; Supervision, I.Y.Z. and R.H.; Project administration, R.H.; Funding acquisition, R.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China with grant number (21973015, 22125301). This work was also supported by the 14th Recruitment Program of Young Professionals of IYZ, the funding of the Innovative research team of high-level local universities in Shanghai, and a key laboratory program of the Education Commission of Shanghai Municipality (ZDSYS14005). The authors would like to acknowledge the Initialization Fund for Talents of Sichuan University (No. YJ202033) and the Special Funds for Scientific and Technological Collaboration of SCU-Zigong (No. 2022CDZG-3). This work was also supported by the State Key Laboratory of Physical Chemistry of Solid Surfaces (No. 202105) in Xiamen University and the Key Laboratory of Spectrochemical Analysis and Instrumentation (Xiamen University), Ministry of Education-SCAI2006.

Data Availability Statement

Data will be provided upon request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jia, Y.; Khanal, S.K.; Zhang, H.; Chen, G.-H.; Lu, H. Sulfamethoxazole degradation in anaerobic sulfate-reducing bacteria sludge system. Water Res. 2017, 119, 12–20. [Google Scholar] [CrossRef]
  2. Langbehn, R.K.; Michels, C.; Soares, H.M. Antibiotics in wastewater: From its occurrence to the biological removal by environmentally conscious technologies. Environ. Pollut. 2021, 275, 116603. [Google Scholar] [CrossRef] [PubMed]
  3. Karkman, A.; Thi Thuy, D.; Walsh, F.; Virta, M.P.J. Antibiotic-Resistance Genes in Waste Water. Trends Microbiol. 2018, 26, 220–228. [Google Scholar] [CrossRef]
  4. Hu, Z.; Cai, J.; Song, G.; Tian, Y.; Zhou, M. Anodic oxidation of organic pollutants: Anode fabrication, process hybrid and environmental applications. Curr. Opin. Electrochem. 2021, 26, 100659. [Google Scholar] [CrossRef]
  5. Krishnan, R.Y.; Manikandan, S.; Subbaiya, R.; Biruntha, M.; Govarthanan, M.; Karmegam, N. Removal of emerging micropollutants originating from pharmaceuticals and personal care products (PPCPs) in water and wastewater by advanced oxidation processes: A review. Environ. Technol. Innov. 2021, 23, 101757. [Google Scholar] [CrossRef]
  6. Xiong, Z.; Zhang, H.; Zhang, W.; Lai, B.; Yao, G. Removal of nitrophenols and their derivatives by chemical redox: A review. Chem. Eng. J. 2019, 359, 13–31. [Google Scholar] [CrossRef]
  7. Feng, Y.; Yang, L.; Liu, J.; Logan, B.E. Electrochemical technologies for wastewater treatment and resource reclamation. Environ. Sci.-Wat. Res. Technol. 2016, 2, 800–831. [Google Scholar] [CrossRef]
  8. Garcia-Segura, S.; Ocon, J.D.; Chong, M.N. Electrochemical oxidation remediation of real wastewater effluents—A review. Process Saf. Environ. Protect. 2018, 113, 48–67. [Google Scholar] [CrossRef]
  9. Saravanan, A.; Deivayanai, V.C.; Kumar, P.S.; Rangasamy, G.; Hemavathy, R.V.; Harshana, T.; Gayathri, N.; Alagumalai, K. A detailed review on advanced oxidation process in treatment of wastewater: Mechanism, challenges and future outlook. Chemosphere 2022, 308, 136524. [Google Scholar] [CrossRef]
  10. Sires, I.; Brillas, E.; Oturan, M.A.; Rodrigo, M.A.; Panizza, M. Electrochemical advanced oxidation processes: Today and tomorrow. A review. Environ. Sci. Pollut. Res. 2014, 21, 8336–8367. [Google Scholar] [CrossRef]
  11. Luo, Z.; Liu, M.; Tang, D.; Xu, Y.; Ran, H.; He, J.; Chen, K.; Sun, J. High H2O2 selectivity and enhanced Fe2+ regeneration toward an effective electro-Fenton process based on a self-doped porous biochar cathode. Appl. Catal. B-Environ. 2022, 315, 121523. [Google Scholar] [CrossRef]
  12. Zhang, M.-H.; Dong, H.; Zhao, L.; Wang, D.-X.; Meng, D. A review on Fenton process for organic wastewater treatment based on optimization perspective. Sci. Total Environ. 2019, 670, 110–121. [Google Scholar] [CrossRef]
  13. Abdalrhman, A.S.; El-Din, M.G. Degradation of organics in real oil sands process water by electro-oxidation using graphite and dimensionally stable anodes. Chem. Eng. J. 2020, 389, 124406. [Google Scholar] [CrossRef]
