Next Article in Journal
Water-Holding Properties of Clinoptilolite/Sodium Polyacrylate-Modified Compacted Clay Cover of Tailing Pond
Next Article in Special Issue
Low-Temperature Reduction Synthesis of γ–Fe2O3−x@biochar Catalysts and Their Combining with Peroxymonosulfate for Quinclorac Degradation
Previous Article in Journal
Long-Term Multi-Sensory Gamma Stimulation of Dementia Patients: A Case Series Report
Previous Article in Special Issue
Evaluating the Performance of Ball-Milled Silk Fibroin Films for Simultaneous Adsorption of Eight Pharmaceuticals from Water
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis and Characterization of g-C3N4/Ag3PO4/TiO2/PVDF Membrane with Remarkable Self-Cleaning Properties for Rhodamine B Removal

1
Hunan Key Laboratory of Applied Environmental Photocatalysis, Changsha University, Changsha 410022, China
2
College of Environmental Science and Engineering, Hunan University, Changsha 410082, China
*
Authors to whom correspondence should be addressed.
Int. J. Environ. Res. Public Health 2022, 19(23), 15551; https://doi.org/10.3390/ijerph192315551
Submission received: 31 October 2022 / Revised: 17 November 2022 / Accepted: 20 November 2022 / Published: 23 November 2022

Abstract

:
g-C3N4/Ag3PO4/TiO2 nanocomposite materials were loaded onto a polyvinylidene fluoride (PVDF) membrane using a phase inversion method to obtain a photocatalytic flat membrane for dye removal. The morphology, structure, and photocatalytic activity of the g-C3N4/Ag3PO4/TiO2 nanoparticles and composite membrane were evaluated. The g-C3N4/Ag3PO4/TiO2/PVDF membrane exhibited superior morphology, hydrophilic properties, and antifouling performance compared with the raw PVDF membrane. Four-stage filtration was performed to evaluate the self-cleaning and antifouling capacity of the g-C3N4/Ag3PO4/TiO2/PVDF membrane. Upon irradiating the composite membrane with visible light for 30 min, its irreversible fouling resistance (Rir) was low (9%), and its flux recovery rate (FRR) was high (71.0%) after five filtration cycles. The removal rate of rhodamine B (RhB) from the composite membrane under visible light irradiation reached 98.1% owing to the high photocatalytic activity of the membrane, which was superior to that of raw PVDF membrane (42.5%). A mechanism of photocatalytic composite membranes for RhB degradation was proposed. Therefore, this study is expected to broaden prospects in the field of membrane filtration technology.

1. Introduction

Rhodamine B (RhB) is one of the most popular dyes for printing, textile, and leather industries, which could cause pollution to the environment if not treated appropriately. Traditional physical methods, such as physical adsorption, coagulation, and chemical oxidation, and biological methods have been unable to achieve the efficient treatment of dye wastewater. Membranes with good selectivity, high efficiency, and excellent stability are widely used to treat municipal drinking water and wastewater. However, membrane fouling and further treatment of concentrated contaminant solutions are bottlenecks that have hindered the development of membrane separation technologies [1,2]. Surface modifications, which increase the hydrophilicity and improve the self-cleaning ability of membranes, are effective methods for enhancing the antifouling properties of membranes. Polyvinylidene fluoride (PVDF) has been widely employed in many areas because it involves a simple preparation process, can be easily modified, presents high strength, and is inexpensive [3,4]. However, PVDF membranes should be modified to confer on them hydrophilic properties and improve their anti-pollution performance. Among the various modification methods studied, the introduction of inorganic nanocomponents into PVDF membranes is an effective approach because nanoparticles present surface and interface effects; moreover, a small amount of dopant can confer on membranes many new functions. Recently, doping PVDF membranes with nanoparticles with photocatalytic properties, such as Fe(NO)3 [5], RbxWO3 [6,7], C3N4 [8,9,10], ZrO2 [11,12], ZnO [13,14], Ag3PO4 [15,16], TiO2 [17,18], has become a popular research topic, because the fabricated photocatalyst-loaded membranes effectively integrate photocatalytic oxidation degradation of pollutants and separation within a single device. Since photocatalysts can effectively degrade contaminants in solutions and on the membrane surface, photocatalyst-loaded membranes present outstanding removal efficiency and remarkable anti-pollution performance [19,20].
Graphitic carbon nitride (g-C3N4), which primarily comprises C and N, presents remarkable application potential owing to its excellent photocatalytic performance and chemical stability. However, the photocatalytic activity of pure g-C3N4 has the disadvantages of a high carrier recombination rate, sparse adsorption sites, and active sites above 450 nm, which result in a volatile photocatalytic efficiency [21,22]. The recombination rate of photogenerated carriers in composites fabricated using two or more semiconductors (or heterojunction structures) is lower than that of their individual components; therefore, the photocatalytic performance of composites is superior to that of their individual components. Numerous semiconductor materials, such as MoO3 [23], TiO2 [24], ZnO [25], SnO2 [26], ZrO2 [27], BIOI [28], Ag3PO4 [29,30,31], and graphene oxide (GO) [8,32], were coupled with g-C3N4 to accelerate the separation of photogenerated electron–hole pairs and increase the bandgap energy. Among them, Ag3PO4 has been applied as a highly effective compound for the visible-light-driven photodegradation of various organic pollutants in an aqueous solution; however, it is prone to photocorrosion in practical applications. The electric field that typically forms upon coupling g-C3N4 with Ag3PO4 can significantly increase the transfer rate of photogenerated charge carriers [29]. The simultaneous addition of TiO2 and Ag3PO4 to g-C3N4 inhibits the photocorrosion of Ag3PO4 because the band positions of TiO2 and Ag3PO4 match. Furthermore, other methods aimed at improving photocatalytic activity by combining g-C3N4 with Ag3PO4 and other Ag-based compounds or TiO2 to form heterojunctions have been successfully developed. Abbasi-Asl et al. [30] prepared a TiO2/Ag3PO4/g-C3N4 semiconductor composite and used it for the degradation of metronidazole (MNZ). The photodegradation efficiency of the composite nanoparticles for MNZ reached 97.18% under optimal experimental conditions. The comparison experiment showed that the photocatalytic efficiency of ternary composites (TiO2/Ag3PO4/g-C3N4) was far better than that of pure and binary samples. Cui et al. [22] fabricated g-C3N4/Ag3PO4/PVDF photocatalytic porous membranes using a phase inversion method and utilized them for RhB removal. The removal efficiency of the g-C3N4/Ag3PO4/PVDF membrane for RhB reached 97%, which was superior to those of the g-C3N4/PVDF and pure PVDF membranes (85% and 41%, respectively).
In this study, a novel type of hydrophilic PVDF membrane loaded with g-C3N4/Ag3PO4/TiO2 photocatalytic nanocomposite was prepared via blending and phase inversion. The morphology, structure, and photocatalytic activity of the g-C3N4/Ag3PO4/TiO2 nanoparticles and composite membrane were characterized and evaluated. A four-stage filtration experiment was used to evaluate the membrane performance for the separation of bovine serum albumin (BSA) and the membrane self-cleaning mechanism. Additionally, the photocatalytic degradation and rejection rate of the composite membrane for RhB in solution were evaluated. The renewability of the composite membrane over repeated RhB photodegradation and adsorption cycles was also evaluated. A mechanism of photocatalytic composite membranes for RhB degradation was proposed.

