Next Article in Journal
Optimizing Sampling Strategies for Estimating Riverine Nutrient Loads in the Yiluo River Watershed, China
Previous Article in Journal
Assessing the Impacts of Changing Connectivity of Hydropower Dams on the Distribution of Fish Species in the 3S Rivers, a Tributary of the Lower Mekong
Previous Article in Special Issue
Electro-Assisted Fe3+/Persulfate System for the Degradation of Bezafibrate in Water: Kinetics, Degradation Mechanism, and Toxicity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

CuFeS2/MXene-Modified Polyvinylidene Fluoride Membrane for Antibiotics Removal through Peroxymonosulfate Activation

1
Key Laboratory of Drinking Water Safety and Distribution Technology of Zhejiang Province, College of Civil Engineering and Architecture, Zhejiang University, Hangzhou 310058, China
2
Future Water Laboratory, Innovation Center of Yangtze River Delta, Zhejiang University, Jiaxing 314000, China
*
Author to whom correspondence should be addressed.
Water 2024, 16(11), 1504; https://doi.org/10.3390/w16111504
Submission received: 22 April 2024 / Revised: 18 May 2024 / Accepted: 21 May 2024 / Published: 24 May 2024

Abstract

:
In this research, the CuFeS2/MXene-modified polyvinylidene fluoride (PVDF) membrane was prepared to activate peroxymonosulfate (PMS) to remove moxifloxacin (MOX) and its morphology; surface functional groups and hydrophilicity were also studied. The parameters of the catalytic membrane/PMS system were optimized, with an optimal loading of 4 mg/cm2 and a PMS dosage of 0.20 mM. High filtration pressure, alkaline conditions, and impurities in water could inhibit MOX removal. After continuous filtration, the removal efficiency of MOX using the catalytic membrane/PMS system and PVDF membrane was 68.2% and 9.9%, respectively. Batch filtration could remove 87.8% MOX by the extra 10 min contact time between the catalytic membrane and solution. During the filtration process, CuFeS2/MXene on the surface of the catalytic membrane activated PMS to produce SO4•−, HO, and 1O2, and MOX was removed through adsorption and degradation. Taking humic acid (HA) as the model foulant, reversible fouling resistance in the catalytic membrane/PMS system was 22.8% of the PVDF membrane. The catalytic membrane/PMS system weakened the formation of the cake layer by oxidizing HA into smaller pollutants and followed the intermediate blocking cake filtration model. The novelty of this research was to develop a CuFeS2/MXene–PVDF membrane-activated PMS system and explore its application in antibiotics removal.

Graphical Abstract

1. Introduction

The polyvinylidene fluoride (PVDF) membrane had excellent mechanical strength, chemical stability, and thermal stability. Its ultrafiltration process was widely used in drinking water production, wastewater treatment, reverse osmosis pretreatment, and so on [1]. However, membrane separation technology was only a physical process [2], and membrane fouling easily occurred due to the hydrophobicity and low surface energy of the PVDF membrane [3]. Typically, ultrafiltration membranes were not effective in removing antibiotics. PVDF ultrafiltration membranes removed 13.3% of sulfamethoxazole and 11.9% of tetracycline by filtering artificial water containing different antibiotics (50 μg/L), respectively [4].
Catalytic membranes were prepared by modifying the membranes with catalysts, which could couple with advanced oxidation technology. Combining oxidative degradation and membrane separation, catalytic membranes received attention for the removal of antibiotics. Alpatova et al. [5] immobilized Fe2O3 on a PVDF membrane in order to catalyze H2O2 to remove organic contaminants, which was beneficial for the reuse of heterogeneous catalysis. Li et al. [6] prepared the ZnO@Ti3C2Tx membrane to generate HO and O2 under visible light, showing a good separation performance for six cycles with RhB rejection > 88.48%. Zhou et al. [7] prepared the D-UiO-66/graphite/PVDF membrane, producing HO and O2 at a low current density (0.01 mA/cm2), with the removal efficiencies of four antibiotics more than 96.6%. Yue et al. [8] activated PMS with the Co2+/MXene membrane to effectively remove tetracycline (>97%) within 5 min and concluded that filtration enhanced the contact between PMS and Co2+. Heterogeneous catalysts activated persulfate to produce SO4•− with excellent oxidation properties, and the reaction system did not consume energy. The choice of catalytic materials was crucial for catalytic membranes to activate persulfate.
Metal oxides, zero-valent metals and the metal–organic framework were commonly used to activate persulfate, accompanied by changes in the valence state of metal elements. Polymetallic catalysts were effective at activating persulfate because the coupling reactions between different metals could increase the production of SO4•− [9]. Metal sulfides exhibited high activity in catalytic reactions, and S2- could promote the redox circulation of multi-valence metals in persulfate activation reactions [10]. Chalcopyrite (CuFeS2) was a kind of bimetallic mineral with abundant reserves and a wide distribution, which had potential applications in catalytic reactions. Cu+ and Fe2+ played a bimetallic synergy in PMS activation, and the reduction in S2− promoted the regeneration of the lower valence metal. Due to the efficient electron transfer properties, CuFeS2 was used to thermally activate H2O2 to degrade dye Rhodamine B [11]. Based on experimental data and theoretical calculations, Huang et al. [12] concluded that the Cu(I)/Fe(III) redox couple in CuFeS2 was crucial to activating peroxydisulfate. Nie et al. [13] reported that CuFeS2 could be used to activate PMS to degrade bisphenol A, with a degradation efficiency of 99.7% and a mineralization efficiency of 75% within 20 min. Zhou et al. [14] activated peroxymonosulfate (PMS) with 1 g/L of CuFeS2 and 95.7% sulfisoxazole, which was degraded within 5 min. CuFeS2/PMS could still degrade 83.1% bisphenol S in the five-cycle experiment [15], indicating its excellent reusability. These studies indicate that CuFeS2 is an efficient catalyst for PMS activation.
However, heterogeneous metal catalysts still had particle aggregation and metal ion leaching, which could be alleviated by combining with carbon-based materials. C3N4–Fe–rGO prevented the acid leaching and agglomeration of Fe, which is conducive to maintaining the activity of the reaction site [16]. Biochar could improve the stability and dispersion of catalysts [17]. Carbon nanotubes could promote adsorption and electron transfer among the catalysts, PMS, and pollutants [18]. MXene could immobilize metal ions due to the surface with rich functional groups. Xia et al. [19] found that CoOOH was well dispersed on MXene and concluded that MXene, as a supporting material, could enhance the catalytic activity. For Fe3O4@MXene, Ti on the MXene surface was helpful for the acceleration of the Fe(II)/Fe(III) cycle, maintaining the activation of sodium percarbonate [20]. Xie et al. [21] prepared FeS2@MXene for water electrolysis and concluded that MXene could reduce the agglomeration phenomenon of FeS2 and regulate the electron density. MXene could be used as the carrier of the transition metal catalyst and play a synergistic role in the activation reaction of persulfate. Therefore, it was hypothesized that the combination of CuFeS2 and MXene could effectively activate PMS. CuFeS2/MXene catalytic membranes could potentially be utilized for the removal of antibiotics in water treatment applications.
In addition, some research showed that catalytic membranes could play a role in mitigating the formation of membrane fouling. The g-C3N4 composite membrane prepared by Li et al. [22] exhibited a self-cleaning performance under visible light, and polyvinyl alcohol (PVA) coated on the membrane surface could enhance hydrophilicity. Zhao et al. [23] prepare ad FeOCl-functionalized ceramic membrane to activate peroxyacetic acid to generate HO and 1O2 as the main species, reducing the extent of pore blocking and cake layer formation. Bai et al. [24] loaded layered double oxides (MnAl-LMO, CuAl-LMO, and CoAl-LMO) onto ceramic membranes to activate persulfate, resulting in the production of SO4•−, HO, and 1O2, which could reduce membrane fouling and improve water quality.
The objective of this research was to synthesize the CuFeS2/MXene–PVDF membrane and investigate the antibiotic removal performance, its influences, and its mechanism using the catalytic membrane/PMS system. Due to the frequent detection in aquatic environments and potential risk to the ecosystem, moxifloxacin (MOX) was selected as the target pollutant in this research. MOX removal efficiency and the water flux of the membrane were investigated. Parameters of the catalytic membrane/PMS system were optimized, and the effects of operating conditions and the water matrix were studied. The mechanisms of MOX removal by the catalytic membrane/PMS system and the anti-fouling property of membranes were studied. Compared with other studies, a novel catalytic membrane was prepared to remove MOX through the synergistic activation of PMS and membrane rejection performance. This research provided the theoretical reference for antibiotics removal using PVDF membrane-coupled advanced oxidation technology.

