Next Article in Journal
Novel Graphene-Based Materials as a Tool for Improving Long-Term Storage of Cultural Heritage
Next Article in Special Issue
Development of Pore Pressure in Cementitious Materials under Low Thermal Effects: Evidence from Optimization of Pore Structure by Incorporation of Fly Ash
Previous Article in Journal
Clean H2 Production by Lignin-Assisted Electrolysis in a Polymer Electrolyte Membrane Flow Reactor
Previous Article in Special Issue
Accelerated Carbonation of Vibro-Compacted Porous Concrete for Eco-Friendly Precast Elements
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

An Overview of Photocatalytic Membrane Degradation Development

by
Mojtaba Binazadeh
1,
Jamal Rasouli
1,
Samad Sabbaghi
2,*,
Seyyed Mojtaba Mousavi
3,
Seyyed Alireza Hashemi
4 and
Chin Wei Lai
5,*
1
Department of Chemical Engineering, School of Chemical and Petroleum Engineering, Shiraz University, Shiraz 71557-13876, Iran
2
Department of Nano-Chemical Engineering, Faculty of Advanced Technologies, Shiraz University, Shiraz 71557-13876, Iran
3
Department of Chemical Engineering, National Taiwan University of Science and Technology, Taipei City 106335, Taiwan
4
Nanomaterials and Polymer Nanocomposites Laboratory, School of Engineering, University of British Columbia, Kelowna, BC V1V 1V7, Canada
5
Nanotechnology & Catalysis Research Centre, University Malaya, Kuala Lumpur 50603, Malaysia
*
Authors to whom correspondence should be addressed.
Materials 2023, 16(9), 3526; https://doi.org/10.3390/ma16093526
Submission received: 3 February 2023 / Revised: 9 March 2023 / Accepted: 27 March 2023 / Published: 4 May 2023
(This article belongs to the Special Issue Transforming Industrial Waste into Sustainable Construction Materials)

Abstract

:
Environmental pollution has become a worldwide issue. Rapid industrial and agricultural practices have increased organic contaminants in water supplies. Hence, many strategies have been developed to address this concern. In order to supply clean water for various applications, high-performance treatment technology is required to effectively remove organic and inorganic contaminants. Utilizing photocatalytic membrane reactors (PMRs) has shown promise as a viable alternative process in the water and wastewater industry due to its efficiency, low cost, simplicity, and low environmental impact. PMRs are commonly categorized into two main categories: those with the photocatalyst suspended in solution and those with the photocatalyst immobilized in/on a membrane. Herein, the working and fouling mechanisms in PMRs membranes are investigated; the interplay of fouling and photocatalytic activity and the development of fouling prevention strategies are elucidated; and the significance of photocatalysis in membrane fouling mechanisms such as pore plugging and cake layering is thoroughly explored.