2. Material and Methods

2.1. Materials

Commercial PVDF (MW 200,000 Da) was purchased from Guangdong Zhan Yang Co., Ltd. (Guangzhou, China). Melamine, titanium dioxide, N,N-dimethylacetamide (DMAc), polyvinylpyrrolidone (PVP), silver nitrate (AgNO3), disodium phosphate (Na2HPO4), BSA, RhB, and hydrochloric acid (HCl) were acquired from Aladdin.

2.2. Synthesis of g-C3N4

g-C3N4 was prepared via melamine pyrolysis using a one-step synthesis method. Melamine (10 g) was added to a porcelain crucible with a lid and calcinated at 550 °C under a heating rate of 5 °C/min in a muffle furnace for 2 h, and then it was cooled to 25 °C. The obtained milky yellow powder was purified with a 1.0 M HCl solution at 150 °C under a heating rate of 3 °C/min for 4 h. Eventually, the fabricated solid was collected through centrifugation, followed by drying, and ground into a uniform powder using an agate mortar and pestle prior to further use [31,33].

2.3. Synthesis of g-C3N4/Ag3PO4/TiO2 Nanocomposites

Ag3PO4 was synthesized via precipitation. g-C3N4/Ag3PO4/TiO2 composite nanoparticles were prepared simultaneously as follows: 0.4 g of g-C3N4, 1.0 g of PVP, and 0.4 g of TiO2 were added to 80 mL of distilled water (suspension 1). After 15 min of intermittent ultrasonication, the suspension was subjected to magnetic stirring for 30 min. Next, 5.1 g AgNO3 was added to suspension 1 (mixture 2). Thereafter, 100 mL of a 0.1 mol/L Na2HPO4 solution was added dropwise to mixture 2 using a burette. After the mixture was stirred for 3 h in the dark, a yellow sediment was separated via centrifugation, followed by rinsing three times with water and ethanol, alternately. Lastly, the residue was dried to a constant weight in a vacuum oven at 50 °C.

2.4. Preparation of Photocatalytic Composite Membranes

PVDF-based photocatalytic membranes using PVP as the pore former were synthesized by blending g-C3N4/TiO2/Ag3PO4 nanoparticles with PVDF using phase inversion technique. In brief, 1.5 g of g-C3N4/TiO2/Ag3PO4 nanoparticles, 2.0 g of PVP, and 40.0 g of DMAc were added to a 250 mL beaker placed in a 60 °C oil bath, and the reactants were stirred at a constant speed until the mixture became transparent. Thereafter, 7.5 g of PVDF powder was added to the beaker step by step with continuous stirring for 6 h at 60 °C to form a homogeneous and transparent casting mixture. Next, the mixture was placed in a vacuum-drying oven for 12 h at 60 °C to remove air bubbles. After degassing at 20 °C and 70% humidity, the dopant mixture was uniformly spread on a clean glass plate using a coater with 200 μm gap. The glass plate was allowed to rest in ambient air for 30 s, and then it was placed in a coagulation bath at 20 °C. Lastly, the prepared membrane was placed in ultrapure water for 24 h. Figure 1 shows the schematics of the preparation of the nanoparticles and the photocatalytic composite membranes.

2.5. Characterization of the g-C3N4/Ag3PO4/TiO2 Nanocomposites

The crystalline structure and phase purity of the photocatalyst powders were analyzed using an Ultima4 (Rigaku) X-ray diffraction (XRD) instrument. The morphology of the photocatalyst powders was evaluated using a JSM 7001F (JEOL) SEM apparatus. An F-4500 (Hitachi Corp.) photoluminescence (PL) spectrometer was used to investigate the photochemical properties of the samples.

2.6. Membrane Characterization

The roughness and morphology of the membrane surfaces were evaluated using the SEM JSM 7001F (JEOL, Tokyo, Japan) and a Dimension Icon (Bruker, Madison, WI, USA) AFM device, respectively. TGA experiments were conducted using a Q5000IR (TA Instruments, New Castle, DE, USA) system. The interfacial interactions and crystallinity in the membranes were analyzed using an IS10 (Thermo Fisher, Waltham, MA, USA) FTIR spectrometer. The mechanical performance of the membranes was measured using a CMT4204 electronic universal testing machine (XinSanSi, Shanghai, China) at room temperature. The hydrophilic performance of the membranes was estimated by testing their water contact angles with an OCA 40 (Dataphysics, San Jose, CA, USA) goniometer. The sessile drop technique with dynamic contact angle measurement was used, and contact angles were measured for 30 s [34].

2.7. Basic Properties of Membranes

2.7.1. Permeability Measurements

The prepared membranes were placed in a dead-end stirred ultrafiltration cup to evaluate their permeability (Figure 2). The effective filtration area of the ultrafiltration cup was 50.2 cm2. Permeability measurements were performed under an operating pressure of 0.1 MPa, as follows: The membrane was pre-pressured with deionized water until the permeate flux and operating pressure were stabilized. The volume of deionized water passing through the membrane was determined within 5 min. The permeate flux was calculated as follows:
J 0   = V At
where J0 is initial permeate flux of deionized water (L/(m2 h)), V, A, and t are the volume of deionized water collected within 5 min (L), effective filtration area (m2), and time employed during filtration (h), respectively.
The pore size and porosity of the membrane were measured by a weighing method. The operation process and calculation formula were similar to the existing reports [35]. The porosity (ε) was calculated as follows:
ε =   m w m d ρ b ×   V d
where md and mw are the weight of the membrane before and after immersion in n-butanol, ρb is the density of n-butanol, and Vd is the dry membrane volume.2.7.2. Membrane reusability
After the membranes were pre-pressurized with deionized water, the feed solution was changed to a 0.2 g/L BSA solution, and the stable flux values (JP) were calculated using Equation (1). The concentration of BSA in the permeate was determined after 30 min, and the rejection of BSA (R) was calculated as follows:
R ( % ) = ( 1 C C 0 ) × 100 %
where C0 (mg/L) and C (mg/L) are related to the contaminant concentrations before and after filtration, respectively.
Subsequently, the fouled membrane was rinsed repeatedly with deionized water, and its permeate flux (Jrw) was recorded after 10 min. Lastly, the membrane was illuminated using a Xe lamp, and the water flux (Jw) was determined after 30 min. This process was repeated five times. The flux recovery ratio (FRR), total fouling ratio (Rt), reversible fouling ratio (Rr), and irreversible fouling ratio (Rir) of the membranes were calculated as follows:
FRR = J w J 0 × 100 %
R t   = J 0   J p J 0 × 100 %
R r   = J w   J p J 0 × 100 %
R ir   = J 0   J w J 0 × 100 %

2.7.2. Removal of RhB Using the Photocatalytic Membrane

A 10 mg/L solution of RhB, which was used to simulate wastewater, was filtered using an ultrafiltration cup and the photocatalytic membrane under the illumination of a 300 W Xe lamp. During this process, a piece of 420 nm cutoff filter was applied to cover the top of the ultrafiltration cup. The system was kept in dark for 30 min to reach adsorption equilibrium before light illumination. During the visible light illumination, 3 mL of permeate solution was collected from the ultrafiltration cup every 7.5 min and analyzed using an ultraviolet (UV)–visible (Vis) spectrophotometer at 554 nm to determine the concentration of residual RhB. The rejection of RhB was calculated using Equation (3). The time evolution of the concentration of RhB was fitted using the following pseudo-first-order kinetic equation:
ln C C 0 = kt
where k (min−1) represents the rate constant and t is the time employed during filtration.