2. Materials and Methods

2.1. Materials

Moxifloxacin (MOX, 99%), PMS (KHSO5·0.5KHSO4·0.5K2SO4), CuCl, phosphoric acid (H3PO4), acetonitrile (CH3CN), methyl alcohol (MeOH) and humic acid (HA) were purchased from Aladdin, Shanghai, China. LiF, FeCl3·6H2O, polyethylene glycol (PEG), and glutaraldehyde (GA) were obtained from Macklin, Shanghai, China. HCl, NaOH, NaHCO3, Na2S2O3·5H2O, (NH2)2CS, anhydrous ethanol, tertiary butyl alcohol (TBA), and L-histidine (L-his) were acquired from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China). Ti3AlC2 (MAX) powder (400 mesh) was provided by HWRK Chemical, Beijing, China. H3PO4 and CH3CN were of the HPLC grade, while other chemicals were analytical grade except for MOX. Ultrapure water was used throughout the experiments. The PVDF membrane (80 mm, 0.1 µm) was provided by Longjin Membrane Technology Co., Ltd. (Nantong, China). In total, 50 mg/L of the MOX stock solution was prepared by dissolving 50 mg of MOX in 1 L of ultrapure water. The MOX stock solution was diluted to 2 mg/L as the initial concentration in the experiments.

2.2. Preparation of CuFeS2/MXene–PVDF Membrane

In this research, CuFeS2/MXene was synthesized by the hydrothermal method (Note S1). CuFeS2/MXene was ultrasonically dispersed in 50 mL of ultrapure water and deposited onto a PVDF membrane by vacuum filtration. Then, 50 mL of PEG (20 wt%) and GA (20 wt%) solutions were successively poured onto the membrane and pressurized with 0.05 MPa N2. By changing the amount of CuFeS2/MXene, the catalytic membrane with different catalyst loadings (2~4 mg/cm2 CuFeS2/MXene) could be prepared.

2.3. Membrane Characterization Method

The morphologies and microstructures of CuFeS2, the PVDF membrane and the CuFeS2/MXene–PVDF membrane were observed using a scanning electron microscope (SEM, FEG650, FEI, Mitchel Field, Long Island, NY, USA). The crystal structures of CuFeS2, CuFeS2/MXene, and CuFeS2/MXene–PVDF were characterized with an X-ray diffractometer (XRD-6000, Shimadzu, Kyoto, Japan). Elemental composition was detected by X-ray photoelectron spectroscopy (XPS, Escalab 250Xi, Thermo Fisher, Waltham, WA, USA). The functional groups on the surface of membranes were determined by Fourier transform infrared spectrometer (FTIR, Spectrum Two, Perkin-Elmer, Waltham, WA, USA). The contact angles of membranes were measured using a contact angle meter (DCA20, Dataphysics, Filderstadt, Germany).

2.4. Experimental Procedure

For the continuous filtration experiment, the catalytic membrane was fixed on the ultrafiltration cup to filter 2 L of MOX under N2 pressure, taking 1 mL sample from the filtrate and adding Na2S2O3 immediately to terminate the reaction. The effects on the MOX removal by the catalytic membrane/PMS system were studied, such as CuFeS2/MXene loading, PMS dosage, and filtration pressure. The effects of the water matrix on MOX removal were investigated by adjusting the initial pH with HCl or NaOH and adding NaHCO3 and HA as impurities in water. For batch filtration, the solution should be in contact with the catalytic membrane for 10 min before filtration. Quenching experiments and electron paramagnetic resonance techniques (EPR, EMXplus-9.5/12, Bruker, Karlsruhe, Germany) were conducted to speculate on the main reactive species. After the addition of the quencher (MeOH, TBA, or L-his), the catalytic membrane (loading = 4 mg/cm2, area = 38.5 cm2) activated 0.20 mM PMS to remove 250 mL of MOX under static conditions.
The catalytic membrane was completely submerged in water, stirring mechanically at different speeds for 1 h before being taken out and dried. The mass loss of modified materials could be evaluated by the weighing method. The membrane was used to filter 300 mL of HA (10 mg/L) at 0.02 MPa pressure, continuously weighing the mass of the filtered liquid. Based on filtration volume (V), filtration time (T), and water flux (J), equations of the relevant models were fitted and calculated, which could then be used to analyze the formation mechanism of membrane fouling.

2.5. Analysis Methods

The concentration of MOX was analyzed by high-performance liquid chromatography (HPLC, Agilent 1200 Series, Agilent Technologies, Santa Clara, CA, USA) with an XDB C-18 (5 μm, 4.6 × 150 mm) column and UV detector. The detection wavelength was 290 nm, the mobile phase was acetonitrile/water (25/75, V/V), and the flow rate was 1 mL/min.