1. Introduction

Industrial organic pollutants such as hormones, pesticides, pharmaceutically active substances, and personal care products have become a growing environmental concern due to their continuous release into the waterways and subsequent detrimental effects on human health, plants, soil, and aquatic systems [1,2,3,4]. Over a long period, antibiotics in their active form (low doses) and their main components can alter microbial communities in bodies of water [5,6]. Studying water samples of wastewater treatment plants in Beijing, China, revealed high concentrations of tetracycline, sulfonamides, and quinolones [7].
To remove contaminants and lessen their adverse effects on the environment and human health, a number of purification approaches are available, such as chemical coagulation [8], biodegradation [9], flotation [10,11], absorption [12], and adsorption [13,14]. Despite their benefits, conventional procedures have drawbacks, such as their inability to entirely remove contaminants. For instance, there are a number of drawbacks to the adsorption technique used for wastewater treatment, such as the challenging nature of separating the adsorbent from the solution and the high cost of the adsorbent [14,15]. Therefore, it is crucial to develop a novel, incredibly efficient method to eliminate contaminants from wastewater before they enter the environment [16].
The development of innovative green chemical technologies and methods in organic synthesis and environmental converters has emerged as one of the most pressing topics for chemical researchers in recent decades, particularly those working with heterogeneous photocatalysis (HPC). Too much emphasis is placed on exploiting the sun as a renewable energy source, and visible light photocatalysts play a critical role in this regard [17,18]. Progress in photocatalysis, an improved oxidation process, has been rapid in recent years. This technology has gained a reputation for being sustainable and efficient at reducing energy use and harming the environment [19,20]. The membrane separation process is also one of the most common and widely used methods for treating contaminated water. Membrane technology is simple to operate, requires little space, and allows product recovery at high efficiency and high selectivity. Membrane separation is done at normal operating temperatures and does not require phase change [21,22]. Membranes are classified into different types depending on the material, structure, geometry, manufacturing method, and type of process used. Organic, inorganic, and composite materials may be used for production of flat, hollow fiber, or tubular membranes [23,24,25,26]. Based on the substance employed, membrane production is primarily divided into organic (polymeric) and inorganic membranes [27,28]. Inorganic membranes are those that incorporate metals, oxides, or elementary carbon in their structure, such as zeolite, ceramic, and metallic membranes, while organic membranes are those constructed of nonporous polymeric materials, including polyethersulfone, polysulfone, polymethylpentene, cellulose acetate, polycarbonate, polyimide, polyetherimide, polydimethylsiloxane, and polyphenylene oxide [27,29,30]. The separation techniques used in these systems are usually vacuumed filtration or pressure filtration using microfiltration, ultrafiltration, or nanofiltration membranes [31,32]. Polymeric synthetic membranes make up a sizable portion of industrial membranes. Compared to mineral membranes, these membranes are more commonly used because as they are easier to build, less expensive, and more processable [33].
Combining photocatalysis and membrane filtration is a widely used method that can be effective for the removal various contaminants from the aqueous system. Therefore, the design and development of photocatalytic membrane reactors (PMRs) have become a popular and effective method for treating polluted water. The number of articles published annually on PMRs is depicted in Figure 1. PMRs differs from conventional photocatalytic reactors as the photocatalyst remains in the system and treated water selectively permeates through the membrane. PMRs enhance process efficiency, controllability, and stability. They also minimize installation space, energy requirement, and eliminate extra expenditures associated with flocculation, coagulation, and sedimentation [34,35]. The photocatalyst in the PMR system can photodegrade pollutant molecules stuck to the membrane surface using visible light, sunshine energy, and UV radiation [36].
Both polymeric and mineral membranes have shown excellent self-cleaning and anti-fouling properties through photodegradation when exposed to visible light and UV irradiation [37,38]. A photocatalytic membrane may utilize the absorbed energy from the irradiation source and breaks down pollutants that adhere to its surface; thus, such a membrane is self-cleaning and anti-fouling [38].
The main focus of this research is how to use PMRs to clean water and waste water. The photocatalyst, membrane, and light source are the main components of PMRs that are hereby discussed in depth. Membrane fouling is a serious operational concern in slurry-type reactors; thus, its effects on permeate flux and system efficiency are described in length, along with solutions for mitigating the impacts of membrane fouling. Several of the most crucial operational variables that affect PMRs performance are covered. The standards for creating and developing PMRs are provided. Lastly, the most recent developments in the utilization of visible light are also discussed, along with efforts to circumvent some of the inherent drawbacks of PMRs, such as a moderate loss of photoactivity, constrained processing speeds due to mass transfer issues, membrane leakage, and unsatisfactory system lifetimes due to photocatalyst leaching.

2. Mechanism of Photocatalysis and Membrane Processes

2.1. Photocatalytic Degradation Mechanism

As schematically shown in Figure 2a, degradation reactions are driven by the electrons transferred from the valence to the conduction band. Typically, a photocatalyst’s band-gap energy ( E a ) should be equal to or less than the emitted photon energy [39,40,41]. Electron transfer results in formation of an associated hole ( h VB + ) in the valence band [42,43]. Electron-hole pairs promote both oxidation and reduction of the adsorbed layer by generation of radicals [17,44]. Radicals are active oxidizing and reducing species that attack to and degrade contaminants in the aqueous solution [45,46]. A substrate reduction potential below that of hole ( h VB + ) results in the oxidation of substrates in the valence band, whereas a substrate reduction potential higher than that of electron ( e CB ) results in the reduction of substrates in the conduction band [47,48]. Hydroxyl radicals ( OH ) are the main species responsible for the photodegradation of pollutants. Another oxidant is reactive oxygen species (ROS), such as superoxide oxygen radicals ( O 2 ) [47]. Figure 2b shows photocatalytic degradation of cefixime, as an example of pharmaceutical pollutants, by Fe2O3@TiO2 [13].