3. Results and Discussion

3.1. Characterization of the g-C3N4/Ag3PO4/TiO2 Nanocomposites

SEM images of g-C3N4, Ag3PO4/TiO2, and g-C3N4/Ag3PO4/TiO2 are shown in Figure 3. A two-dimensional, micrometer-sized, solid agglomerated structure of g-C3N4 was homogeneously formed (Figure 3a). g-C3N4 presented a typical thin lamellar structure of graphite-phase C3N4 with distinct folds, which conferred on it its large specific surface area. Ag3PO4 comprised 300–500 nm spheres, which connected tightly with the ultra-thin g-C3N4 layers (Figure 3b). The Ag3PO4/TiO2 nanoparticles presented a regular spherical shape similar to that of the pure Ag3PO4 particles that attached to the g-C3N4 sheets (Figure 3c,d).
Figure 4a shows the XRD patterns of pure g-C3N4, Ag3PO4/TiO2, and g-C3N4/Ag3PO4/TiO2. The characteristic peaks in the XRD pattern of pure g-C3N4 at 2θ = 27.8° (strong) and 12.6° (weak) were attributed to the (002) and (100) planes of g-C3N4, respectively [36]. The peaks at 20.80°, 29.68°, 33.31°, 36.59°, 42.54°, 47.78°, 52.60°, 55.05°, 57.29°, 61.62°, 65.81°, 69.94°,71.89°, and 73.85° in the XRD pattern of Ag3PO4 corresponded to planes (110), (200), (210), (211), (220), (310), (222), (320), (321), (400), (330), (420), (421), and (322), respectively, of cubic-phase Ag3PO4. No impurity peaks were present in the XRD patterns of the samples, and the sharp peaks indicate good crystallization of the samples [33]. It is noteworthy that in the ternary catalyst, the (002) and (100) planes of g-C3N4 disappeared, while the diffraction peak of Ag3PO4 was more prominent, which indicated that the crystal of Ag3PO4 grew. As shown in Figure 4b, it was clear that the PL peak intensity of C3N4 was the highest, that is, the photogenerated electron–hole recombination of its monomer was the easiest. However, the spectral PL intensity of g-C3N4/Ag3PO4/TiO2 composites was significantly suppressed in comparison to g-C3N4 or Ag3PO4/TiO2, which showed that the ternary composite catalyst had the highest photogenerated current-carrying separation efficiency and the lowest recombination rate of the photogenerated charge carriers [37]. That is to say, the composite of g-C3N4/Ag3PO4/TiO2 can effectively improve the diffusion rate and charge mobility of photogenerated charge.

3.2. Membrane Characterizations

3.2.1. Membrane Morphology

Figure 5a,b show the SEM images of the surface and cross-section of the raw PVDF membrane, respectively, whereas Figure 5c,d show those of the g-C3N4/Ag3PO4/TiO2/PVDF membrane. Surface and cross-sectional images were used to study the dispersion of nanoparticles on the PVDF membrane and probe the thickness and homogeneity of the coating layer. As shown in Figure 5a,c, the raw PVDF membrane displayed a cavity-like structure, and the composite membrane presented a dense top layer supported by a porous layer in the surface image. All cross-sections of membranes showed an asymmetric microscopic porous structure and spongy layers surrounding a finger-like macroscopic cavity. By contrast, the finger-like pore wall of the composite membrane appeared rougher and more porous. The addition of nanoparticles to PVDF promoted the demixing rate by increasing the thermodynamic instability of the membrane. During phase inversion, g-C3N4/Ag3PO4/TiO2 nanoparticles strongly adsorbed water molecules, which provided abundant sites for the infiltration of water molecules and led to pore formation in the surface layer of the membrane.
Figure 6 shows the AFM images of the raw PVDF and g-C3N4/Ag3PO4/TiO2/PVDF membranes. Unlike the PVDF membrane, the g-C3N4/Ag3PO4/TiO2/PVDF membrane presented distinct ridges and valleys on its surface; furthermore, more holes were obtained on the surface of the composite membrane than on the surface of the raw PVDF membrane. The higher surface roughness of the composite membrane was attributed to the nodular structures that formed on the top layer of the membrane owing to the aggregation of the photocatalytic material. The Ra values of the raw PVDF and g-C3N4/Ag3PO4/TiO2/PVDF membranes were 93 and 309 nm, respectively, and the Rq values of the raw and composite membranes were 113 and 395 nm, respectively. These data indicated that the surface of the composite membrane was rougher than that of the raw membrane, which provided optimal conditions for adequate contact between photocatalytic materials and contaminants in solutions. This was attributed to the nodular structure generated via the aggregation of photocatalytic material on the membrane surface [8,38]. It was worth noting the results did not match with reference [15] which indicated that the addition of a ternary composite of Ag3PO4/GO/APTES decreased the roughness of PVDF membranes.