2.6. Computational Methods

The density functional theory (DFT) was used to study the degradation pathway of MOX. The molecular formula of MOX was imported into GaussView 5.0 software (Gaussian Inc., Wallingford, CT, USA), and the DFT calculation was performed in Gaussian 09 software (Gaussian Inc., USA). The parameters were set corresponding to the B3LYP/6-31 + G (d, p) energy level theory, Natural Bond Orbitals, and aqueous phase model. The three-dimensional coordinates and Hirschfeld charges of each atom of MOX after geometric optimization were obtained. Then, the condensed Fukui function (CFF) was calculated according to Equations (1) and (2) [25]. Atom K with high values of f0 and f was susceptible to free radical attack and electrophilic attack, respectively.
Free   radical   attack                        f   k 0 = (   q N 1 k   q N + 1 k ) / 2
Electrophilic   attack                        f   k = (   q N 1 k   q N k )
where qk is the atom charge of atom K at the corresponding state; N, N − 1, and N + 1 indicate that the MOX molecule has a charge of 0, +1, and −1, respectively.
After the calculation file (.fch) of MOX was output from Gaussian 09, the Mayer analysis menu in Multiwfn software 3.8 (Tian Lu, China) was used to obtain the Mayer bonds of MOX. The Mayer bonds represented the stability of the chemical bond.
The water flux J (L m−2 h−1) of membranes could be calculated according to Equation (3) as follows:
J = V A T
where V (L) is the filter volume; A is the effective area membrane (3.85 × 10−3 m2); and T (h) represents the filtration time.
The catalytic performance of the membrane was evaluated by calculating the removal efficiency of MOX (Equation (4)).
R = 1 C p C 0
where Cp and C0 represent the MOX concentrations of the filtrate and influent, respectively.
HA (10 mg/L TOC) was selected as the model foulant, and the membrane resistance R1, reversible fouling resistance R2, and irreversible fouling resistance R3 of the membranes were obtained from Equations (5)–(7) as follows:
R 1 = Δ P μ J 0
R 2 = Δ P μ J 1 Δ P μ J 2
R 3 = Δ P μ J 2 Δ P μ J 0
where μ represents dynamic viscosity (8.9 × 10−4 Pa·s); ∆P represents filtration pressure; J0 (before filtering HA), J1 (after filtering HA) and J2 (after hydraulic cleaning) represents the water flux J (L m−2 h−1) of membranes, respectively.

3. Results and Discussion

3.1. Membrane Characterization

SEM could observe the surface morphology of CuFeS2/MXene, the PVDF membrane, and the catalytic membrane. As a type of MXene, Ti3C2Tx with a typical layered structure was prepared by etching Ti3AlC2 (Figure 1a). In Figure 1b, a large amount of granular CuFeS2 was loaded onto the surface of MXene. As shown in Figure 1c, the PVDF membrane was a loose porous network structure with uneven pore size. The size of CuFeS2/MXene particles was larger than PVDF membrane pores, indicating that CuFeS2/MXene could accumulate on the surface of the PVDF membrane. In Figure 1d, the presence of CuFeS2/MXene and the formation of pores were observed on the membrane surface.
The phase composition and structure were inspected by XRD, as shown in Figure 2a. According to the standard XRD data for CuFeS2 (JCPDS 83-0983), the diffraction signals at 2θ = 29.4°, 49.0°, 57.9°, and 79.5° corresponded to the (1 1 2), (2 0 4), (3 1 2), and (3 1 6) planes of CuFeS2. The strong diffraction signals and sharp peak shapes indicated that CuFeS2 was crystallographically complete, and the arrangement of the atoms inside was relatively regular. Xu et al. [26] used XRD to characterize CuFeS2 and observed similar characteristic peaks. The characteristic peaks of the composite at 18.2°, 39.0°, and 60.1° were correlated with the (0 0 4), (1 0 4), and (1 1 0) planes of MXene [19], indicating the successful synthesis of CuFeS2/MXene. For CuFeS2/MXene, the peak corresponding to the (0 0 2) plane was shifted to a lower angle compared to the standard XRD data for Ti3AlC2 (JCPDS 52-0875), suggesting that some particles were loaded onto the layered structure of MXene [20]. In addition, the low intensity of some peaks belonging to MXene may be due to the formation of CuFeS2 on the MXene surface. XPS was applied to analyze the elemental composition of CuFeS2, CuFeS2/MXene, and the CuFeS2/MXene–PVDF membrane, and the results are presented in Figure 2b–d. Cu, Fe, and S were obtained from CuFeS2, and Ti was derived from MXene. The results of SEM, XRD, and XPS indicate the successful synthesis of the CuFeS2/MXene–PVDF membrane.
The functional groups of membranes were analyzed by FTIR (Figure 3). The characteristic peak at 839 cm−1, which corresponded to the β phase of the PVDF membrane, was clear [27]. The absorbance of 2943 and 1402 cm−1 was due to the stretching vibration of -CH2, and the peaks at 1170, 1073, and 878 cm−1 were related to -CF2. Li et al. [28] obtained similar FTIR spectra when characterizing PVDF membranes. The FTIR spectrum of the catalytic membrane showed a broad band in the range of 3200~3500 cm−1 and a characteristic peak at 1730 cm−1, indicating that -OH and -CHO were introduced on the membrane surface. Strong signals of CuFeS2 in FTIR spectra [29] included 1147, 1120, 1040, 993 and 960 cm−1. Yu et al. [30] observed the FTIR characteristic peak of CuFeS2 at 570 cm−1. The catalytic membrane prepared in this research also showed strong absorbance at the corresponding wave number, indicating the presence of CuFeS2 on the surface of the membrane. According to the results of previous observations of MXene [31], 1630 and 655 cm−1 in Figure 3 correspond to -COOH and Ti-O, respectively.
The hydrophilicity of the surfaces of membranes could be quantified through the contact angles, and the results are shown in Figure 4. The PVDF membrane exhibited a contact angle of 131.9°, and it was presumed that its hydrophobic quality was attributed to the repulsion of water molecules by the C-F bond. The catalytic membrane had super hydrophilicity, with a contact angle of 3.2°. PEG and GA were added as adhesives during the preparation of the catalytic membrane, so hydrophilic groups such as -OH and -CHO were introduced. In addition, the zeta potentials of PVDF and the catalytic membrane were −2.31 mV and −0.72 mV at pH = 5.75 (initial pH of the filtered solution), respectively. When the pH value was 5.22, the zeta potential of the catalytic membrane was zero. Zhao et al. measured the isoelectric point of the PVDF membrane at pH = 3.4 and concluded that the detected zeta potential values of COF-LZU1, which modified the membranes slightly, increased with the enhancement of surface hydrophilicity [32].

3.2. Parameter Optimization of the Catalytic Membrane/PMS System

Catalytic membranes could degrade antibiotics through advanced oxidation techniques. Before the filtration experiment, the bonding strength between modified materials and the PVDF membrane was investigated by shear water flow (Table S1). Only when using vacuum filtration did CuFeS2/MXene easily fall off the PVDF membrane. However, after coating PEG and GA on the membrane surface, the loss of modified materials was significantly reduced. It was speculated that PEG and GA could effectively adhere CuFeS2/MXene to the surface of the PVDF membrane. At 0.02 MPa filtration pressure, the MOX removal efficiencies by different systems were investigated in continuous filtration (Figure 5a). The removal efficiency of MOX by the CuFeS2/MXene–PVDF membrane that activated the PMS system was 31.1%. Compared with other systems (removing 9.9%~13.4% MOX), it could be concluded that the catalytic membrane/PMS system promoted the removal of MOX through oxidative degradation. During the filtration process, CuFeS2/MXene on the surface of the catalytic membrane activated PMS and generated reactive species.
Using CuFeS2/MXene–PVDF membranes with different catalyst loadings to activate PMS, the removal efficiencies of MOX were 21.1%, 31.1%, and 48.2% (Figure 5b), respectively. The increase in loadings could promote MOX removal in the catalytic membrane/PMS system because more CuFeS2/MXene indicated that there were more reaction sites in the PMS activation. At 0.02 MPa filtration pressure, the water flux of the PVDF membrane was 630.7 L m−2 h−1. As the CuFeS2/MXene loading on the membrane surface increased from 1 to 4 mg/cm2, the water flux decreased from 496.8 to 344.8 L m−2 h−1, promoting PMS activation by increasing the contact time between PMS and the catalytic membrane. Zheng et al. [2] used a catalytic membrane to activate PMS to remove pollutants and concluded that the kobs values of RhB degradation increased with the increasing Cu-MOF-74 loadings. Figure 5c showed the removal efficiencies of MOX from different catalytic membrane/PMS systems, indicating that the catalytic performance of CuFeS2/MXene was better than that of CuFeS2. It was presumed that CuFeS2 and MXene played a synergistic role in activating PMS. The addition of PMS (0.20, 0.16, and 0.12 mM) resulted in MOX removal efficiencies of 68.2%, 54.1%, and 48.2% (Figure 5d), respectively. It showed that the increase in the PMS concentration was advantageous for MOX removal. With the progress of filtration, the reaction sites on the catalytic membrane surface may be covered by pollutants [33], which could hinder the reaction between CuFeS2/MXene and PMS. After selecting an appropriate loading (4 mg/cm2) and PMS dosage (0.20 mM), the following experiments were carried out. In this research, the removal efficiency of 2 L of MOX by the catalytic membrane/PMS system was 68.2%, which was higher than that of the PVDF membrane (9.9%). The slight removal of antibiotics by the PVDF membrane was attributed to solute adsorption into the membrane’s porous structure [34]. Wang et al. [35] activated PMS with the Ti catalytic membrane and TiO2/Ti catalytic membrane, and the removal efficiency of ciprofloxacin was 6.9% and 7.2%, respectively. It indicated that CuFeS2/MXene activated PMS effectively and promoted MOX removal through oxidative degradation.