2.2. Mechanism of the Membrane Filtration Process

Pressure, concentration, or electric potential differences are the driving forces of membrane separation techniques. PMRs operate by pressure and concentration differences. Generally, surface water recycling for non-drinking purposes involves the utilization of ultrafiltration (UF) and microfiltration (MF) membranes. Nanofiltration membranes (NF) with molecular weight cut-offs (MWCO) between 150 and 350 Da are suitable for wastewater treatment and environmental cleanups due to their excellent inorganic ion removal capabilities [21].
The contaminated feed solution passes through the membrane under pressure. Membrane in a PMR retains the photocatalyst and pollutants and preferentially permeates water [49,50,51] as depicted in Figure 3a. In photocatalysis conditions at PMRs, polymeris membranes are highly susceptible to (1) abrasion by the photocatalyst and (2) degradation assisted by hydroxyl radicals. Thus, inorganic membranes are preferable due to their superior chemical, mechanical, and thermal resistance. One drawback of inorganic membranes is their cost. A thinner membrane enhances permeate flux due to its lower hydraulic resistance. However, if the membrane is too thin, it may be vulnerable to damage. Average pore size diameter, chemical, mechanical and thermal stability, cost, and lifetime are other important parameters that must be tailored for each membrane application. A list of membranes used for water decontamination is reported in Table 1.
As depicted in Figure 3 there are four main solute transfer mechanisms in membranes. The main transport mechanism for a membrane process is determined by relative magnitude of (1) average pore size diameter (d), (2) size difference of transferable and nontransferable molecules, (3) mean free path of transferable molecule (λ), and (4) pore network structure. According to Figure 3b, presence of large pores in the membrane (d/λ > 20) results in convective flux. Convective flux through membrane accelerates the process; however, convection mechanism is only applicable in PMRs when there is a large size difference between contaminants/photocatalyst and water molecules. If such size difference does not exist, membranes with smaller pore size are required which transport water molecules by permeation. The permeation rate is affected by various factors, including pressure and concentration gradient, size and shape of permeate, and pore size, thickness and chemical structure of membrane. There are two main forces that drive the permeation: hydraulic pressure, which pushes solvent molecules through the membrane, and osmotic pressure, which opposes the flow of solvent due to the presence of dissolved species in the wastewater [60]. The Knudsen diffusion mechanism is dominant when d/λ < 0.2 [61]. Knudsen diffusion occurs only during gas transfer across nano porous membrane; thus, it is useful in removing gases and volatile organic compounds from wastewater [62,63]. Molecular sieves may also be used to purify water [64]. Solution–diffusion is the main transport in dense membranes. Solution–diffusion does not apply to water purification; however, it is the main mechanism for hydrogen transport in Pd-based membranes [65].
Figure 3. Separation mechanisms in membrane: (a) typical membrane, (b) convective flow, (c) convection–permeation, (d) Knudsen diffusion, (e) molecular sieving and (f) solution–diffusion [65]. Adapted with permission from Elsevier. Copyright 2021.
Figure 3. Separation mechanisms in membrane: (a) typical membrane, (b) convective flow, (c) convection–permeation, (d) Knudsen diffusion, (e) molecular sieving and (f) solution–diffusion [65]. Adapted with permission from Elsevier. Copyright 2021.
Materials 16 03526 g003
During the filtration process, pollutants adhere to the membrane surface and decrease membrane permeability. This phenomenon is calling membrane fouling which is one of the most important challenges of membrane processes [66,67].
Adsorption, accumulation, and precipitation are three mechanisms that may simultaneously occur to produce fouling [68,69]. To improve membrane performance, its surface is modified to maximize its affinity towards the permeating solvent and minimize its affinity towards the fouling agents [70]. Chemical modification, UV irradiation, applied electric field, aeration, and plasma treatment may be utilized to tune hydrophilicity of the membrane surface [71]. Filtration at critical flux has a modest flux drop and minimized irreversible fouling; however, it reduces output. Finally, to maintain the performance of the membrane system, cleaning and maintenance strategies such as backwashing and chemical cleaning are employed routinely during operation [72,73].