  14. Sopaj, F.; Rodrigo, M.A.; Oturan, N.; Podvorica, F.I.; Pinson, J.; Oturan, M.A. Influence of the anode materials on the electrochemical oxidation efficiency. Application to oxidative degradation of the pharmaceutical amoxicillin. Chem. Eng. J. 2015, 262, 286–294. [Google Scholar] [CrossRef]
  15. Hoyos-Arbelaez, J.; Vazquez, M.; Contreras-Calderon, J. Electrochemical methods as a tool for determining the antioxidant capacity of food and beverages: A review. Food Chem. 2017, 221, 1371–1381. [Google Scholar] [CrossRef] [PubMed]
  16. Zhang, L.; Qin, D.; Feng, J.; Tang, T.; Cheng, H. Rapid quantitative detection of luteolin using an electrochemical sensor based on electrospinning of carbon nanofibers doped with single-walled carbon nanoangles. Anal. Methods 2023, 15, 3073–3083. [Google Scholar] [CrossRef] [PubMed]
  17. Dai, H.; Wang, N.; Wang, D.; Ma, H.; Lin, M. An electrochemical sensor based on phytic acid functionalized polypyrrole/graphene oxide nanocomposites for simultaneous determination of Cd(II) and Pb(II). Chem. Eng. J. 2016, 299, 150–155. [Google Scholar] [CrossRef]
  18. Ratautaite, V.; Boguzaite, R.; Brazys, E.; Ramanaviciene, A.; Ciplys, E.; Juozapaitis, M.; Slibinskas, R.; Bechelany, M.; Ramanavicius, A. Molecularly imprinted polypyrrole based sensor for the detection of SARS-CoV-2 spike glycoprotein. Electrochim. Acta 2022, 403, 139581. [Google Scholar] [CrossRef] [PubMed]
  19. Bu, Y.; Wang, C.; Zhang, W.; Yang, X.; Ding, J.; Gao, G. Electrical Pulse-Driven Periodic Self-Repair of Cu-Ni Tandem Catalyst for Efficient Ammonia Synthesis from Nitrate. Angew. Chem.-Int. Edit. 2023, 62, 17337. [Google Scholar] [CrossRef]
  20. Wang, J.; Yao, J.; Wang, L.; Xue, Q.; Hu, Z.; Pan, B. Multivariate optimization of the pulse electrochemical oxidation for treating recalcitrant dye wastewater. Sep. Purif. Technol. 2020, 230, 115851. [Google Scholar] [CrossRef]
  21. Lei, J.; Xu, Z.; Xu, H.; Qiao, D.; Liao, Z.; Yan, W.; Wang, Y. Pulsed electrochemical oxidation of acid Red G and crystal violet by PbO2 anode. J. Environ. Chem. Eng. 2020, 8, 103773. [Google Scholar] [CrossRef]
  22. Wei, Z.; Xu, H.; Lei, Z.; Yi, X.; Feng, C.; Dang, Z. A binder-free electrode for efficient H2O2 formation and Fe2+ regeneration and its application to an electro-Fenton process for removing organics in iron-laden acid wastewater. Chin. Chem. Lett. 2022, 33, 920–925. [Google Scholar] [CrossRef]
  23. Poza-Nogueiras, V.; Pazos, M.; Angeles Sanroman, M.; Gonzalez-Romero, E. Double benefit of electrochemical techniques: Treatment and electroanalysis for remediation of water polluted with organic compounds. Electrochim. Acta 2019, 320, 134628. [Google Scholar] [CrossRef]
  24. Wei, J.; Zhu, X.; Ni, J. Electrochemical oxidation of phenol at boron-doped diamond electrode in pulse current mode. Electrochim. Acta 2011, 56, 5310–5315. [Google Scholar] [CrossRef]
  25. Li, J.; Lin, H.; Zhu, K.; Zhang, H. Degradation of Acid Orange 7 using peroxymonosulfate catalyzed by granulated activated carbon and enhanced by electrolysis. Chemosphere 2017, 188, 139–147. [Google Scholar] [CrossRef] [PubMed]
  26. Jing, J.; Wang, X.; Zhou, M. Electro-enhanced activation of peroxymonosulfate by a novel perovskite-Ti4O7 composite anode with ultra-high efficiency and low energy consumption: The generation and dominant role of singlet oxygen. Water Res. 2023, 232, 119682. [Google Scholar] [CrossRef]
  27. Zhao, Y.; Sun, M.; Zhao, Y.; Wang, L.; Lu, D.; Ma, J. Electrified ceramic membrane actuates non-radical mediated peroxymonosulfate activation for highly efficient water decontamination. Water Res. 2022, 225, 119140. [Google Scholar] [CrossRef] [PubMed]
  28. Wang, Y.; Sun, T.; Tong, L.; Gao, Y.; Zhang, H.; Zhang, Y.; Wang, Z.; Zhu, S. Non-free Fe dominated PMS activation for enhancing electro-Fenton efficiency in neutral wastewater. J. Electroanal. Chem. 2023, 928, 117062. [Google Scholar] [CrossRef]