3.3. Removal Efficiency under Varying Conditions

The effects of the operating conditions and water matrix on MOX removal by the catalytic membrane/PMS system were studied. Figure 6a shows the MOX removal at different filtration pressures. When the filtration pressure increased from 0.02 MPa to 0.08 MPa, the removal efficiency of MOX decreased from 68.2% to 18.5%. Low filtration pressure was conducive to the MOX removal by the catalytic membrane/PMS system. At 0.08 MPa, it was speculated that MOX was mainly removed by adsorption because the catalytic membrane could not effectively activate PMS within a short contact time. Under the different filtration pressures (0.08, 0.05, and 0.02 MPa), the water flux of the catalytic membrane was 1150.3, 548.3, and 344.8 L m−2 h−1, respectively. Filtration pressure affected the oxidative degradation effect by affecting the contact time between PMS and the catalytic membrane. Yue et al. [8] obtained a similar view that the degradation efficiency of antibiotics gradually decreased with the increase in pressure due to the short residual time of PMS on the catalytic membrane. In Figure 6b, the MOX removal efficiency of batch filtration increased by 19.6% compared with the original filtration method. It illustrated that the batch filtration was beneficial to the activation and the oxidative degradation of MOX.
By changing the initial pH value in the solution, the effect of a strong acid or alkali on MOX removal efficiency was investigated (Figure 7a). The initial pH of 2 mg/L of the MOX solution was 5.75, which was close to the theoretical value calculated based on the dissociation coefficient [36]. Filtering 1.5 L of the solution when the initial pH values were 3.10, 5.75, and 10.89, the removal efficiency of MOX was 83.4%, 82.5%, and 74.9%, respectively. This shows that the catalytic membrane/PMS system is more suitable for removing MOX under acidic conditions. pH could affect PMS activation by changing the fractions of different PMS species [37] and the surface charge of the catalyst. When the initial pH was 3.10, PMS almost existed in the form of HSO5, and the surface of CuFeS2/MXene was positively charged. It promoted the activation of PMS through the electrostatic interaction between CuFeS2/MXene and PMS. When the initial pH was 10.89, the main component of PMS was HSO52−, which repelled CuFeS2/MXene with a negative surface charge. In addition, SO4•− was easy to use to generate HO under alkaline conditions [38] (Equation (8)), and the oxidation performance and action time of SO4•− were greater than that of HO [39].
SO4•− + OH → HO + SO42−
Impurities in the water could also interfere with the degradation of MOX due to scavenging and adsorption competition. The effects of NOM and anions on the catalytic membrane/PMS system were investigated by selecting HA and HCO3, respectively, as shown in Figure 7b. When HA and HCO3 were used as impurities in water, the removal efficiencies of MOX were 80.7% and 77.3%, respectively. HCO3 may react with SO4•− and HO to form active substances with a lower oxidation capacity (Equations (9) and (10)) [15,40], which weakenes the degradation of MOX. Free radicals were more likely to attach to HA than pollutants and then react with functional groups in HA [41]. HA preempted the active sites on the catalyst surface through strong π-π stacking [42], so the interaction between CuFeS2/MXene and PMS could be blocked.
SO4•− + HCO3 → HCO3 + SO42−
HO + HCO3 → CO3•− + H2O

3.4. Mechanism of MOX Removal by Catalytic Membrane

When the solution was filtered by a catalytic membrane, MOX adsorption and PMS activation could occur. According to the FTIR spectral results, the catalytic membrane surface contained functional groups, such as -OH, -CHO, and -COOH, which could adsorb MOX. Zou et al. [43] concluded that the oxygen-containing functional groups in adsorbents were an important mechanism for adsorbing MOX due to the stronger H-bonding effect. CuFeS2/MXene loaded on the membrane surface could activate PMS to produce reactive species, and the activation mechanism was studied. Cu+ and Fe2+ on the CuFeS2 surface reacted with PMS to generate SO4•−, which was accompanied by the increase in the valence states of metals (Cu2+ and Fe3+). Then, SO4•− and HSO5 could further generate HO and 1O2 through their reaction with H2O [44]. The low valence sulfur [13], Cu+ [45], and MXene [46] contained in CuFeS2/MXene could promote the regeneration of Cu2+/Fe3+ into Cu+/Fe2+, respectively.
MeOH was used to eliminate HO and SO4•−; TBA was chosen as a quencher of HO, and L-His could quench 1O2. Using the catalytic membrane/PMS system to remove MOX under static conditions, Figure 8a shows the effects of quenchers on MOX removal. With the addition of TBA, L-his, or MeOH, the removal efficiency of MOX was 70.1%, 68.2% and 26.6%, respectively. It can be inferred that SO4•−, HO•, and 1O2 are beneficial for MOX removal. Using DMPO and TEMP as the capture of radicals, the characteristic signal peaks of DMPO-HO, DMPO-SO4•−, and TEMP-1O2 are observed in the EPR spectrum (Figure 8b). The quenching experiments and EPR results show that SO4•−, HO, and 1O2 play an important role in MOX degradation.
CFF was used to illustrate the regioselectivity of radicals attacking (Table S2), where f0 and f corresponded to radical attack and electrophilic attack, respectively. The values of f0 and f of the reactive sites were usually larger than other regions, such as the atoms (Num. 1, 2, 5, 13, 19, 21, and 29). In Figure 9a, different colors showed the numerical ranges of the Mayer bonds of MOX. Mayer bonds of N1-C14 and C22-O27 were 1.26 and 1.60, respectively, so it was difficult to remove N and O. The Mayer bonds of N1-C2, N1-C5, C10-C21, N19-C24, and C13-F29 were 0.62, 0.49, 0.78, 0.91, and 0.93, respectively, which were easily broken under the attack of active species. Therefore, the degradation of MOX may involve the cleavage of the nitrogen-containing heterocycle (NCH) (C1-C2 and C1-C5), the removal of the cyclopropyl group (C24-N19), decarboxylation (C10-C21) and defluorination (C13-F29). Similar MOX degradation pathways have been reported in other studies, such as g-C3N4/photocatalysis [47] and FeMnCoOx/PMS [48]. The mechanism of MOX removal by the catalytic membrane/PMS system included adsorption and degradation, as shown in Figure 9b. CuFeS2/MXene on the membrane surface was in contact with PMS, generating reactive species (SO4•−, HO and 1O2) for the degradation of MOX on the membrane surface and in the solution. On the one hand, the PVDF membrane served as an effective carrier for CuFeS2/MXene, alleviating the problems of catalyst loss and difficulties in recycling. On the other hand, CuFeS2/MXene could enhance MOX removal by the PVDF membrane through advanced oxidation.