Characterization of Membranes

Membrane production procedure may include phase inversion, interfacial polymerization, and coating. The casting solution and cooling bath utilized in the phase inversion process have a substantial effect on the tortuosity, pore size distribution, pore network. morphology, microstructure, and mechanical properties. Fouling, conditioning, chemical exposure, disintegration, cleaning, the aging process can result in reversible or irreversible alterations to the physical and chemical properties of membranes; thus, membrane characterization is an essential part of membrane manufacturing and maintenance [74]. Characterization techniques such as XRD, SEM, TEM, TGA, DSC, BET, Zeta potential analysis, and FTIR may be employed. SEM images in Figure 4a,b show cross-section PES membrane and irregular porous surface of PVDF membrane, respectively [75,76]. TEM image of Figure 4c shows components of a hierarchical layer of a TiO2 nanowire membrane [77].

3. Configurations of PMRs

A typical PMR includes a light source, membrane, and photocatalyst [78]. The design and configuration of the PMR dictates process efficiency and controllability. PMRs are classified into two types based on the photocatalyst loading. As schematically illustrated in Figure 5 and Figure 6, suspended photocatalytic membrane reactors (SPMR) are those in which the photocatalyst is suspended, whereas immobilized photocatalytic membrane reactors (IPMR) are those in which the photocatalyst is fixed on a carrier material such as quartz, stainless steel, glass, limestone, or zeolite [79]. When a photocatalyst is immobilized on a support, the active surface available to solution particles is drastically decreased, resulting in a loss of photoactivity [80]. The active surface increases significantly when the photocatalyst is suspended; however, after detoxification, the photocatalyst particles must be separated from the treated water. Table 2 shows the numerous applications of PMRs in wastewater treatment processes.
Benefits of PMRs over traditional photoreactors include: (1) the ability to regulate the residence inside the reactor; (2) continuous operation; (3) the containment of the contaminants and photocatalyst within the reaction environment; (4) enhancing process efficiency and stability; and (5) reduced reactor volume and operating costs [36,42,81].
Table 2. Application of PMRS in wastewater treatment process.
Table 2. Application of PMRS in wastewater treatment process.
PhotocatalystPollutantType of PMRPhotocatalyst Dosage (wt%)Pollutant Concentration (mg·L−1)Light SourceTime (min)Degradation (%)Ref.
P-doped g-C3N4 (PCN) and coated on an Al2O3 substratephenol and methyl blue 10-visible-92 and 90[82]
MIL-88B(Fe) and coated onto an Al2O3 substratephenol --visible--[83]
immobilized N-doped TiO2diclofenacSPMR--visible--[84]
MIL-53(Fe)/PVDF mixed-matrix membranetetracyclineIPMR5-UV-93[85]
UVA/TiO2-MFoxytetracyclineSuspended vs immobilized TiO2-P25-5visible30>90[86]
polysulfone/H2O2-g-C3N4 mixed matrix membranehumic acid-10-visible-93.5[87]
NH2-MIL125(Ti) MOFmethyl blueimmobilized and suspended2-UV-60 and 97[88]
TiO2ketoprofenSPMR-10 -61[79]
TiO2-WO3/PANICr (VI)SPMR5-Visible6098.5[89]
TiO2/UV-AnitrateSPMR--UV-65–90[90]
TiO2ketoprofenSPMR-10--75[79]
Sb2O3/CuBi2O4methyl blueSPMR1010Visible 94.6[91]
ZnO/WO3phenolSPMR-30 792.5[92]

4. Photocatalytic Degradation of Pollutants

Many organic and inorganic substances, especially toxic or refractory substances are resistant to biological degradation. After the discovery of the photocatalytic splitting of water in 1972 by Fujishima and Honda [93], scientists and researchers turned their attention to semiconductor photocatalysts that could destroy resistant contaminants that were difficult or impossible to remove by other methods [94]. Photocatalytic degradation has emerged as one of the most sustainable, energy efficient, cost-effective and non-hazardous and environmentally friendly processes for contaminants removal from water which uses light as the energy source [95]. The catalyst’s photonic activation mode, which replaces thermal activation, is the primary distinction between photocatalysis and traditional catalysis [96]. Photocatalysts do not contain heavy metal and do not require strong oxidants/reducing agents for activation. Photocatalysis degradation products are harmless [47].