3.2.2. Basic Properties of Membranes

Some basic properties of two types of membranes are shown in Table 1 and Figure 7. The result of the porosity test was consistent with that of the SEM. The thickness of the composite membrane was higher than that of the raw membrane owing to the presence of nanoparticles. Tensile strength measurements were used to investigate the mechanical properties of raw and composite membranes. As shown in Table 1, the tensile stress of the g-C3N4/Ag3PO4/TiO2/PVDF membrane was lower than that of the raw PVDF membrane. That can be explained by the fact that more porous structures could decrease the membrane’s mechanical performance. Likewise, agglomeration resulting from excessive incorporation of g-C3N4/Ag3PO4/TiO2 made the dispersion of particles in the polymer matrix nonuniform. When the membrane worked under the action of an external force, the internal stress concentration polarization took place, resulting in the decreased mechanical performance [5].
As shown in Figure 7a, there are no obvious differences in the FTIR spectra between the raw and composite membranes. The obtained results show that relative overlapping is the dominant phenomenon in these spectra. The distinct peaks at 1440 and 1170 cm−1 in the FTIR spectra of the raw PVDF and g-C3N4/Ag3PO4/TiO2/PVDF membranes were attributed to the C–H deformation vibration and the C–F vibration, respectively. Furthermore, the absence of the characteristic absorption peaks of Ag3PO4 in the FTIR spectra of the g-C3N4/Ag3PO4/TiO2/PVDF membrane was attributed to the relatively low amount of Ag3PO4 or overlap with other peaks [39].
Two types of membranes were heated at 15 °C/min at temperatures ranging from 25 °C to 700 °C. The TGA curves of the raw PVDF and g-C3N4/Ag3PO4/TiO2/PVDF membranes, which were used to evaluate the thermal stability of the membranes (Figure 7b), showed that the TGA curves of the raw and composite membranes were comparable, which was attributed to the decomposition and evaporation of small polymer fragments at high temperatures. The significant mass loss (72.3%) in the range of 306–473 °C was attributed to the thermal decomposition of the PVDF backbone. A distinct weight loss was exhibited in the TGA curve of g-C3N4/Ag3PO4/TiO2/PVDF as the temperature was increased from 317 to 540 °C, and the weight loss rate was ~60.9% at 700 °C. The weight loss rate and quantity of the g-C3N4/Ag3PO4/TiO2/PVDF membrane were lower than those of the raw PVDF membrane. This was attributed to the addition of composite particles forming hydrogen or coordination bonds between organic macromolecular interactions, hindering the thermal motion of molecules and the movement of macromolecular chains. Consequently, the energy needed to break the bonds increased, and the thermal stability of the membrane increased [35]. Cacho-Bailo [40] obtained similar results by adding a metal–organic skeleton to polysulfone membranes.
Surface hydrophilicity is a critical factor for the anti-pollution performance of membranes. Generally, the lower the contact angle, the higher the membrane’s hydrophilic performance; in other words, the greater the resistance against fouling. The contact angle of the raw PVDF membrane (83.15°) was higher than that of the composite membrane (70.32°) (Figure 7c). This was attributed to the nanoparticles loaded on the top layer of the PVDF membrane. Overall, the addition of g-C3N4/Ag3PO4/TiO2 to PVDF improved membrane hydrophilicity. Generally, the more hydrophilic composites added, the stabler and denser the hydration layer which could rule out more contaminant molecules was obtained. At the same time, hydrogen bonds formed between g-C3N4 and water molecules, which promoted the diffusion of water molecules in the membrane and weakened the adsorption of contaminants on the membrane surface [41,42]. The hydrophilicity improvement of the membrane also resulted from the hydroxyl radical produced by the composite nanoparticles. In addition, according to previous studies [43,44], surface roughness can improve membrane wettability owing to the formation of composite nanoparticle layers. A few contaminant molecules could accumulate in the “valleys” on the surface accordingly. These results were consistent with the aforementioned SEM and AFM results.

3.2.3. Antifouling Performance of Membranes

The ratios of reversible and irreversible fouling (Rir and Rr, respectively) to the total fouling (Rt) of the membranes are critical parameters for evaluating the antifouling performance of membranes (Figure 8a,b). Rir represents the fouling molecules that are tightly anchored to the membrane surface and trapped within its pores, whereas Rr primarily represents the foulant that is loosely bound to the membrane surface. Before visible light irradiation, the Rir values of the membranes were higher than the corresponding Rr values (Figure 8a,b), indicating that irreversible fouling was the predominant process. The Rir value of the composite membrane decreased considerably, from 61.1% to 8.9%, whereas the Rr value increased significantly, from 33.7% to 85.3%. For the composite membrane, a fraction of irreversible fouling was converted into reversible fouling. Notably, the Rr values of the g-C3N4/Ag3PO4/TiO2/PVDF membrane were remarkably greater than those of the raw PVDF membrane under both experimental conditions. Therefore, the addition of photocatalytic nanomaterials to PVDF improved the antifouling ability of the raw PVDF membrane.
In practice, it is necessary to evaluate the antifouling properties of membranes. Therefore, we subjected both membranes to four-stage filtration experiments for five cycles. The FRR values of both membranes decreased gradually after each cycle (Figure 8c), which was attributed to the accumulation of irreversible sums and the adsorption of dirt in the membrane pores. However, the composite nanoparticles promoted the self-cleaning ability of the membrane and lowered the FRR value of the g-C3N4/Ag3PO4/TiO2/PVDF membrane from 90.1% to 71.0%. In contrast, the FRR value of the raw PVDF membrane decreased from 78.9% to 50.3% after five filtration cycles. This was attributed to the photocatalytic performance and high hydrophilicity of the g-C3N4/Ag3PO4/TiO2 nanoparticles on the membrane surface, which could prevent the accumulation of foulant on the membrane surface during filtration. The BSA rejection of the raw and composited membranes was higher than 90% during the five filtration cycles. The addition of g-C3N4/Ag3PO4/TiO2 to PVDF increased the BSA rejection and flux of the PVDF membrane.

3.2.4. Photocatalytic Properties of Membranes

Figure 9a,b show the photocatalytic degradation efficiency and kinetic model of RhB, respectively. The blank test demonstrated that RhB rejection did not change in the absence of a membrane, so the photodecomposition of RhB was negligible.
The amount of RhB decreased only by 42.5% after 120 min of filtration using the raw PVDF membrane (Figure 9a). RhB removal by the raw PVDF membrane occurred exclusively via absorption. In contrast, the degradation efficiency of the g-C3N4/Ag3PO4/TiO2/PVDF membrane for RhB was significantly higher (98.1%) under visible light irradiation. The kinetic equation −ln(C/C0) = kt was used to evaluate the photodegradation of RhB (Figure 9b). The k value of the raw PVDF membrane (0.00549 min−1) was negligible as it was approximately seven times lower than that of the composite membrane (0.03618 min−1) under the same experimental conditions. These results indicate that the g-C3N4/Ag3PO4/TiO2 coating improved the photocatalytic activity of the PVDF membrane. The UV–Vis absorption spectra of an RhB solution subjected to photodegradation using the g-C3N4/Ag3PO4/TiO2/PVDF membrane under visible light irradiation for 90 min are shown in Figure 9c. The maximum absorption peaks at 554 nm declined significantly, consistent with the results shown in Figure 9a.
The membranes were used to remove RhB for five cycles, and the experimental data provided information about the reusability and stability of the g-C3N4/Ag3PO4/TiO2/PVDF membrane. The removal rates of RhB using the composite membrane and a dead-end stirred ultrafiltration cup under visible light irradiation for five cycles were 98.1%, 97.0%, 96.5%, 95.6%, and 95.3%, given chronologically (Figure 10a,b). After five cycles, the photocatalytic performance of the composite membrane was still as excellent as in its original state. The low attenuation rate after the five experimental cycles indicates the excellent stability and regeneration capacity of the composite membrane, for the reason that the g-C3N4/Ag3PO4/TiO2 nanoparticles are stably bound in the membrane without loss.