3.5. Anti-Fouling Property

HA could produce membrane fouling through adhesion and blockage, which was selected as a common foulant to assess the anti-fouling performance of membranes. The contamination of the membrane surface was reversible and could be removed by hydraulic cleaning, while irreversible membrane fouling caused by pore blockage and adsorption required chemical reagents to restore water flux. Based on the water flux of membranes and Equations (5)–(7), the resistances in different membrane treatment systems were calculated, as shown in Figure 10. For the PVDF membrane, the oxidation of PMS itself could reduce the total resistance and reversible fouling resistance by 7.5% and 21.0%, respectively. However, PMS also oxidized HA into smaller pollutants, which was easier to enter the PVDF membrane’s pores, and irreversible fouling resistance increased by 13.2%. For the catalytic membrane, HA caused accumulation on the membrane surface, and reversible fouling resistance was even more serious. However, due to the small pore size of the catalytic membrane, HA could not easily enter the membrane pore, so irreversible fouling resistance was lower than that of the PVDF membrane. The calculated total resistance and reversible fouling resistance of the catalytic membrane/PMS system were 69.9% and 22.8% of the PVDF membrane, respectively. This indicates that membrane fouling could be effectively alleviated by the catalytic membrane/PMS system. The catalytic membrane could activate PMS efficiently, reducing the degree of membrane fouling. On the one hand, in situ, PMS activation occurred when the feed water passed through the catalytic layer of the membrane, which could enhance the repulsive interactions between pollutants and membranes according to the interface free energy calculations [24]. On the other hand, the active species produced by activating PMS could promote the degradation of pollutants with high molecular weight into smaller fractions and oxidize the hydrophobic matter into hydrophilic matter [7,49]. Although the loading of CuFeS2/MXene increased the inherent resistance of the catalytic membrane, the total resistance and reversible fouling resistance were reduced through PMS activation.
Hermia modeled and analyzed the filtering process (Equation (11)), and the different values of n corresponded to four single models [50]. The PVDF membrane and catalytic membrane were used to filter HA in order to study the formation mechanism of membrane fouling (Figure 11 and Table S3). The Cake filtration model (R2 = 0.9944) might explain the formation of PVDF membrane fouling compared with other models (R2 = 0.9276~0.9710). Thus, it could be assumed that HA formed a cake layer on the surface of the PVDF membrane.
d 2 t / d V 2 = k ( dt / dV ) n
For the catalytic membrane/PMS system, the fitting effects of the cake filtration model (R2 = 0.9885) and intermediate blocking model (R2 = 0.9755) were better than the other models. The fitting effects of single models were not satisfactory, but the intermediate blocking cake filtration model (Equation (12)) [51] resulted in a better fitting effect (R2 = 0.9991). It was shown how the cake layer tended to transform into intermediate blocking, and Simon et al. [52] also found a similar conclusion using caustic cleaning to alleviate membrane fouling. It is speculated that the catalytic membrane could activate PMS to produce reactive species, and HA can be oxidized into smaller pollutants to enter the pores. The catalytic membrane/PMS system alleviated membrane fouling by oxidizing HA and weakening the formation of the cake layer, as follows:
1 K i ln ( 1 + K i ( 1 + 2 K c J 0 2 t 1 ) K c J 0 ) = y
where J0 is the initial water flux; y is the filtration volume of the membrane per unit area; and Ki (m−1) and Kc (s/m2) are the constants of the intermediate blocking model and cake filtration model, respectively.

4. Conclusions

The CuFeS2/MXene–PVDF membrane was prepared to active PMS and remove MOX, and it was characterized by the measurement of SEM, FTIR, and contact angles. CuFeS2/MXene accumulated on the surface of the PVDF membrane, leading to the formation of new pores. The coating of PEG and GA enhanced the bonding strength between CuFeS2/MXene and the PVDF membrane surface and also introduced hydrophilic functional groups, such as -OH and -CHO. The contact angles of the PVDF membrane and catalytic membrane were 131.9° and 3.2°, respectively, indicating superhydrophilicity for the latter. Optimal parameters for the catalytic membrane/PMS system included loading 4 mg/cm2 of CuFeS2/MXene onto the PVDF membrane and adding 0.20 mM of PMS. In the continuous filtration of a 2 L solution, the catalytic membrane/PMS system exhibited a 58.3% higher removal efficiency of MOX compared to the PVDF membrane alone. Low filtration pressure and batch filtration could promote PMS activation by enhancing the contact reaction between the catalytic membrane and solution. However, OH, HCO3, and HA could inhibit MOX removal by affecting the generation and transformation of free radicals. Besides adsorption, the catalytic membrane also activated PMS to generate SO4•−, HO, and 1O2, which could degrade MOX through the removal of the NCH and cyclopropyl group, decarboxylation, and defluorination. Based on the fouling resistance calculation and the intermediate blocking cake filtration model (R2 = 0.9991), the catalytic membrane/PMS system could reduce reversible membrane fouling and weaken the formation of the cake layer. These findings indicate the potential application of catalytic membrane-activated persulfate technology in the field of water treatment.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/w16111504/s1, Note S1: Synthesis of CuFeS2/MXene; Table S1: The mass loss of modified materials; Table S2: Data related to the Fukui function; Table S3: Single model and its correlation coefficient.

Author Contributions

Methodology, writing—original draft preparation, D.Z.; data curation, K.L.; supervision, writing—review and editing, L.F.; investigation, H.C. All authors have read and agreed to the published version of the manuscript.

Funding

The work was supported by the National Key R&D Program of China (No. 2023YFC3208203, No. 2023YFF0614500), and the Zhejiang Provincial Natural Science Foundation of China (No. LY24E080005).