4.1. Photocatalytic Degradation of Pharmaceutical Compounds

pH is one of the most vital factors that affects photodegradation efficiency. The efficiency of TiO2 and ZnO nanoparticles in removing acetaminophen from the water was studies by Ahed et al. [97]. The outcomes of the study demonstrated that ZnO was more effective at eliminating acetaminophen. Ahed et al. proved that pH is an important factor in photodegradation. Their synthesized ZnO photodegraded 97% of acetaminophen at pH = 9 in 1 h of exposure. They demonstrated that at neutral pHs, photocatalytic degradation rates are higher than at acidic pHs. Sabouni et al. [98] investigated the elimination of progesterone, ibuprofen, and naproxen using ZnO photocatalyst. They studied the initial concentration of pollutants and the photocatalyst loading. They found that ZnO photocatalyst is quite efficient in eliminating all three pollutants. They reported progesterone, ibuprofen, and naproxen as having a degradation efficiency of 92.3%, 94.5%, and 98.7%, respectively. The degradation of paracetamol using TiO2 and Fe2O3 photocatalysts was studied by Abdelwahab et al. SEM, TEM, XRD, FTIR, Raman spectroscopy, and VSM analyses were used to characterize the synoecized photocatalysts. Their results revealed that paracetamol degradation increased when TiO2 loading in the TiO2/Fe2O3 composite was increased [99]. The FTIR spectra of iron glycolate, Fe2O3, and 50% TiO2/Fe2O3 reported by Abdelwahab et al. are displayed in Figure 7a. The peaks at 3432 cm−1 and 1618 cm−1 are attributed to the O-H stretching and bending vibrations of adsorbed water or EG, respectively. Peaks between 2850–2950 cm−1 are indicative of C-H vibrations. Between 1120 and 1470 cm−1, the CH2 bending vibration peaks could be seen. C-O stretching vibrations were responsible for two sharp peaks at about 1050 and 1085 cm−1. Peaks of the Fe-O stretching vibration were seen at the 467–660 cm−1. All of the peaks related to the glycolate moiety have disappeared in the Fe2O3 FTIR spectrum, which was obtained after the iron glycolate was calcined at 350 °C [99].
To examine the crystalline structure and phase purity of the prepared sample, the XRD patterns photocatalysts synthesized by Abdelwahab et al. are reported at Figure 7b. When the iron glycolate sheets are stacked, a prominent low-angle diffraction peak at 11° is observed. The diffraction pattern of the Fe2O3 product created by calcining the iron glycolate at 350 °C shows that the main diffraction peaks match with the α-Fe2O3 rhombohedral structural pattern. However, two peaks at 2θ of 30 and 43° that are either indicative of Fe3O4 or γ-Fe2O3 show that the acquired sample could be made up of α-Fe2O3 and other iron oxide crystal forms. the relative magnitude hematite (α-Fe2O3) and maghemite (γ-Fe2O3) and/or magnetite (Fe3O4) are 70% and 30%, respectively. Furthermore, the new peaks at 2θ of 25.2, 37.4, 48, and 54° of the TiO2/Fe2O3 samples can be correlated to the (101), (004), (200), and (105) planes of anatase TiO2, demonstrating that the TiO2 crystallites were placed onto the magnetic core Fe2O3. The average crystallite size of TiO2, calculated by Scherrer’s equation is 12 nm for the anatase (101) peak [99].
Table 3 shows the photocatalytic degradation of different pharmaceuticals. Clearly, heterogeneous photocatalysts with a variety of nanostructures can effectively degrade pharmaceuticals in aqueous solutions.

4.2. Photocatalytic Degradation of Dye Compounds

Dyes are routinely used in the textile, food, beverage, printing, and pharmaceutical industries. These compounds may be harmful and cancerous will barricade sunlight from reaching water bodies, impacting natural aquatic processes such as photosynthesis and other biodegradation operations [112,113]. Long-lasting and non-degradable colored pollutants must be removed before entering the environment since their entry leads to the aquatic ecosystem becoming toxic and dangerous to humans [114]. The photocatalytic process has been suggested as one of the more successful methods. In photocatalysis, the degradation is begun with OH radicals breaking the azo bond (-N=N-), which is one of the weakest chemical bonds in the dye molecules’ chemical structure [115]. The process’s intermediates will then undergo a radical chain reaction with the oxygen molecules, finally breaking down to produce water and carbon dioxide [115]. In recent years, numerous studies on the removal of contaminants from actual wastewater have been conducted. Table 4 illustrates the photocatalytic degradation of a variety of dyes.