3.3. Mechanism of the g-C3N4/Ag3PO4/TiO2/PVDF Membrane

The photodegradation mechanism of RhB by the C3N4/Ag3PO4/TiO2/PVDF membrane is shown in Figure 11. First, RhB molecules were electrostatically adsorbed onto the surface and within the pores of the membrane. Subsequently, the C3N4/Ag3PO4/TiO2 nanoparticles became excited under visible light irradiation and photogenerated electrons and holes and migrated to the surface of the nanocomposite.
According to previous studies [30,45,46], the bandgap energies of C3N4, Ag3PO4, and TiO2 are approximately 2.76, 2.31, and 3.20 eV, respectively. The conduction band (CB) of g-C3N4 (–1.16 eV) is more negative than those of TiO2 (−0.29 eV) and Ag3PO4 (+0.299 eV); therefore, superoxide radicals (•O2) can be generated as the photoinduced electrons in the CB of g-C3N4 reduce the absorbed O2. Similarly, the electrons in the CB of TiO2 migrated to Ag3PO4. Furthermore, the valence band (VB) potential of Ag3PO4 (+2.61 eV) was more positive than those of g-C3N4 (+1.60eV) and the H2O/•OH pair (+2.38 eV); therefore, •OH active species can be formed via the oxidation of H2O by the holes in the VB of Ag3PO4 [47]. Subsequently, the strong oxidizing radicals (•O2 and •OH) and photogenerated holes could thoroughly oxidize RhB into CO2 and H2O. In addition, Ag+ ions can be reduced to metallic Ag with the help of generating enough energy, which can be excited by Ag3PO4 and g-C3N4 under light irradiation. The photogenerated electrons in the CB of Ag3PO4 can be transferred to metallic Ag via the constructed Ag bridge and bind to the holes in the VB of g-C3N4 [47,48]. According to previously published papers [49,50,51], the hybrid C3N4/Ag3PO4/TiO2 ternary composite photocatalyst had a higher electron–hole pair separation efficiency and a lower recombination rate than the binary composite photocatalyst.

4. Conclusions

In summary, g-C3N4/Ag3PO4/TiO2 nanocomposites were prepared and used as precursors to fabricate g-C3N4/Ag3PO4/TiO2/PVDF photocatalyst membranes using the phase inversion method. The g-C3N4/Ag3PO4/TiO2/PVDF membrane presented excellent self-cleaning properties and high RhB removal performance during membrane separation and photodegradation. Therefore, our results demonstrated that the nanocomposites improved the antifouling and self-cleaning performance of the membrane. Furthermore, the visible-light-driven photocatalytic degradation of RhB on the membrane surface by the g-C3N4/Ag3PO4/TiO2 nanoparticles conferred the composite membrane’s excellent self-cleaning ability and remarkable regeneration capacity.
The photocatalytic performance of the g-C3N4/Ag3PO4/TiO2/PVDF membrane has been improved compared to previous studies of TiO2-based photocatalytic membranes. Mericq et al. [52] demonstrated that the TiO2 nanoparticles improved the structure of the membrane as well as its antifouling performance under UV irradiation. Zhang et al. [41] found that Ag/g-C3N4 membranes degraded a maximum of 77% of methyl orange in 100 min. Cui et al. [22] fabricated g-C3N4/Ag3PO4/PVDF membranes for RhB removal (97%), which were superior to those of g-C3N4/PVDF and pure PVDF membranes (85% and 41%, respectively). It is worth noting that they did not take the mechanical properties of the membrane and the exudation of nanoparticles into account. In fact, the incorporation of g-C3N4/Ag3PO4/TiO2 which resulted in the decreased mechanical performance of the membrane cannot be ignored. Hence, the further improvement of the mechanical strength of the membrane is one of the keys to the application of photocatalytic membranes in the future. Overall, this study can expand the application of g-C3N4/Ag3PO4/TiO2 as an advanced membrane material and the use of composite membranes for wastewater treatment.

Author Contributions

Conceptualization, X.L.; data curation, R.L.; formal analysis, Y.P.; investigation, Q.W.; methodology, X.L.; project administration, X.L.; resources, J.H.; supervision, J.H.; validation, H.P.; visualization, H.P. and L.W.; writing—original draft, R.L.; writing—review and editing, K.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the National Natural Science Foundation of China (51308076, 52000013); Natural Science Foundation of Hunan of China (2020JJ4643, 2022JJ30634); and Innovative Province Construction Special Fund of Hunan province (2020SK2016).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The datasets generated during and/or analyzed during the current study are available from the corresponding author on reasonable request.

Conflicts of Interest

The authors declare that they have no conflict of interest.