Data Availability Statement

All data generated or analyzed during this study are included in this published article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Ji, J.; Liu, F.; Hashim, N.A.; Abed, M.R.M.; Li, K. Poly (vinylidene fluoride) (PVDF) membranes for fluid separation. React. Funct. Polym. 2015, 86, 134–153. [Google Scholar] [CrossRef]
  2. Zheng, H.; Zhou, Y.; Wang, D.; Zhu, M.; Sun, X.; Jiang, S.; Fan, Y.; Zhang, D.; Zhang, L. Surface-functionalized PVDF membranes by facile synthetic Cu-MOF-74 for enhanced contaminant degradation and antifouling performance. Colloids Surf. A. 2022, 651, 129640. [Google Scholar] [CrossRef]
  3. Li, J.; Sun, J.; Ren, L.; Lei, T.; Li, J.; Jin, J.; Luo, S.; Qin, S.; Gao, C. Properties and preparation of TiO2-HAP@PVDF composite ultrafiltration membranes. Polym. Compos. 2023, 44, 7499–7509. [Google Scholar] [CrossRef]
  4. Liu, H.; Zhang, H.; Dong, X.; Wu, C.; Lichtfouse, E. Removal of antibiotics from black water by a membrane filtration-visible light photocatalytic system. J. Water Process Eng. 2023, 53, 103605. [Google Scholar] [CrossRef]
  5. Alpatova, A.; Meshref, M.; McPhedran, K.N.; Gamal El-Din, M. Composite polyvinylidene fluoride (PVDF) membrane impregnated with Fe2O3 nanoparticles and multiwalled carbon nanotubes for catalytic degradation of organic contaminants. J. Membr. Sci. 2015, 490, 227–235. [Google Scholar] [CrossRef]
  6. Li, Y.; Luo, H.; Ji, W.; Li, S.; Nian, P.; Xu, N.; Ye, N.; Wei, Y. Visible-light-driven photocatalytic ZnO@Ti3C2Tx MXene nanofiltration membranes for enhanced organic dyes removal. Sep. Purif. Technol. 2023, 323, 124420. [Google Scholar] [CrossRef]
  7. Zhou, S.; Zhu, J.; Wang, Z.; Yang, Z.; Yang, W.; Yin, Z. Defective MOFs-based electrocatalytic self-cleaning membrane for wastewater reclamation: Enhanced antibiotics removal, membrane fouling control and mechanisms. Water Res. 2022, 220, 118635. [Google Scholar] [CrossRef] [PubMed]
  8. Yue, R.; Sun, X. A Self-cleaning, catalytic titanium carbide (MXene) membrane for efficient tetracycline degradation through peroxymonosulfate activation: Performance evaluation and mechanism study. Sep. Purif. Technol. 2021, 279, 119796. [Google Scholar] [CrossRef]
  9. Li, X.; Jie, B.; Lin, H.; Deng, Z.; Qian, J.; Yang, Y.; Zhang, X. Application of sulfate radicals-based advanced oxidation technology in degradation of trace organic contaminants (TrOCs): Recent advances and prospects. J. Environ. Manag. 2022, 308, 114664. [Google Scholar] [CrossRef]
  10. Liu, Y.; Du, N.; Liu, X.; Yao, D.; Wu, D.; Li, Z.; Mao, S. ZnS-embedded porous carbon for peroxydisulfate activation: Enhanced electron transfer for bisphenol A degradation. Environ. Sci. Ecotechnol. 2024, 19, 100338. [Google Scholar] [CrossRef]
  11. Yang, J.; Jia, K.; Lu, S.; Cao, Y.; Boczkaj, G.; Wang, C. Thermally activated natural chalcopyrite for Fenton-like degradation of Rhodamine B: Catalyst characterization, performance evaluation, and catalytic mechanism. J. Environ. Chem. Eng. 2024, 12, 111469. [Google Scholar] [CrossRef]
  12. Huang, J.; Zhou, Y.; Deng, S.; Shangguan, Y.; Wang, R.; Ge, Q.; Feng, X.; Yang, Z.; Ji, Y.; Fan, T.; et al. Photo-assisted reductive cleavage and catalytic hydrolysis-mediated persulfate activation by mixed redox-couple-involved CuFeS2 for efficient trichloroethylene oxidation in groundwater. Water Res. 2022, 222, 118885. [Google Scholar] [CrossRef] [PubMed]
  13. Nie, W.; Mao, Q.; Ding, Y.; Hu, Y.; Tang, H. Highly efficient catalysis of chalcopyrite with surface bonded ferrous species for activation of peroxymonosulfate toward degradation of bisphenol A: A mechanism study. J. Hazard. Mater. 2019, 364, 59–68. [Google Scholar] [CrossRef]
  14. Zhou, W.; Li, Y.; Zhang, M.; Ying, G.; Feng, Y. Highly Efficient Degradation of Sulfisoxazole by Natural Chalcopyrite-Activated Peroxymonosulfate: Reactive Species and Effects of Water Matrices. Water-Sui 2022, 14, 3450. [Google Scholar] [CrossRef]
  15. Peng, J.; Zhou, H.; Liu, W.; Ao, Z.; Ji, H.; Liu, Y.; Su, S.; Yao, G.; Lai, B. Insights into heterogeneous catalytic activation of peroxymonosulfate by natural chalcopyrite: pH-dependent radical generation, degradation pathway and mechanism. Chem. Eng. J. 2020, 397, 125387. [Google Scholar] [CrossRef]
  16. Zuo, S.; Jin, X.; Wang, X.; Lu, Y.; Zhu, Q.; Wang, J.; Liu, W.; Du, Y.; Wang, J. Sandwich structure stabilized atomic Fe catalyst for highly efficient Fenton-like reaction at all pH values. Appl. Catal. B Environ. 2021, 282, 119551. [Google Scholar] [CrossRef]
  17. Chen, X.; Li, H.; Lai, L.; Zhang, Y.; Chen, Y.; Li, X.; Liu, B.; Wang, H. Peroxymonosulfate activation using MnFe2O4 modified biochar for organic pollutants degradation: Performance and mechanisms. Sep. Purif. Technol. 2023, 308, 122886. [Google Scholar] [CrossRef]
  18. Zhang, J.; Ma, Y.; Sun, Y.; Zhu, Y.; Wang, L.; Lin, F.; Ma, Y.; Ji, W.; Li, Y.; Wang, L. Enhancing deep mineralization of refractory benzotriazole via carbon nanotubes-intercalated cobalt copper bimetallic oxide nanosheets activated peroxymonosulfate process: Mechanism, degradation pathway and toxicity. J. Colloid Interface Sci. 2022, 628, 448–462. [Google Scholar] [CrossRef]
  19. Xia, X.; Deng, L.; Yang, L.; Shi, Z. Facile synthesis of CoOOH@MXene to activate peroxymonosulfate for efficient degradation of sulfamethoxazole: Performance and mechanism investigation. Environ. Sci. Pollut. Res. 2022, 29, 52995–53008. [Google Scholar] [CrossRef]
  20. Xu, X.; Zhou, Y.; Li, S.; Lin, Y.; Zhang, S.; Guan, Q. Activation of sodium percarbonate with Fe3O4@MXene composite for the efficient removal of bisphenol A. J. Environ. Chem. Eng. 2022, 10, 108702. [Google Scholar] [CrossRef]
  21. Xie, Y.; Yu, H.; Deng, L.; Amin, R.S.; Yu, D.; Fetohi, A.E.; Maximov, M.Y.; Li, L.; El-Khatib, K.M.; Peng, S. Anchoring stable FeS2 nanoparticles on MXene nanosheets via interface engineering for efficient water splitting. Inorg. Chem. Front. 2022, 9, 662–669. [Google Scholar] [CrossRef]