  29. Fu, H.; Zhao, P.; Xu, S.; Cheng, G.; Li, Z.; Li, Y.; Li, K.; Ma, S. Fabrication of Fe3O4 and graphitized porous biochar composites for activating peroxymonosulfate to degrade p-hydroxybenzoic acid: Insights on the mechanism. Chem. Eng. J. 2019, 375, 121980. [Google Scholar] [CrossRef]
  30. Wang, J.; Long, X.; Zhang, I.Y.; Huang, R. Pulsed versus direct current electrochemical co-catalytic peroxymonosulfate-based system: Elevated degradation and energy efficiency with enhanced oxidation mechanisms. J. Hazard. Mater. 2023, 458, 132004. [Google Scholar] [CrossRef]
  31. Nidheesh, P.V.; Ganiyu, S.O.; Martinez-Huitle, C.A.; Mousset, E.; Olvera-Vargas, H.; Trellu, C.; Zhou, M.; Oturan, M.A. Recent advances in electro-Fenton process and its emerging applications. Crit. Rev. Environ. Sci. Technol. 2023, 53, 887–913. [Google Scholar] [CrossRef]
  32. Chen, X.; Han, Y.; Gao, P.; Li, H. New insight into the mechanism of electro-assisted pyrite minerals activation of peroxymonosulfate: Synergistic effects, activation sites and electron transfer. Sep. Purif. Technol. 2021, 274, 118817. [Google Scholar] [CrossRef]
  33. Chen, X.; Zhao, N.; Hu, X. A novel strategy of pulsed electro-assisted pyrite activation of peroxymonosulfate for the degradation of tetracycline hydrochloride. Sep. Purif. Technol. 2022, 280, 119781. [Google Scholar] [CrossRef]
  34. Zhou, P.; Ren, W.; Nie, G.; Li, X.; Duan, X.; Zhang, Y.; Wang, S. Fast and Long-Lasting Iron(III) Reduction by Boron Toward Green and Accelerated Fenton Chemistry. Angew. Chem.-Int. Edit. 2020, 59, 16517–16526. [Google Scholar] [CrossRef]
  35. Meng, S.; Zhou, P.; Sun, Y.; Zhang, P.; Zhou, C.; Xiong, Z.; Zhang, H.; Liang, J.; Lai, B. Reducing agents enhanced Fenton-like oxidation (Fe(III)/Peroxydisulfate): Substrate specific reactivity of reactive oxygen species. Water Res. 2022, 218, 118412. [Google Scholar] [CrossRef]
  36. Wang, J.; Zhou, W.; Li, J.; Yang, C.; Meng, X.; Gao, J. Insights into the effects of pulsed parameters on H2O2 synthesis by two-electron oxygen reduction under pulsed electrocatalysis. Electrochem. Commun. 2023, 146, 107414. [Google Scholar] [CrossRef]
  37. Chen, L.; Li, X.; Zhang, J.; Fang, J.; Huang, Y.; Wang, P.; Ma, J. Production of Hydroxyl Radical via the Activation of Hydrogen Peroxide by Hydroxylamine. Environ. Sci. Technol. 2015, 49, 10373–10379. [Google Scholar] [CrossRef] [PubMed]
  38. Shi, H.; Zhou, Y.-T.; Yao, R.-Q.; Wan, W.-B.; Ge, X.; Zhang, W.; Wen, Z.; Lang, X.-Y.; Zheng, W.-T.; Jiang, Q. Spontaneously separated intermetallic Co3Mo from nanoporous copper as versatile electrocatalysts for highly efficient water splitting. Nat. Commun. 2020, 11, 2940. [Google Scholar] [CrossRef]
  39. Lu, Z.; Xu, W.; Zhu, W.; Yang, Q.; Lei, X.; Liu, J.; Li, Y.; Sun, X.; Duan, X. Three-dimensional NiFe layered double hydroxide film for high-efficiency oxygen evolution reaction. Chem. Commun. 2014, 50, 6479–6482. [Google Scholar] [CrossRef]
  40. Bolton, J.R.; Bircher, K.G.; Tumas, W.; Tolman, C.A. Figures-of-merit for the technical development and application of advanced oxidation technologies for both electric- and solar-driven systems (IUPAC Technical Report). Pure Appl. Chem. 2001, 73, 627–637. [Google Scholar] [CrossRef]
  41. Lv, F.; Sun, M.; Hu, Y.; Xu, J.; Huang, W.; Han, N.; Huang, B.; Li, Y. Near-unity electrochemical conversion of nitrate to ammonia on crystalline nickel porphyrin-based covalent organic frameworks. Energy Environ. Sci. 2023, 16, 201–209. [Google Scholar] [CrossRef]
  42. Ma, X.; He, C.; Yan, Y.; Chen, J.; Feng, H.; Hu, J.; Zhu, H.; Xia, Y. Energy-efficient electrochemical degradation of ciprofloxacin by a Ti-foam/PbO2-GN composite electrode: Electrode characteristics, parameter optimization, and reaction mechanism. Chemosphere 2023, 315, 137739. [Google Scholar] [CrossRef]
  43. Long, X.; Huang, R.; Li, Y.; Wang, J.; Zhang, M.; Zhang, I.Y. Understanding the electro-cocatalytic peroxymonosulfate-based systems with BDD versus DSA anodes: Radical versus nonradical dominated degradation mechanisms. Sep. Purif. Technol. 2023, 309, 123120. [Google Scholar] [CrossRef]