4.3. Photocatalytic Degradation of Hydrocarbons

Industrial use of hydrocarbons inevitably contaminates natural waters through improper disposal or leaching from landfills, spills, or leaks in underground pipes. Hydrocarbon contaminants endanger human health if they enter drinking water. The presence of these contaminants in water hinders light penetration into the water and affects the diffusion/solubility of gases required for aquatic plant respiration which ultimately leads to plant death; therefore, it may affect the food supply chain.
Schnabel et al. [127] used semiconductor titanium dioxide to remove hydrocarbons. This study demonstrated that a variety of photocatalyst designs, when exposed to ultraviolet (UV) light, can remove the non-polar material in diesel fuel. They reported that the floating foam glass catalyst with TiO2 coating reduces the concentration from an initial concentration of 668 mg/L to 329 mg/L in 16 h. The contaminant concentration is reduced by 401 mg/L and 55 mg/L when glass fiber and steel grit was used, respectively [127].
Nirmala Rani et al. [128] studied the elimination of three polycyclic aromatic hydrocarbons (PAHs) in mixed or separate states using titanium oxide photocatalysts and MS membranes under UV irradiation in a PMR. They reported the degradation efficiency of 100, 94.1, and 97% in an aqueous mixture containing 1000 µg /L phenanthrene (PHE), 5000 µg/L naphthalene (NAP), and 1000 g/L acenaphthene (ANA), respectively, after 180 min UV irradiation at photocatalyst loading of 0.5 g/L. When the compounds were used as a sole compound, the elimination percentages of PHE, NAP, and ANA were 99.3, 92.8, and 95.3, respectively, under similar operating conditions [129]. The photocatalytic degradation of several hydrocarbons is reported in Table 5.

4.4. Photocatalytic Degradation of other Pollutants

Photocatalyst have been used to degrade pollutants listed in Table 6. Pitchaimani Veera Kumar et al. [139] employed zinc oxide nano stars (ZnONSt) coupled with Ag and Pd to photocatalytically degrade herbicides and pesticides. Ag at ZnONSt and Pd at ZnONSt photocatalysts accelerated the degradation of existing pollutants as they facilitate the interfacial charge transfer process [139]. In order to remove the herbicide ametrine, Rodrigo Pereira Cavalcante et al. [140] utilized a titanium dioxide photocatalyst. They were able to completely remove the ametrine after 60 min irradiation of simulated sunlight (using a 1000 W Xenon lamp), and then they used 0.4 g/L of photocatalyst to detoxify the solution. Samsudin et al. [141] utilized BiVO4/g-C3N4 integrated with Pt to purify poultry during sun light exposure. The as-synthesized photocatalyst demonstrated 93.5% COD removal from the starting concentration of 2152 mg/L in 3 h. The photocatalyst showed strong recyclability and photostabiliy.

5. PMRs

5.1. Operating Factors and Limits of PMRs

5.1.1. Operating Mode

Both dead-end and cross-flow PMRs can be used in photocatalytic systems. In the dead-end configuration, the whole stream is filtered by passing through a membrane (permeate). As a consequence, the concentration of the nontransferable components rises, resulting in creation of a filter cake on a membrane’s surface, as well as a reduction in membrane permeability and photocatalytic efficiency. In the absence of turbulency, e.g., stirring, there will be insufficient contact between the contaminants, photocatalyst, and the light source [152]. Wang et al. [153] investigated a novel photocatalytic membrane created via pressure-driven filtration load with a ZnO/N-g-C3N4 composite via glutaraldehyde as a crosslinker. SEM, XPS, and FTIR were employed to verify the photocatalyst loading on membrane surface. They found that in both immersion and filtration models, the photocatalytic capabilities of the ZnO/N-g-C3N4 composite membrane were effective for the decomposition of tetracycline, ofloxacin, and ciprofloxacin under visible light (>420 nm). For tetracycline at 5 mg/L and 10 mg/L concentrations a ZnO/N-g-C3N4 loading of 1.12 g/cm2 resulted in 100% and 80% degradation, respectively. They concluded that a prolonged reaction time on the membrane surface, low trans-membrane pressure (0.005 MPa), and narrow membrane size were advantageous for the elimination of antibiotics in the filtration processes.