  22. Li, R.; Ren, Y.; Zhao, P.; Wang, J.; Liu, J.; Zhang, Y. Graphitic carbon nitride (g-C3N4) nanosheets functionalized composite membrane with self-cleaning and antibacterial performance. J. Hazard. Mater. 2019, 365, 606–614. [Google Scholar] [CrossRef] [PubMed]
  23. Zhao, Y.; Zhao, Y.; Yu, X. Peracetic acid integrated catalytic ceramic membrane filtration for enhanced membrane fouling control: Performance evaluation and mechanism analysis. Water Res. 2022, 220, 118710. [Google Scholar] [CrossRef] [PubMed]
  24. Bai, Z.; Gao, S.; Yu, H.; Liu, X.; Tian, J. Layered metal oxides loaded ceramic membrane activating peroxymonosulfate for mitigation of NOM membrane fouling. Water Res. 2022, 222, 118928. [Google Scholar] [CrossRef] [PubMed]
  25. Shi, J.; Zhang, B.; Wang, W.; Zhang, W.; Du, P.; Liu, W.; Xing, X.; Ding, D.; Lv, G.; Lv, Q.; et al. In situ produced hydrogen peroxide by biosynthesized palladium nanoparticles and natural clay mineral for highly-efficient carbamazepine degradation. Chem. Eng. J. 2021, 426, 131567. [Google Scholar] [CrossRef]
  26. Xu, X.; Tang, D.; Cai, J.; Xi, B.; Zhang, Y.; Pi, L.; Mao, X. Heterogeneous activation of peroxymonocarbonate by chalcopyrite (CuFeS2) for efficient degradation of 2,4-dichlorophenol in simulated groundwater. Appl. Catal. B Environ. 2019, 251, 273–282. [Google Scholar] [CrossRef]
  27. Huang, F.; Li, Q.; Ji, G.; Tu, J.; Ding, N.; Qu, Q.; Liu, G. Oil/water separation using a lauric acid-modified, superhydrophobic cellulose composite membrane. Mater. Chem. Phys. 2021, 266, 124493. [Google Scholar] [CrossRef]
  28. Li, N.; Xiao, C.; An, S.; Hu, X. Preparation and properties of PVDF/PVA hollow fiber membranes. Desalination 2010, 250, 530–537. [Google Scholar] [CrossRef]
  29. Reyes-Bozo, L.; Escudey, M.; Vyhmeister, E.; Higueras, P.; Godoy-Faúndez, A.; Salazar, J.L.; Valdés-González, H.; Wolf-Sepúlveda, G.; Herrera-Urbina, R. Adsorption of biosolids and their main components on chalcopyrite, molybdenite and pyrite: Zeta potential and FTIR spectroscopy studies. Miner. Eng. 2015, 78, 128–135. [Google Scholar] [CrossRef]
  30. Yu, H.; Xu, J.; Hu, Y.; Zhang, H.; Zhang, C.; Qiu, C.; Wang, X.; Liu, B.; Wei, L.; Li, J. Synthesis and characterization of CuFeS2 and Se doped CuFeS2−xSex nanoparticles. J. Mater. Sci.-Mater. Electron. 2019, 30, 12269–12274. [Google Scholar] [CrossRef]
  31. Jun, B.; Kim, S.; Rho, H.; Park, C.M.; Yoon, Y. Ultrasound-assisted Ti3C2Tx MXene adsorption of dyes: Removal performance and mechanism analyses via dynamic light scattering. Chemosphere 2020, 254, 126827. [Google Scholar] [CrossRef] [PubMed]
  32. Zhao, Y.; Wang, C.; Ren, L.; Zhang, B.; Shao, J. Facile preparation of COF-LZU1 modified membrane via electrostatic spraying for emerging contaminants treatment in direct contact membrane distillation. Sep. Purif. Technol. 2024, 337, 126411. [Google Scholar] [CrossRef]
  33. Xiang, S.; Dong, H.; Li, Y.; Xiao, J.; Dong, Q.; Hou, X.; Chu, D. Novel flower-like Fe-Mo composite for peroxydisulfate activation toward efficient degradation of carbamazepine. Sep. Purif. Technol. 2023, 305, 122487. [Google Scholar] [CrossRef]
  34. Utomo, D.P.; Kusworo, T.D.; Kumoro, A.C.; Othman, M.H.D. Developing a versatile and resilient PVDF/CeO2@GO-COOH photocatalytic membrane for efficient treatment of antibiotic-contaminated wastewater. J. Water Process Eng. 2023, 56, 104353. [Google Scholar] [CrossRef]
  35. Wang, L.; Wu, W.; Lin, L.; Wang, Z.; Hou, X.; Wang, L.; Zhang, F.; Li, Y.; Xie, H. Activation of peroxymonosulfate for ciprofloxacin degradation by a novel oxygen-deficient CoTiO3/TiO2NTs/Ti composite catalytic membrane: Electron transfer pathway and underlying mechanism. Sep. Purif. Technol. 2024, 330, 125511. [Google Scholar] [CrossRef]
  36. Neves, P.; Leite, A.; Rangel, M.; de Castro, B.; Gameiro, P. Influence of structural factors on the enhanced activity of moxifloxacin: A fluorescence and EPR spectroscopic study. Anal. Bioanal. Chem. 2007, 387, 1543–1552. [Google Scholar] [CrossRef]
  37. Tan, C.; Gao, N.; Fu, D.; Deng, J.; Deng, L. Efficient degradation of paracetamol with nanoscaled magnetic CoFe2O4 and MnFe2O4 as a heterogeneous catalyst of peroxymonosulfate. Sep. Purif. Technol. 2017, 175, 47–57. [Google Scholar] [CrossRef]
  38. Zhang, T.; Chen, Y.; Leiknes, T. Oxidation of refractory benzothiazoles with PMS/CuFe2O4: Kinetics and transformation intermediates. Environ. Sci. Technol. 2016, 50, 5864–5873. [Google Scholar] [CrossRef] [PubMed]
  39. Zhang, X.; Gan, X.; Cao, S.; Shang, J.; Cheng, X. Efficient Removal of Rhodamine B in Wastewater via Activation of Persulfate by MnO2 with Different Morphologies. Water-Sui 2023, 15, 735. [Google Scholar] [CrossRef]
  40. Li, J.; Yan, J.; Yao, G.; Zhang, Y.; Li, X.; Lai, B. Improving the degradation of atrazine in the three-dimensional (3D) electrochemical process using CuFe2O4 as both particle electrode and catalyst for persulfate activation. Chem. Eng. J. 2019, 361, 1317–1332. [Google Scholar] [CrossRef]
  41. Xu, P.; Wang, P.; Li, X.; Wei, R.; Wang, X.; Yang, C.; Shen, T.; Zheng, T.; Zhang, G. Efficient peroxymonosulfate activation by CuO-Fe2O3/MXene composite for atrazine degradation: Performance, coexisting matter influence and mechanism. Chem. Eng. J. 2022, 440, 135863. [Google Scholar] [CrossRef]
  42. Liu, L.; Mi, H.; Zhang, M.; Sun, F.; Zhan, R.; Zhao, H.; He, S.; Zhou, L. Efficient moxifloxacin degradation by CoFe2O4 magnetic nanoparticles activated peroxymonosulfate: Kinetics, pathways and mechanisms. Chem. Eng. J. 2021, 407, 127201. [Google Scholar] [CrossRef]
  43. Zou, M.; Tian, W.; Chu, M.; Gao, H.; Zhang, D. Biochar composite derived from cellulase hydrolysis apple branch for quinolone antibiotics enhanced removal: Precursor pyrolysis performance, functional group introduction and adsorption mechanisms. Environ. Pollut. 2022, 313, 120104. [Google Scholar] [CrossRef] [PubMed]