References

  1. Zhou, A.; Jia, R.; Wang, Y.L.; Sun, S.H.; Xin, X.D.; Wang, M.Q.; Zhao, Q.H.; Zhu, H.H. Abatement of sulfadiazine in water under a modified ultrafiltration membrane (PVDF-PVP-TiO2-dopamine) filtration-photocatalysis system. Sep. Purif. Technol. 2020, 234, 116099. [Google Scholar] [CrossRef]
  2. Li, X.; He, S.B.; Feng, C.L.; Zhu, Y.K.; Pang, Y.; Hou, J.; Luo, K.; Liao, X.S. Non-competitive and competitive adsorption of Pb2+, Cd2+ and Zn2+ ions onto SDS in process of micellar-enhanced ultrafiltration. Sustainability 2018, 10, 92. [Google Scholar] [CrossRef] [Green Version]
  3. Wan, J.; Huang, J.H.; Yu, H.B.; Liu, L.S.; Shi, Y.H.; Liu, C.H. Fabrication of self-assembled 0D-2D Bi2MoO6-g-C3N4 photocatalytic composite membrane based on PDA intermediate coating with visible light self-cleaning performance. J. Colloid Interface Sci. 2021, 601, 229–241. [Google Scholar] [CrossRef] [PubMed]
  4. Lan, T.; Huang, J.H.; Ouyang, Y.C.; Yi, K.X.; Yu, H.B.; Zhang, W.; Zhang, C.Y.; Li, S.Z. QQ-PAC core-shell structured quorum quenching beads for potential membrane antifouling properties. Enzym. Microb. Technol. 2021, 148, 109813. [Google Scholar] [CrossRef]
  5. Yang, C.Y.; Wang, P.; Li, J.N.; Wang, Q.; Xu, P.; You, S.J.; Zheng, Q.Z.; Zhang, G.S. Photocatalytic PVDF ultrafiltration membrane blended with visible-light responsive Fe(III)-TiO2 catalyst: Degradation kinetics, catalytic performance and reusability. Chem. Eng. J. 2021, 417, 129340. [Google Scholar] [CrossRef]
  6. Naseem, S.; Wu, C.M.; Motora, K.G. Novel multifunctional RbxWO3@Fe3O4 immobilized Janus membranes for desalination and synergic-photocatalytic water purification. Desalination 2021, 517, 115256. [Google Scholar] [CrossRef]
  7. Huang, X.Z.; Liu, H.W.; Zhang, X.; Jiang, H.R. High Performance All-Solid-State Flexible Micro-Pseudocapacitor Based on Hierarchically Nanostructured Tungsten Trioxide Composite. ACS Appl. Mater. Interfaces 2015, 7, 27845–27852. [Google Scholar] [CrossRef] [PubMed]
  8. Shi, Y.H.; Huang, J.H.; Zeng, G.M.; Cheng, W.J.; Hu, J.L.; Shi, L.X.; Yi, K.X. Evaluation of self-cleaning performance of the modified g-C3N4 and GO based PVDF membrane toward oil-in-water separation under visible-light. Chemosphere 2019, 230, 40–50. [Google Scholar] [CrossRef]
  9. Kolesnyk, I.; Kujawa, J.; Bubela, H.; Konovalova, V.; Burban, A.; Cyganiuk, A.; Kujawski, W. Photocatalytic properties of PVDF membranes modified with g-C3N4 in the process of Rhodamines decomposition. Sep. Purif. Technol. 2020, 250, 117231. [Google Scholar] [CrossRef]
  10. Huang, J.H.; Hu, J.L.; Shi, Y.H.; Zeng, G.M.; Cheng, W.J.; Yu, H.B.; Gu, Y.L.; Shi, L.X.; Yi, K.X. Evaluation of self-cleaning and photocatalytic properties of modified g-C3N4 based PVDF membranes driven by visible light. J. Colloid Interface Sci. 2019, 541, 356–366. [Google Scholar] [CrossRef]
  11. Chen, X.J.; Huang, G.; Li, Y.P.; An, C.J.; Feng, R.F.; Wu, Y.H.; Shen, J. Functional PVDF ultrafiltration membrane for Tetrabromobisphenol-A (TBBPA) removal with high water recovery. Water Res. 2020, 181, 115952. [Google Scholar] [CrossRef] [PubMed]
  12. Zhang, Y.Q.; Liu, P.L. Preparation of porous ZrO2 solid superacid shell/void/TiO2 core particles and effect of doping them on PVDF membranes properties. Chem. Eng. Sci. 2015, 135, 67–75. [Google Scholar] [CrossRef]
  13. Chen, X.J.; Huang, G.; An, C.J.; Feng, R.F.; Wu, Y.H.; Huang, C. Plasma-induced PAA-ZnO coated PVDF membrane for oily wastewater treatment: Preparation, optimization, and characterization through Taguchi OA design and synchrotron-based X-ray analysis. J. Membr. Sci. 2019, 582, 70–82. [Google Scholar] [CrossRef]
  14. Ayyaru, S.; Dinh, T.T.L.; Ahn, Y.H. Enhanced antifouling performance of PVDF ultrafiltration membrane by blending zinc oxide with support of graphene oxide nanoparticle. Chemosphere 2020, 241, 125068. [Google Scholar] [CrossRef] [PubMed]
  15. Zhang, R.; Li, Y.; Cai, Y.; Han, Y.; Zhang, T.Q.; Liu, Y.; Zeng, K.L.; Zhao, C. Photocatalytic Poly(vinylidene fluoride) membrane of Ag3PO4/GO/APTES for water treatment. Colloids Surf. A 2020, 597, 124779. [Google Scholar] [CrossRef]
  16. Fan, G.D.; Chen, C.G.; Chen, X.L.; Li, Z.S.; Bao, S.L.; Luo, J.; Tang, D.S.; Yan, Z.S. Enhancing the antifouling and rejection properties of PVDF membrane by Ag3PO4-GO modification. Sci. Total Environ. 2021, 801, 149611. [Google Scholar] [CrossRef]
  17. Safarpour, M.; Khataee, A.; Vatanpour, V. Effect of reduced graphene oxide/TiO2 nanocomposite with different molar ratios on the performance of PVDF ultrafiltration membranes. Sep. Purif. Technol. 2015, 140, 32–42. [Google Scholar] [CrossRef]
  18. Sakarkar, S.; Muthukumran, S.; Jegatheesan, V. Factors affecting the degradation of remazol turquoise blue (RTB) dye by titanium dioxide (TiO2) entrapped photocatalytic membrane. J. Environ. Manag. 2020, 272, 111090. [Google Scholar] [CrossRef]
  19. Che, H.N.; Liu, L.H.; Che, G.B.; Dong, H.J.; Liu, C.B.; Li, C.M. Control of energy band, layer structure and vacancy defect of graphitic carbon nitride by intercalated hydrogen bond effect of NO−3 toward improving photocatalytic performance. Biochem. Eng. J. 2019, 357, 209–219. [Google Scholar] [CrossRef]
  20. Yu, H.B.; Huang, J.H.; Jiang, L.B.; Yuan, X.Z.; Yi, K.X.; Zhang, W.; Zhang, J.; Chen, H.Y. Steering photo-excitons towards active sites: Intensified substrates affinity and spatial charge separation for photocatalytic molecular oxygen activation and pollutant removal. Chem. Eng. J. 2021, 408, 127334. [Google Scholar] [CrossRef]
  21. Cai, G.H.; Wang, J.P.; Wu, X.Q.; Zhan, Y.Y.; Liang, S.J. Scalable one-pot synthesis of porous 0D/2D C3N4 nanocomposites for efficient visible-light driven photocatalytic hydrogen evolution. Appl. Surf. Sci. 2018, 459, 224–232. [Google Scholar] [CrossRef]
  22. Cui, Y.H.; Yang, L.L.; Meng, M.J.; Zhang, Q.; Li, B.R.; Wu, Y.L.; Zhang, Y.L.; Lang, J.H.; Li, C.X. Facile preparation of antifouling g-C3N4/Ag3PO4 nanocomposite photocatalytic polyvinylidene fluoride membranes for effective removal of rhodamine B. Korean J. Chem. Eng. 2019, 36, 236–247. [Google Scholar] [CrossRef]