  44. Kashani, M.R.K.; Wang, Q.; Khatebasreh, M.; Li, X.; Asadi, A.M.S.; Boczkaj, G.; Ghanbari, F. Sequential treatment of landfill leachate by electrocoagulation/aeration, PMS/ZVI/UV and electro-Fenton: Performance, biodegradability and toxicity studies. J. Environ. Manag. 2023, 338, 117781. [Google Scholar] [CrossRef]
  45. Eary, L.E.; Rai, D. Chromate Reduction by Subsurface Soils under Acidic Conditions. Soil Sci. Soc. Am. J. 1991, 55, 676–683. [Google Scholar] [CrossRef]
  46. Tanaka, E.; Hatanaka, N.; Muguruma, M.; Ishikawa, T.; Nakayama, T. Influence of anions on the formation of artificial steel rust particles prepared from acidic aqueous Fe(III) solution. Corros. Sci. 2013, 66, 136–141. [Google Scholar] [CrossRef]
  47. Lee, J.; von Gunten, U.; Kim, J.-H. Persulfate-Based Advanced Oxidation: Critical Assessment of Opportunities and Roadblocks. Environ. Sci. Technol. 2020, 54, 3064–3081. [Google Scholar] [CrossRef]
  48. Hong, Y.; Zhou, H.; Xiong, Z.; Liu, Y.; Yao, G.; Lai, B. Heterogeneous activation of peroxymonosulfate by CoMgFe-LDO for degradation of carbamazepine: Efficiency, mechanism and degradation pathways. Chem. Eng. J. 2020, 391, 123604. [Google Scholar] [CrossRef]
  49. Liu, X.; Fu, Q.; Xu, P.; Zhu, P.; Yang, Z.; Crittenden, J.C. Rapid determination of monopersulfate with bromide ion-catalyzed oxidation of 2, 2′-azino-bis(3-ethylbenzothiazoline-6-sulfonic acid (ABTS). Chem. Eng. J. 2022, 433, 133551. [Google Scholar] [CrossRef]
  50. Zou, J.; Huang, Y.; Zhu, L.; Cui, Z.; Yuan, B. Multi-wavelength spectrophotometric measurement of persulfates using 2,2′-azino-bis(3-ethylbenzothiazoline-6-sulfonate) (ABTS) as indicator. Spectroc. Acta Pt. A-Molec. Biomolec. Spectr. 2019, 216, 214–220. [Google Scholar] [CrossRef]
  51. Xue, Q.; Yu, B.; Dagnew, M.; Li, W.; Ding, H.; Zhang, J.; Sun, Z.; Wang, P.; Zhao, C. Rapid in situ regeneration of phenol-saturated activated carbon fiber by an electro-permonosulfate-ozone process: Performance, operators and mechanism. J. Environ. Chem. Eng. 2024, 12, 111932. [Google Scholar] [CrossRef]
  52. Liu, J.; Zhang, X.; Li, J.; Feng, Y.; Feng, L.; Han, S.; Pan, T.; Zhang, T.; Wu, S.; Ke, Z.; et al. Study on the performance efficiency, mechanism, power consumption and biochemical properties of E/Ce(IV)/PMS on the enhanced removal of RB19. Environ. Res. 2023, 232, 116271. [Google Scholar] [CrossRef]
  53. Miao, F.; Liu, Z.; Kang, X.; Cheng, C.; Mao, X.; Li, R.; Lin, H.; Zhang, H. Electro-enhanced heterogeneous activation of peroxymonosulfate via acceleration of Fe(III)/Fe(II) redox cycle on Fe-B catalyst. Electrochim. Acta 2021, 377, 138073. [Google Scholar] [CrossRef]
  54. Moreira, F.C.; Boaventura, R.A.R.; Brillas, E.; Vilar, V.J.P. Electrochemical advanced oxidation processes: A review on their application to synthetic and real wastewaters. Appl. Catal. B-Environ. 2017, 202, 217–261. [Google Scholar] [CrossRef]
  55. Li, J.; Li, Y.; Xiong, Z.; Yao, G.; Lai, B. The electrochemical advanced oxidation processes coupling of oxidants for organic pollutants degradation: A mini-review. Chin. Chem. Lett. 2019, 30, 2139–2146. [Google Scholar] [CrossRef]
  56. Oh, W.-D.; Dong, Z.; Lim, T.-T. Generation of sulfate radical through heterogeneous catalysis for organic contaminants removal: Current development, challenges and prospects. Appl. Catal. B-Environ. Energy 2016, 194, 169–201. [Google Scholar] [CrossRef]
  57. Liu, Z.; Ding, H.; Zhao, C.; Wang, T.; Wang, P.; Dionysiou, D.D. Electrochemical activation of peroxymonosulfate with ACF cathode: Kinetics, influencing factors, mechanism, and application potential. Water Res. 2019, 159, 111–121. [Google Scholar] [CrossRef]
  58. Lai, L.; Yan, J.; Li, J.; Lai, B. Co/Al2O3-EPM as peroxymonosulfate activator for sulfamethoxazole removal: Performance, biotoxicity, degradation pathways and mechanism. Chem. Eng. J. 2018, 343, 676–688. [Google Scholar] [CrossRef]