5.1.2. Photocatalyst Type and Characteristics

Key parameters that significantly affect photocatalytic efficiency include the type of photocatalyst, the photocatalyst’s physicochemical properties (band gap energy, particle size distribution, crystallographic structure, and chemical makeup), and the photocatalyst’s concentration in the reacting environment. As mentioned previously, photons by energy equal to or greater than the band gap energy can be absorbed in photocatalytic activities, resulting in the creation of electron-hole pairs. However, photocatalysts that require visible light to function are more intriguing, and it has become a challenge for PMR systems to use a light source that is both environmentally friendly and economically viable [154]. For example, TiO2-supported photocatalyst is the most frequently employed in PMR in suspended form due to its great photochemical stability in aquatic solutions, robust catalytic activity, reasonably long lifespan of electron-hole pairs, low cost, and low toxicity. This material is inactive when exposed to visible light. As a result, TiO2 can only absorb around 5% of the solar radiation that is in the UV spectrum [155]. Ahmad et al. [156]. stated that a developed composite ceramic membrane may benefit from a synergy of dead-end filtration and cross-flow filtration while being subjected to intermittent UV irradiation in order to efficiently prevent membrane fouling. To remove organic dye impurities in a photocatalytic membrane reactor, a partly coated TiO2 (pc-TiO2) layer was made with the assistance of cheap polyvinyl chloride (PVC) to make gaps in a porous Al2O3 membrane substrate. Their study revealed that the pc-TiO2/Al2O3 composite membrane has superior water flux and anti-fouling capabilities compared to the uniformly coated TiO2/Al2O3 (UC-TiO2/Al2O3) membrane. The photocatalytic activity of UC-and pc-TiO2/Al2O3 composite membranes was significantly enhanced during cross-flow membrane filtration in comparison to that of the Al2O3 bare membrane substrate.

5.1.3. Light Source

When light is shone on photocatalysts, photons with energy higher than or equal to the band gap are absorbed, valence band electrons are shifted to the conduction band. The oxidation and reduction reactions that occur are due to the production of electron-hole pairs [4]. Consequently, the kind and intensity of light have a significant impact on the performance of photocatalysis [157].
The sol-gel procedure and the dip-coating may be used to create a nanostructured TiO2 film from titanium tetraisopropoxide. Sol-gel nanostructured TiO2 (anatase phase) film was investigated for its photocatalytic degradation of azithromycin to determine the most efficient degradation pathway for application in wastewater treatment. At the pH of 10 and UV-C irradiation maximum degradation was achieved. The LED irradiation source with emission wavelength of 365 nm was not as efficient as the UV-C lamp. The LED bulb, however, may be a “real-world” option due to its low price, high energy efficiency, and low environmental impact [158]. Shang et al. [159]. investigated the antibacterial activities of TiO2 photocatalysts under various light sources, exclusively under visible light. They discovered that by doping metal ions and nonmetal ions on TiO2 and compounding with polymers, they could increase the photocatalytic activity to the visible light region, improve the surface characteristics, and enhance the contact area with bacteria. Reactive oxygen species (ROS) and hydroxyl free radicals damage the cell membrane, DNA, and enzymes.

5.2. Degradation of Pharmaceutical Compounds via PMR

Pharmaceutical compounds are structurally complex and environmentally stable. They typically contain abundant aromatic rings. Thus, conventional wastewater treatment methods cannot effectively remove them. Hence, the use of PMRs has become a popular solution for removal of pharmaceutical compounds from water. Fang et al. [153] synthesized a ZnO/N-g-C3N4 composite and immobilized it on a commercial polymer membrane via GA as the crosslinker to breakdown antibiotics under visible light in a PMR setup. Additionally, the ZnO/N-g-C3N4 composite photocatalytic membrane properties were evaluated using the immersion model and the filtration model. The amount of ZnO/N-g-C3N4 loading and GA concentration were significant for the photocatalytic abilities of composite membranes, according to an immersion model. The outcome demonstrated that the filtering model results in a greater antibiotic decomposition at longer photocatalytic reaction time, that could be attained by narrow membrane pore sizes, low TMP, and reduced flow. Using a similar photocatalytic membrane fabrication procedure using pristine membrane with MWCO of 50 kDa, improved TC degradation (71.7%) was obtained by the immersion approach which is attributed to the decreased flux and increased retention time. The Photocatalytic membrane degradation of pharmaceutical compounds is shown in Table 7.