  23. Liu, L.S.; Huang, J.H.; Yu, H.B.; Wan, J.; Liu, L.Y.; Yi, K.X.; Zhang, W.; Zhang, C.Y. Construction of MoO3 nanopaticles /g-C3N4 nanosheets 0D/2D heterojuntion photocatalysts for enhanced photocatalytic degradation of antibiotic pollutant. Chemosphere 2021, 282, 131049. [Google Scholar] [CrossRef] [PubMed]
  24. Ni, S.; Fu, Z.; Li, L.; Ma, M.; Liu, Y. Step-scheme heterojunction g-C3N4/TiO2 for efficient photocatalytic degradation of tetracycline hydrochloride under, UV light. Colloids Surf. A 2022, 649, 129475. [Google Scholar] [CrossRef]
  25. Sun, Q.; Sun, Y.; Zhou, M.; Cheng, H.; Chen, H.; Dorus, B.; Lu, M.; Le, T. A 2D/3D g-C3N4/ZnO heterojunction enhanced visible-light driven photocatalytic activity for sulfonamides degradation. Ceram. Int. 2022, 48, 7283–7290. [Google Scholar] [CrossRef]
  26. Van, K.N.; Huu, H.T.; Nguyen Thi, V.N.; Le Thi, T.L.; Truong, D.H.; Truong, T.T.; Dao, N.N.; Vo, V.; Tran, D.L.; Vasseghian, Y. Facile construction of S-scheme SnO2/g-C3N4 photocatalyst for improved photoactivity. Chemosphere 2022, 289, 133120. [Google Scholar] [CrossRef] [PubMed]
  27. Khezami, L.; Ben Aissa, M.A.; Modwi, A.; Ismail, M.; Guesmi, A.; Algethami, F.K.; Ben Ticha, M.; Assadi, A.A.; Nguyen-Tri, P. Harmonizing the photocatalytic activity of g-C3N4 nanosheets by ZrO2 stuffing: From fabrication to experimental study for the wastewater treatment. Biochem. Eng. J. 2022, 182, 108411. [Google Scholar] [CrossRef]
  28. Yin, C.; Liu, Y.L.; Lv, X.Y.; Lv, S.Y.; Cheng, H.; Kang, X.R.; Li, X. Carbon dots as heterojunction transport mediators effectively enhance BiOI/g-C3N4 synergistic persulfate degradation of antibiotics. Appl. Surf. Sci. 2022, 601, 154249. [Google Scholar] [CrossRef]
  29. Raeisi-Kheirabadi, N.; Nezamzadeh-Ejhieh, A. A Z-scheme g-C3N4/Ag3PO4 nanocomposite: Its photocatalytic activity and capability for water splitting. Int. J. Hydrog. Energy 2020, 45, 33381–33395. [Google Scholar] [CrossRef]
  30. Abbasi-Asl, H.; Sabzehmeidani, M.M.; Ghaedi, M. Efficient degradation of metronidazole antibiotic by TiO2/Ag3PO4/g–C3N4 ternary composite photocatalyst in a continuous flow-loop photoreactor. J. Environ. 2021, 9, 105963. [Google Scholar] [CrossRef]
  31. Liu, L.; Qi, Y.H.; Lu, J.R.; Lin, S.L.; An, W.J.; Liang, Y.H.; Cui, W.Q. A stable Ag3PO4@g-C3N4 hybrid core@shell composite with enhanced visible light photocatalytic degradation. Appl. Catal. B Environ. 2016, 183, 133–141. [Google Scholar] [CrossRef]
  32. Ikreedeegh, R.R.; Tahir, M. Facile fabrication of well-designed 2D/2D porous g-C3N4–GO nanocomposite for photocatalytic methane reforming (DRM) with CO2 towards enhanced syngas production under visible light. Fuel 2021, 305, 121558. [Google Scholar] [CrossRef]
  33. He, P.Z.; Song, L.M.; Zhang, S.J.; Wu, X.Q.; Wei, Q.W. Synthesis of g-C3N4/Ag3PO4 heterojunction with enhanced photocatalytic performance. Mater. Res. Bull. 2014, 51, 432–437. [Google Scholar] [CrossRef]
  34. Rosman, N.; Norharyati Wan Salleh, W.; Aqilah Mohd Razali, N.; Nurain Ahmad, S.Z.; Hafiza Ismail, N.; Aziz, F.; Harun, Z.; Fauzi Ismail, A.; Yusof, N. Ibuprofen removal through photocatalytic filtration using antifouling PVDF-ZnO/Ag2CO3/Ag2O nanocomposite membrane. Mater. Today Proc. 2021, 42, 69–74. [Google Scholar] [CrossRef]
  35. Pan, T.D.; Liu, Y.; Li, Z.J.; Fan, J.; Wang, L.; Liu, J.; Shou, W. A Sm-doped Egeria-densa-like ZnO nanowires@PVDF nanofiber membrane for high-efficiency water clean. Sci. Total Environ. 2020, 737, 139818. [Google Scholar] [CrossRef] [PubMed]
  36. Lan, Z.A.; Zhang, G.G.; Wang, X.C. A facile synthesis of Br-modified g-C3N4 semiconductors for photoredox water splitting. Appl. Catal. B Environ. 2016, 192, 116–125. [Google Scholar] [CrossRef]
  37. Li, S.Y.; Zhang, M.; Qu, Z.H.; Cui, X.; Liu, Z.Y.; Piao, C.C.; Li, S.G.; Wang, J.; Song, Y.T. Fabrication of highly active Z-scheme Ag/g-C3N4-Ag-Ag3PO4 (1 1 0) photocatalyst photocatalyst for visible light photocatalytic degradation of levofloxacin with simultaneous hydrogen production. Biochem. Eng. J. 2020, 382, 122394. [Google Scholar] [CrossRef]
  38. Arumugham, T.; Amimodu, R.G.; Kaleekkal, N.J.; Rana, D. Nano CuO/g-C3N4 sheets-based ultrafiltration membrane with enhanced interfacial affinity, antifouling and protein separation performances for water treatment application. J. Environ. Sci. 2019, 82, 57–69. [Google Scholar] [CrossRef]
  39. Shao, L.Q.; Jiang, D.L.; Xiao, P.; Zhu, L.M.; Meng, S.C.; Chen, M. Enhancement of g-C3N4 nanosheets photocatalysis by synergistic interaction of ZnS microsphere and RGO inducing multistep charge transfer. Appl. Catal. B Environ. 2016, 198, 200–210. [Google Scholar] [CrossRef]
  40. Cacho-Bailo, F.; Téllez, C.; Coronas, J. Interactive Thermal Effects on Metal–Organic Framework Polymer Composite Membranes. Chem. Eur. J. 2016, 22, 9533–9536. [Google Scholar] [CrossRef]
  41. Zhang, M.Y.; Liu, Z.Y.; Gao, Y.; Shu, L. Ag modified g-C3N4 composite entrapped PES UF membrane with visible-light-driven photocatalytic antifouling performance. RSC Adv. 2017, 7, 42919–42928. [Google Scholar] [CrossRef] [Green Version]
  42. Seyyed Shahabi, S.; Azizi, N.; Vatanpour, V. Synthesis and characterization of novel g-C3N4 modified thin film nanocomposite reverse osmosis membranes to enhance desalination performance and fouling resistance. Sep. Purif. Technol. 2019, 215, 430–440. [Google Scholar] [CrossRef]
  43. Chen, J.X.; Li, Z.Y.; Wang, C.B.; Wu, H.; Liu, G. Synthesis and characterization of g-C3N4 nanosheet modified polyamide nanofiltration membranes with good permeation and antifouling properties. RSC Adv. 2016, 6, 112148–112157. [Google Scholar] [CrossRef]
  44. Ravichandran, K.; Sindhuja, E. Fabrication of cost effective g-C3N4+Ag activated ZnO photocatalyst in thin film form for enhanced visible light responsive dye degradation. Mater. Chem. Phys. 2019, 221, 203–215. [Google Scholar] [CrossRef]
  45. Mamba, G.; Mishra, A.K. Graphitic carbon nitride (g-C3N4) nanocomposites:A new and exciting generation of visible light driven photocatalysts for environmental pollution remediation. Appl. Catal. B Environ. 2016, 198, 347–377. [Google Scholar] [CrossRef]