  59. Lei, Y.; Yu, Y.; Lei, X.; Liang, X.; Cheng, S.; Ouyang, G.; Yang, X. Assessing the Use of Probes and Quenchers for Understanding the Reactive Species in Advanced Oxidation Processes. Environ. Sci. Technol. 2023, 57, 5433–5444. [Google Scholar] [CrossRef]
  60. Li, J.; Xiong, Z.; Yu, Y.; Wang, X.; Zhou, H.; Huang, B.; Wu, Z.; Yu, C.; Chen, T.; Pan, Z.; et al. Efficient degradation of carbamazepine by electro-Fenton system without any extra oxidant in the presence of molybdate: The role of slow release of iron ions. Appl. Catal. B-Environ. 2021, 298, 120506. [Google Scholar] [CrossRef]
  61. Long, X.; Xiong, Z.; Huang, R.; Yu, Y.; Zhou, P.; Zhang, H.; Yao, G.; Lai, B. Sustainable Fe(III)/Fe(II) cycles triggered by co-catalyst of weak electrical current in Fe(III)/peroxymonosulfate system: Collaboration of radical and non-radical mechanisms. Appl. Catal. B-Environ. 2022, 317, 121716. [Google Scholar] [CrossRef]
  62. Long, X.; Shi, H.; Huang, R.; Gu, L.; Liu, Y.; He, C.-s.; Du, Y.; Xiong, Z.; Liu, W.; Lai, B. Identifying the evolution of primary oxidation mechanisms and pollutant degradation routes in the electro-cocatalytic Fenton-like systems. J. Hazard. Mater. 2023, 445, 130577. [Google Scholar] [CrossRef]
  63. Fotiou, T.; Triantis, T.M.; Kaloudis, T.; O’Shea, K.E.; Dionysiou, D.D.; Hiskia, A. Assessment of the roles of reactive oxygen species in the UV and visible light photocatalytic degradation of cyanotoxins and water taste and odor compounds using C-TiO2. Water Res. 2016, 90, 52–61. [Google Scholar] [CrossRef]
  64. Lai, W.; Chen, Z.; Ye, S.; Xu, Y.; Xie, G.; Kuang, C.; Li, Y.; Zheng, L.; Wei, L. BiVO4 prepared by the sol-gel doped on graphite felt cathode for ciprofloxacin degradation and mechanism in solar-photo-electro-Fenton. J. Hazard. Mater. 2021, 408, 124621. [Google Scholar] [CrossRef]
  65. Tokumura, M.; Morito, R.; Hatayama, R.; Kawase, Y. Iron redox cycling in hydroxyl radical generation during the photo-Fenton oxidative degradation: Dynamic change of hydroxyl radical concentration. Appl. Catal. B-Environ. 2011, 106, 565–576. [Google Scholar] [CrossRef]
  66. Li, X.; Ma, J.; Gao, Y.; Liu, X.; Wei, Y.; Liang, Z. Enhanced atrazine degradation in the Fe(III)/peroxymonosulfate system via accelerating Fe(II) regeneration by benzoquinone. Chem. Eng. J. 2022, 427, 131995. [Google Scholar] [CrossRef]
  67. Li, X.; Yang, S.; Dzakpasu, M.; Xu, S.; Ding, D.; Wang, G.; Chen, R.; Jin, P.; Wang, X.C. Galvanic corrosion of zero-valent iron to intensify Fe2+ generation for peroxymonosulfate activation. Chem. Eng. J. 2021, 417, 128023. [Google Scholar] [CrossRef]
  68. Prasanna, V.L.; Vijayaraghavan, R. Insight into the Mechanism of Antibacterial Activity of ZnO: Surface Defects Mediated Reactive Oxygen Species Even in the Dark. Langmuir 2015, 31, 9155–9162. [Google Scholar] [CrossRef]
  69. Yu, X.; Jin, X.; Li, M.; Yu, Y.; Liu, H.; Zhou, R.; Yin, A.; Shi, J.; Sun, J.; Zhu, L. Mechanism and security of UV driven sodium percarbonate for sulfamethoxazole degradation using DFT and metabolomic analysis. Environ. Pollut. 2023, 323, 121352. [Google Scholar] [CrossRef]
  70. Zhang, S.; Wei, Y.; Metz, J.; He, S.; Alvarez, P.J.J.; Long, M. Persistent free radicals in biochar enhance superoxide-mediated Fe(III)/Fe (II) cycling and the efficacy of CaO2 Fenton-like treatment. J. Hazard. Mater. 2022, 421, 126805. [Google Scholar] [CrossRef]
  71. Ji, Y.; Fan, Y.; Liu, K.; Kong, D.; Lu, J. Thermo activated persulfate oxidation of antibiotic sulfamethoxazole and structurally related compounds. Water Res. 2015, 87, 1–9. [Google Scholar] [CrossRef]
  72. Li, Y.; Zhao, X.; Yan, Y.; Yan, J.; Pan, Y.; Zhang, Y.; Lai, B. Enhanced sulfamethoxazole degradation by peroxymonosulfate activation with sulfide-modified microscale zero-valent iron (S-mFe0): Performance, mechanisms, and the role of sulfur species. Chem. Eng. J. 2019, 376, 03.178. [Google Scholar] [CrossRef]