  44. Wang, Y.; Yan, P.; Dou, X.; Liu, C.; Zhang, Y.; Song, Z.; Chen, Z.; Xu, B.; Qi, F. Degradation of benzophenone-4 by peroxymonosulfate activated with microwave synthesized well-distributed CuBi2O4 microspheres: Theoretical calculation of degradation mechanism. Appl. Catal. B Environ. 2021, 290, 120048. [Google Scholar] [CrossRef]
  45. Tian, Y.; Li, Q.; Zhang, M.; Nie, Y.; Tian, X.; Yang, C.; Li, Y. pH-dependent oxidation mechanisms over FeCu doped g-C3N4 for ofloxacin degradation via the efficient peroxymonosulfate activation. J. Clean. Prod. 2021, 315, 128207. [Google Scholar] [CrossRef]
  46. Song, H.; Pan, S.; Wang, Y.; Cai, Y.; Zhang, W.; Shen, Y.; Li, C. MXene-mediated electron transfer in Cu(II)/PMS process: From Cu(III) to Cu(I). Sep. Purif. Technol. 2022, 297, 121428. [Google Scholar] [CrossRef]
  47. Mo, F.; Liu, Y.; Xu, Y.; He, Q.; Sun, P.; Dong, X. Photocatalytic elimination of moxifloxacin by two-dimensional graphitic carbon nitride nanosheets: Enhanced activity, degradation mechanism and potential practical application. Sep. Purif. Technol. 2022, 292, 121067. [Google Scholar] [CrossRef]
  48. Xu, A.; Wu, D.; Zhang, R.; Fan, S.; Lebedev, A.T.; Zhang, Y. Bio-synthesis of Co-doped FeMnOx and its efficient activation of peroxymonosulfate for the degradation of moxifloxacin. Chem. Eng. J. 2022, 435, 134695. [Google Scholar] [CrossRef]
  49. Yang, H.; Xue, W.; Liu, M.; Yu, K.; Yu, W. Carbon doped Fe3O4 peroxidase-like nanozyme for mitigating the membrane fouling by NOM at neutral pH. Water Res. 2020, 174, 115637. [Google Scholar] [CrossRef]
  50. Jacob, J.; Prádanos, P.; Calvo, J.I.; Hernández, A.; Jonsson, G. Fouling kinetics and associated dynamics of structural modifications. Colloids Surf. A. 1998, 138, 173–183. [Google Scholar] [CrossRef]
  51. Bolton, G.; Lacasse, D.; Kuriyel, R. Combined models of membrane fouling: Development and application to microfiltration and ultrafiltration of biological fluids. J. Membr. Sci. 2006, 277, 75–84. [Google Scholar] [CrossRef]
  52. Simon, A.; Price, W.E.; Nghiem, L.D. Changes in surface properties and separation efficiency of a nanofiltration membrane after repeated fouling and chemical cleaning cycles. Sep. Purif. Technol. 2013, 113, 42–50. [Google Scholar] [CrossRef]
Figure 1. SEM images of (a) MXene, (b) CuFeS2/MXene, (c) PVDF membrane, and (d) CuFeS2/MXene–PVDF membrane.
Figure 1. SEM images of (a) MXene, (b) CuFeS2/MXene, (c) PVDF membrane, and (d) CuFeS2/MXene–PVDF membrane.
Water 16 01504 g001
Figure 2. (a) XRD patterns of three materials; XPS spectra of (b) CuFeS2, (c) CuFeS2/MXene and (d) CuFeS2/MXene–PVDF membrane.
Figure 2. (a) XRD patterns of three materials; XPS spectra of (b) CuFeS2, (c) CuFeS2/MXene and (d) CuFeS2/MXene–PVDF membrane.
Water 16 01504 g002
Figure 3. FTIR spectra of PVDF membrane and CuFeS2/MXene–PVDF membrane.
Figure 3. FTIR spectra of PVDF membrane and CuFeS2/MXene–PVDF membrane.
Water 16 01504 g003
Figure 4. Contact angles of (a) PVDF membrane and (b) CuFeS2/MXene–PVDF membrane.
Figure 4. Contact angles of (a) PVDF membrane and (b) CuFeS2/MXene–PVDF membrane.
Water 16 01504 g004
Figure 5. Removal of MOX by catalytic membrane/PMS system: (a) different systems, (b) CuFeS2/MXene loading, (c) different catalysts, and (d) PMS dosage.
Figure 5. Removal of MOX by catalytic membrane/PMS system: (a) different systems, (b) CuFeS2/MXene loading, (c) different catalysts, and (d) PMS dosage.
Water 16 01504 g005
Figure 6. The effects of operating conditions on the removal of MOX: (a) filtration pressure and (b) batch filtration.
Figure 6. The effects of operating conditions on the removal of MOX: (a) filtration pressure and (b) batch filtration.
Water 16 01504 g006
Figure 7. The effects of water matrix on the removal of MOX: (a) initial pH value, and (b) impurities in water.
Figure 7. The effects of water matrix on the removal of MOX: (a) initial pH value, and (b) impurities in water.
Water 16 01504 g007
Figure 8. (a) Effects of radical scavengers on the removal efficiencies of MOX; (b) EPR spectra of DMPO-SO4•−, DMPO-HO, and TEMP-1O2 adduct.
Figure 8. (a) Effects of radical scavengers on the removal efficiencies of MOX; (b) EPR spectra of DMPO-SO4•−, DMPO-HO, and TEMP-1O2 adduct.
Water 16 01504 g008
Figure 9. (a) Mayer bonds of MOX; (b) the mechanism of the catalytic membrane/PMS for removing MOX.
Figure 9. (a) Mayer bonds of MOX; (b) the mechanism of the catalytic membrane/PMS for removing MOX.
Water 16 01504 g009
Figure 10. Membrane resistance and membrane fouling resistance.
Figure 10. Membrane resistance and membrane fouling resistance.
Water 16 01504 g010
Figure 11. Membrane fouling model: (a) complete blocking (n = 2), (b) standard blocking (n = 1.5), (c) intermediate blocking (n = 1), and (d) cake filtration (n = 0).
Figure 11. Membrane fouling model: (a) complete blocking (n = 2), (b) standard blocking (n = 1.5), (c) intermediate blocking (n = 1), and (d) cake filtration (n = 0).
Water 16 01504 g011
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

Zhang, D.; Li, K.; Fang, L.; Chen, H. CuFeS2/MXene-Modified Polyvinylidene Fluoride Membrane for Antibiotics Removal through Peroxymonosulfate Activation. Water 2024, 16, 1504. https://doi.org/10.3390/w16111504

AMA Style

Zhang D, Li K, Fang L, Chen H. CuFeS2/MXene-Modified Polyvinylidene Fluoride Membrane for Antibiotics Removal through Peroxymonosulfate Activation. Water. 2024; 16(11):1504. https://doi.org/10.3390/w16111504

Chicago/Turabian Style

Zhang, Dongyang, Kunfu Li, Lei Fang, and Huishan Chen. 2024. "CuFeS2/MXene-Modified Polyvinylidene Fluoride Membrane for Antibiotics Removal through Peroxymonosulfate Activation" Water 16, no. 11: 1504. https://doi.org/10.3390/w16111504

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