  46. Du, Y.E.; Li, W.X.; Bai, Y.; Huangfu, Z.W.; Wang, W.J.; Chai, R.D.; Chen, C.D.; Yang, X.J.; Feng, Q. Facile synthesis of TiO2/Ag3PO4 composites with co-exposed high-energy facets for efficient photodegradation of rhodamine B solution under visible light irradiation. RSC Adv. 2020, 10, 24555–24569. [Google Scholar] [CrossRef]
  47. Lv, J.L.; Dai, K.; Zhang, J.F.; Lu, L.H.; Liang, C.H.; Geng, L.; Wang, Z.L.; Yuan, G.Y.; Zhu, G.P. In situ controllable synthesis of novel surface plasmon resonance-enhanced Ag2WO4/Ag/Bi2MoO6 composite for enhanced and stable visible light photocatalyst. Appl. Surf. Sci. 2017, 391, 507–515. [Google Scholar] [CrossRef]
  48. Liu, W.; Shen, J.; Yang, X.F.; Liu, Q.Q.; Tang, H. Dual Z-scheme g-C3N4/Ag3PO4/Ag2MoO4 ternary composite photocatalyst for solar oxygen evolution from water splitting. Appl. Surf. Sci. 2018, 456, 369–378. [Google Scholar] [CrossRef]
  49. Dong, Z.F.; Wu, Y.; Thirugnanam, N.; Li, G.L. Double Z-scheme ZnO/ZnS/g-C3N4 ternary structure for efficient photocatalytic H2 production. Appl. Surf. Sci. 2018, 430, 293–300. [Google Scholar] [CrossRef]
  50. Wen, J.Q.; Xie, J.; Chen, X.B.; Li, X. A review on g-C3N4-based photocatalysts. Appl. Surf. Sci. 2017, 391, 72–123. [Google Scholar] [CrossRef]
  51. Yu, H.B.; Huang, J.H.; Jiang, L.B.; Leng, L.J.; Yi, K.X.; Zhang, W.; Zhang, C.Y.; Yuan, X.Z. In situ construction of Sn-doped structurally compatible heterojunction with enhanced interfacial electric field for photocatalytic pollutants removal and CO2 reduction. Appl. Catal. B Environ. 2021, 298, 120618. [Google Scholar] [CrossRef]
  52. Méricq, J.P.; Mendret, J.; Brosillon, S.; Faur, C. High performance PVDF-TiO2 membranes for water treatment. Chem. Eng. Sci. 2015, 123, 283–291. [Google Scholar] [CrossRef]
Figure 1. Schematics of nanocomposite synthesis and membrane preparation.
Figure 1. Schematics of nanocomposite synthesis and membrane preparation.
Ijerph 19 15551 g001
Figure 2. Schematic of dead-end stirred ultrafiltration cup setup.
Figure 2. Schematic of dead-end stirred ultrafiltration cup setup.
Ijerph 19 15551 g002
Figure 3. SEM images of (a) g-C3N4, (b) Ag3PO4/TiO2, and (c,d) g-C3N4/Ag3PO4/TiO2.
Figure 3. SEM images of (a) g-C3N4, (b) Ag3PO4/TiO2, and (c,d) g-C3N4/Ag3PO4/TiO2.
Ijerph 19 15551 g003
Figure 4. (a) XRD patterns and (b) PL spectra of g-C3N4, Ag3PO4/TiO2, and g-C3N4/Ag3PO4/TiO2.
Figure 4. (a) XRD patterns and (b) PL spectra of g-C3N4, Ag3PO4/TiO2, and g-C3N4/Ag3PO4/TiO2.
Ijerph 19 15551 g004
Figure 5. SEM images of the (a) top surface and (b) cross-section of the raw PVDF membrane, and (c) top surface and (d) cross-section of the g-C3N4/Ag3PO4/TiO2/PVDF membrane.
Figure 5. SEM images of the (a) top surface and (b) cross-section of the raw PVDF membrane, and (c) top surface and (d) cross-section of the g-C3N4/Ag3PO4/TiO2/PVDF membrane.
Ijerph 19 15551 g005
Figure 6. AFM images of surface of the (a,b) raw PVDF membrane and (c,d) g-C3N4/Ag3PO4/TiO2/PVDF membrane.
Figure 6. AFM images of surface of the (a,b) raw PVDF membrane and (c,d) g-C3N4/Ag3PO4/TiO2/PVDF membrane.
Ijerph 19 15551 g006
Figure 7. (a) FTIR analysis, (b) TGA curves, and (c) dynamic contact angle of the raw PVDF and g-C3N4/Ag3PO4/TiO2/PVDF membranes.
Figure 7. (a) FTIR analysis, (b) TGA curves, and (c) dynamic contact angle of the raw PVDF and g-C3N4/Ag3PO4/TiO2/PVDF membranes.
Ijerph 19 15551 g007
Figure 8. (a) The antifouling property and (b) the ratio of irreversible fouling to reversible fouling of the membranes, and (c) the flux recovery ratio and rejection of BSA (initial BSA concentration of 300 mg/L, pH = 6.9).
Figure 8. (a) The antifouling property and (b) the ratio of irreversible fouling to reversible fouling of the membranes, and (c) the flux recovery ratio and rejection of BSA (initial BSA concentration of 300 mg/L, pH = 6.9).
Ijerph 19 15551 g008
Figure 9. (a) Membrane performance for the photocatalytic degradation of RhB, (b) kinetic model for the photocatalytic degradation of RhB, and (c) changes in the UV–Vis spectra of RhB with reaction time (initial RhB concentration of 10 mg/L, pH = 6.4).
Figure 9. (a) Membrane performance for the photocatalytic degradation of RhB, (b) kinetic model for the photocatalytic degradation of RhB, and (c) changes in the UV–Vis spectra of RhB with reaction time (initial RhB concentration of 10 mg/L, pH = 6.4).
Ijerph 19 15551 g009
Figure 10. (a) Photocatalytic activity and (b) cycling test and reusability of g-C3N4/Ag3PO4/TiO2/PVDF membranes under visible light irradiation for five cycles (initial RhB concentration of 10 mg/L, pH = 6.4).
Figure 10. (a) Photocatalytic activity and (b) cycling test and reusability of g-C3N4/Ag3PO4/TiO2/PVDF membranes under visible light irradiation for five cycles (initial RhB concentration of 10 mg/L, pH = 6.4).
Ijerph 19 15551 g010
Figure 11. Photocatalytic degradation process in the g-C3N4/Ag3PO4/TiO2/PVDF membrane.
Figure 11. Photocatalytic degradation process in the g-C3N4/Ag3PO4/TiO2/PVDF membrane.
Ijerph 19 15551 g011
Table 1. Characteristics of different membranes.
Table 1. Characteristics of different membranes.
MembraneThickness (μm)Porosity (%)Tensile Strength (MPa)
raw membrane176 ± 1970 ± 79.3 ± 1.0
composite membrane186 ± 1785 ± 37.5 ± 1.9
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, R.; Li, X.; Huang, J.; Pang, H.; Wan, Q.; Luo, K.; Pang, Y.; Wang, L. Synthesis and Characterization of g-C3N4/Ag3PO4/TiO2/PVDF Membrane with Remarkable Self-Cleaning Properties for Rhodamine B Removal. Int. J. Environ. Res. Public Health 2022, 19, 15551. https://doi.org/10.3390/ijerph192315551

AMA Style

Liu R, Li X, Huang J, Pang H, Wan Q, Luo K, Pang Y, Wang L. Synthesis and Characterization of g-C3N4/Ag3PO4/TiO2/PVDF Membrane with Remarkable Self-Cleaning Properties for Rhodamine B Removal. International Journal of Environmental Research and Public Health. 2022; 19(23):15551. https://doi.org/10.3390/ijerph192315551

Chicago/Turabian Style

Liu, Renguo, Xue Li, Jinhui Huang, Haoliang Pang, Qiongfang Wan, Kun Luo, Ya Pang, and Lingyu Wang. 2022. "Synthesis and Characterization of g-C3N4/Ag3PO4/TiO2/PVDF Membrane with Remarkable Self-Cleaning Properties for Rhodamine B Removal" International Journal of Environmental Research and Public Health 19, no. 23: 15551. https://doi.org/10.3390/ijerph192315551

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