  73. Qi, C.; Liu, X.; Lin, C.; Zhang, X.; Ma, J.; Tan, H.; Ye, W. Degradation of sulfamethoxazole by microwave-activated persulfate: Kinetics, mechanism and acute toxicity. Chem. Eng. J. 2014, 249, 6–14. [Google Scholar] [CrossRef]
  74. Wang, F.; Fu, H.; Wang, F.-X.; Zhang, X.-W.; Wang, P.; Zhao, C.; Wang, C.-C. Enhanced catalytic sulfamethoxazole degradation via peroxymonosulfate activation over amorphous CoSx@SiO2 nanocages derived from ZIF-67. J. Hazard. Mater. 2022, 423, 126998. [Google Scholar] [CrossRef] [PubMed]
  75. Cheng, H.; Huang, B.; Dai, Y. Engineering BiOX (X = Cl, Br, I) nanostructures for highly efficient photocatalytic applications. Nanoscale 2014, 6, 2009–2026. [Google Scholar] [CrossRef] [PubMed]
  76. Cheng, N.; Zhang, L.; Wang, M.; Shu, J.; Shao, P.; Yang, L.; Meng, X.; Fan, Y.; Li, M. Highly effective recovery of palladium from a spent catalyst by an acid- and oxidation-resistant electrospun fibers as a sorbent. Chem. Eng. J. 2023, 466, 143171. [Google Scholar] [CrossRef]
  77. Wang, Q.; Lu, J.; Jiang, Y.; Yang, S.; Yang, Y.; Wang, Z. FeCo bimetallic metal organic framework nanosheets as peroxymonosulfate activator for selective oxidation of organic pollutants. Chem. Eng. J. 2022, 443, 136483. [Google Scholar] [CrossRef]
  78. Wang, W.-L.; Wu, Q.-Y.; Huang, N.; Wang, T.; Hu, H.-Y. Synergistic effect between UV and chlorine (UV/chlorine) on the degradation of carbamazepine: Influence factors and radical species. Water Res. 2016, 98, 190–198. [Google Scholar] [CrossRef]
  79. Dong, H.; Xu, Q.; Lian, L.; Li, Y.; Wang, S.; Li, C.; Guan, X. Degradation of Organic Contaminants in the Fe(II)/Peroxymonosulfate Process under Acidic Conditions: The Overlooked Rapid Oxidation Stage. Environ. Sci. Technol. 2021, 55, 15390–15399. [Google Scholar] [CrossRef]
  80. Golshan, M.; Kakavandi, B.; Ahmadi, M.; Azizi, M. Photocatalytic activation of peroxymonosulfate by TiO2 anchored on cupper ferrite (TiO2@CuFe2O4) into 2,4-D degradation: Process feasibility, mechanism and pathway. J. Hazard. Mater. 2018, 359, 325–337. [Google Scholar] [CrossRef]
  81. Huang, Y.; Lai, L.; Huang, W.; Zhou, H.; Li, J.; Liu, C.; Lai, B.; Li, N. Effective peroxymonosulfate activation by natural molybdenite for enhanced atrazine degradation: Role of sulfur vacancy, degradation pathways and mechanism. J. Hazard. Mater. 2022, 435, 128899. [Google Scholar] [CrossRef] [PubMed]
  82. Wang, C.; Wang, Y.; Yu, Y.; Cui, X.; Yan, B.; Song, Y.; Li, N.; Chen, G.; Wang, S. Effect of phosphates on oxidative species generation and sulfamethoxazole degradation in a pig manure derived biochar activated peroxymonosulfate system. Sep. Purif. Technol. 2022, 295, 121255. [Google Scholar] [CrossRef]
  83. Duan, P.; Liu, X.; Liu, B.; Akram, M.; Li, Y.; Pan, J.; Yue, Q.; Gao, B.; Xu, X. Effect of phosphate on peroxymonosulfate activation: Accelerating generation of sulfate radical and underlying mechanism. Appl. Catal. B-Environ. 2021, 298, 120532. [Google Scholar] [CrossRef]
  84. Xiao, S.; Cheng, M.; Zhong, H.; Liu, Z.; Liu, Y.; Yang, X.; Liang, Q. Iron-mediated activation of persulfate and peroxymonosulfate in both homogeneous and heterogeneous ways: A review. Chem. Eng. J. 2020, 384, 123265. [Google Scholar] [CrossRef]
  85. Ao, X.; Liu, W. Degradation of sulfamethoxazole by medium pressure UV and oxidants: Peroxymonosulfate, persulfate, and hydrogen peroxide. Chem. Eng. J. 2017, 313, 629–637. [Google Scholar] [CrossRef]
Figure 1. Comparison of performance of different waveform currents on SMX degradation regarding (a) removal of SMX and (b) calculated rate constants (the relationship between the molar concentration of reactants and the rate of a chemical reaction) in different oxidation processes for SMX treatment. Reaction conditions: [SMX] = 2 mg/L, [PMS] = 1.0 mmol/L, [Fe(III)] = 0.6 mmol/L, pH = 3, [Na2SO4] = 20 mmol/L, voltage = 6 V, pulsed frequency = 1 kHz, duty cycle = 50%. Duty cycle refers to the percentage of time during the whole circle that the pulse is applied.
Figure 1. Comparison of performance of different waveform currents on SMX degradation regarding (a) removal of SMX and (b) calculated rate constants (the relationship between the molar concentration of reactants and the rate of a chemical reaction) in different oxidation processes for SMX treatment. Reaction conditions: [SMX] = 2 mg/L, [PMS] = 1.0 mmol/L, [Fe(III)] = 0.6 mmol/L, pH = 3, [Na2SO4] = 20 mmol/L, voltage = 6 V, pulsed frequency = 1 kHz, duty cycle = 50%. Duty cycle refers to the percentage of time during the whole circle that the pulse is applied.
Catalysts 14 00532 g001
Figure 2. Impact of different pH values (a) in SMX degradation using sinusoidal waveform currents with calculated rate constant kobs in (d). Impact of different voltages (b) in SMX degradation using sinusoidal waveform currents with calculated rate constant kobs in (e). Impact of different frequencies (c) in SMX degradation using sinusoidal waveform currents with calculated rate constant kobs in (f). Reaction conditions: [SMX] = 2 mg/L, voltage = 6 V, pulsed frequency = 1 kHz, [PMS] = 1.0 mmol/L, [Fe(III)] = 0.6 mmol/L, pH = 3, [Na2SO4] = 20 mmol/L.
Figure 2. Impact of different pH values (a) in SMX degradation using sinusoidal waveform currents with calculated rate constant kobs in (d). Impact of different voltages (b) in SMX degradation using sinusoidal waveform currents with calculated rate constant kobs in (e). Impact of different frequencies (c) in SMX degradation using sinusoidal waveform currents with calculated rate constant kobs in (f). Reaction conditions: [SMX] = 2 mg/L, voltage = 6 V, pulsed frequency = 1 kHz, [PMS] = 1.0 mmol/L, [Fe(III)] = 0.6 mmol/L, pH = 3, [Na2SO4] = 20 mmol/L.
Catalysts 14 00532 g002
Figure 3. Impact of different concentrations of Fe(III) in (a,c) in SMX degradation using sinusoidal waveform currents with calculated rate constant kobs in (b,d). Reaction conditions: [SMX] = 2 mg/L, voltage = 6 V, pulsed frequency = 1 kHz, [PMS] = 1.0 mmol/L, pH = 3, [Na2SO4] = 20 mmol/L.
Figure 3. Impact of different concentrations of Fe(III) in (a,c) in SMX degradation using sinusoidal waveform currents with calculated rate constant kobs in (b,d). Reaction conditions: [SMX] = 2 mg/L, voltage = 6 V, pulsed frequency = 1 kHz, [PMS] = 1.0 mmol/L, pH = 3, [Na2SO4] = 20 mmol/L.
Catalysts 14 00532 g003
Figure 4. The concentrations of dissolved Fe(II) and total iron ions (a) in SIN/PMS/Fe(III) processes. The concentration of PMS (b) during the experiment in SIN/PMS/Fe(III) system. Reaction conditions: [SMX] = 2 mg/L, voltage = 6 V, pulsed frequency = 1 kHz, [PMS] = 1.0 mmol/L, pH = 3, [Na2SO4] = 20 mmol/L.
Figure 4. The concentrations of dissolved Fe(II) and total iron ions (a) in SIN/PMS/Fe(III) processes. The concentration of PMS (b) during the experiment in SIN/PMS/Fe(III) system. Reaction conditions: [SMX] = 2 mg/L, voltage = 6 V, pulsed frequency = 1 kHz, [PMS] = 1.0 mmol/L, pH = 3, [Na2SO4] = 20 mmol/L.
Catalysts 14 00532 g004
Figure 5. Comparisons of the performance of different oxidation processes were conducted regarding the (a) removal of SMX and (b) calculated rate constants in different oxidation processes for SMX treatment. Reaction conditions: [SMX] = 2 mg/L, [PMS] = 1.0 mmol/L, [Fe(III)] = 0.6 mmol/L, pH = 3, [Na2SO4] = 20 mmol/L, voltage = 6 V, duty cycle = 50%, pulsed frequency = 1 kHz.
Figure 5. Comparisons of the performance of different oxidation processes were conducted regarding the (a) removal of SMX and (b) calculated rate constants in different oxidation processes for SMX treatment. Reaction conditions: [SMX] = 2 mg/L, [PMS] = 1.0 mmol/L, [Fe(III)] = 0.6 mmol/L, pH = 3, [Na2SO4] = 20 mmol/L, voltage = 6 V, duty cycle = 50%, pulsed frequency = 1 kHz.
Catalysts 14 00532 g005
Figure 6. EPR spectra for the detection of (a) OH and SO4•− and (b) 1O2 in the presence of DMPO and TEMP, respectively. Reaction conditions: [SMX] = 2 mg/L, voltage = 6 V, pulsed frequency = 1 kHz, [PMS] = 1.0 mmol/L, [Fe(III)] = 0.6 mmol/L, pH = 3, [Na2SO4] = 20 mmol/L.
Figure 6. EPR spectra for the detection of (a) OH and SO4•− and (b) 1O2 in the presence of DMPO and TEMP, respectively. Reaction conditions: [SMX] = 2 mg/L, voltage = 6 V, pulsed frequency = 1 kHz, [PMS] = 1.0 mmol/L, [Fe(III)] = 0.6 mmol/L, pH = 3, [Na2SO4] = 20 mmol/L.
Catalysts 14 00532 g006
Figure 7. Quenching effects of different scavengers (a) on SMX degradation in SIN/PMS/Fe(III) system. Probe product 7-HC for OH (b) in SIN/PMS/Fe(III) system. Probe product p-HBA for SO4•− and PMSO2 for Fe(IV) (c) in SIN/PMS/Fe(III) system. Probe product MF and DF for superoxide radical (d) in SIN/PMS/Fe(III) system. Reaction conditions: [SMX] = 2 mg/L, voltage = 6 V, pulsed frequency = 1 kHz, [PMS] = 1.0 mmol/L, [Fe(III)] = 0.6 mmol/L, pH = 3, [Na2SO4] = 20 mmol/L.
Figure 7. Quenching effects of different scavengers (a) on SMX degradation in SIN/PMS/Fe(III) system. Probe product 7-HC for OH (b) in SIN/PMS/Fe(III) system. Probe product p-HBA for SO4•− and PMSO2 for Fe(IV) (c) in SIN/PMS/Fe(III) system. Probe product MF and DF for superoxide radical (d) in SIN/PMS/Fe(III) system. Reaction conditions: [SMX] = 2 mg/L, voltage = 6 V, pulsed frequency = 1 kHz, [PMS] = 1.0 mmol/L, [Fe(III)] = 0.6 mmol/L, pH = 3, [Na2SO4] = 20 mmol/L.
Catalysts 14 00532 g007
Figure 8. Proposed degradation pathways of SMX in SIN/PMS/Fe(III) system.
Figure 8. Proposed degradation pathways of SMX in SIN/PMS/Fe(III) system.
Catalysts 14 00532 g008
Figure 9. Toxicity evaluation results of (a) acute toxicity LD50, (b) developmental toxicity, (c) mutagenicity, and (d) bioaccumulation factor of SMX and SMX byproducts in SIN/PMS/Fe(III) process obtained using the Toxicity Estimation Software Tool (T.E.S.T.).
Figure 9. Toxicity evaluation results of (a) acute toxicity LD50, (b) developmental toxicity, (c) mutagenicity, and (d) bioaccumulation factor of SMX and SMX byproducts in SIN/PMS/Fe(III) process obtained using the Toxicity Estimation Software Tool (T.E.S.T.).
Catalysts 14 00532 g009
Figure 10. Influence of different anions including (a) Cl, (b) NO3, (c) H2PO4, and (d) humic acid (HA) and (e) different water backgrounds for SMX degradation in the PE/PMS/Fe(III) system. (f) SMX degradation in five consecutive runs to examine the reusability of the SIN/PMS/Fe(III) system. Reaction conditions: [SMX] = 2 mg/L, voltage = 6 V, pulsed frequency = 1 kHz, [PMS] = 1.0 mmol/L, [Fe(III)] = 0.6 mmol/L, pH = 3, [Na2SO4] = 20 mmol/L.
Figure 10. Influence of different anions including (a) Cl, (b) NO3, (c) H2PO4, and (d) humic acid (HA) and (e) different water backgrounds for SMX degradation in the PE/PMS/Fe(III) system. (f) SMX degradation in five consecutive runs to examine the reusability of the SIN/PMS/Fe(III) system. Reaction conditions: [SMX] = 2 mg/L, voltage = 6 V, pulsed frequency = 1 kHz, [PMS] = 1.0 mmol/L, [Fe(III)] = 0.6 mmol/L, pH = 3, [Na2SO4] = 20 mmol/L.
Catalysts 14 00532 g010
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Fang, J.; Wang, Y.; Wang, J.; Zhang, I.Y.; Huang, R. Application of Different Waveforms of Pulsed Current in the Classical Electro-Cocatalytic Process for Effective Removal of Sulfamethoxazole: Oxidation Mechanisms. Catalysts 2024, 14, 532. https://doi.org/10.3390/catal14080532

AMA Style

Fang J, Wang Y, Wang J, Zhang IY, Huang R. Application of Different Waveforms of Pulsed Current in the Classical Electro-Cocatalytic Process for Effective Removal of Sulfamethoxazole: Oxidation Mechanisms. Catalysts. 2024; 14(8):532. https://doi.org/10.3390/catal14080532

Chicago/Turabian Style

Fang, Jingkai, Yongjian Wang, Jiahao Wang, Igor Ying Zhang, and Rongfu Huang. 2024. "Application of Different Waveforms of Pulsed Current in the Classical Electro-Cocatalytic Process for Effective Removal of Sulfamethoxazole: Oxidation Mechanisms" Catalysts 14, no. 8: 532. https://doi.org/10.3390/catal14080532

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