5.3. Degradation of Dye Compounds via PMR

Synthetic paint is one of the most abundant pollutants in sewage and effluent of industrial plants. The most commonly investigated dyes are rhodamine B, methylene blue, and methyl orange. Conventional wastewater treatment methods are inefficient in removing dyes. Photocatalytic technology is one of the most successful approaches suggested for dye removal.
Dzinun et al. [180] create a TiO2-PVDF photocatalytic membrane by addition of different loading of TiO2 nanoparticles on a PVDF membrane for methylene blue removal. They characterized the resulting photocatalytic membranes with FE-SEM, EDS, and AFM. The results of their custom-designed PMR demonstrated that adding TiO2 nanoparticles to the PVDF membrane speeds up methylene blue removal from waste water. Kolesnyk et al. [181] studied the impact of g-C3N4 loading on a commercial PVDF membrane on rhodamine-B removal. The highest removal rate occurred in the alkaline medium. In this study, the membrane lost approximately 15% of its pores after five cycles (a total of 50 h). Yu et al. [163] synthesized a PSF membrane coated with g-C3N4 and TiO2 nanocomposites to investigate the removal of sulfamethoxazole. They reported that sulfamethoxazole was converted into seven other non-toxic substances using their custom-designed PMR with sun light irradiation [163]. Horowitz et al. [179] investigated the efficiency of Al2O3 coated membranes with pore sizes of 200 and 800 nm for carbamazepine removal at different operating conditions. It was observed that the PMR efficiency under UV irradiation is much higher than that under visible waves. They also reported that the contaminant removal rate increases with temperature. Ma et al. [182] investigated humic acid removal from wastewater using TiO2/Al2O3 photocatalyst and membrane microfiltration processes. They found that light intensity significantly affects humic acid removal. A summary of dye degradation via PMRs is reported in Table 8.

5.4. Degradation of Hydrocarbons via PMR

Rani et al. used a membrane photocatalytic reactor containing suspended TiO2 photocatalytic particles for naphthalene removal. They studied the impact of different operating parameters such as the initial concentration of naphthalene (5–25 mg/L), photocatalyst loading (0.1–0.9 g/L), and pH (3–9) on naphthalene removal rate. The maximum naphthalene removal by separate photocatalysis and membrane was 76.8% and 49.1%, respectively, while a naphthalene removal of 90.2%. could be achieved by PMR [183]. Batch PMRs functioning in the dead-end mode was designed by Moslehyani et al. [184] which could to eliminate 99% of the hydrocarbons from sludge after 2 h. Ag-TiO2-coated alumina membrane in a dead-end configuration to degrade rhodamine rate of 1.007 mg/m2h1 [185]. Despite the promising outcomes, the researchers emphasized that the dead-end process leads to the buildup of separated substrates on the membrane surface and ultimately forms a cake layer, which decreases photocatalytic efficiency. A summary of hydrocarbon degradation via PMRs is reported in Table 9.

5.5. Degradation of Other Pollutants via PMR

Pollutants such as toxins, detergents, and heavy metals that do not fall into the above categories are also treated by PMRs which can be seen in Table 10. For example, it can be seen from Table 10, 99.9% of toxic hexavalent chromium (Cr (VI)) at an initial concentration of 10 mg/L could be removed by PMR containing Chitosan-sodium alginate/Fe-doped WO3 photocatalyst and PES membrane after 240 min using a 300 W Xe light source [186].
Table 8. Review on Photocatalytic Membrane degradation of dye compounds.
Table 8. Review on Photocatalytic Membrane degradation of dye compounds.
PollutantPollutant Concentration (mg/L)Photocatalyst/Synthesis MethodMembrane/Pore Size (µm)Light SourceTime (min)Degradation (%)Ref.
Phenol50TiO2-GO/modified Hummer’sPVDF, PAA/0.45100W UV-C lamp 60[187]
RhodamineB10CNTs/MCU-C3N4/GOPVDF300 W Xe lamp, 98.31[188]
RhodamineB-TiO2PVDF/0.08–0.23 UV-C lamps 95[189]
Eosin yellow100N,Pd co-doped TiO2/Polysulfone500 W Xenon lamp18092[190]
Phenol5g-C3N4/CNTsAl2O3/0.297300 W Xe lamp6094[191]
Methylene blue1TiO2Al2O3/0.02–0.2UV-LED 80[192]
Methylene blue-ZnWO4/NiAl-LDH/HydrothermalPVDF/0.45 for pure PVDF 12093.97[193]
Methylene blue500Co/PC/g-C3N4 Xe Lamp 300 W36099[194]
Methylene blue1NbCxOy/NbOx/g-C3N4--480100[195]
RhodamineB10TMPyP/SPSf/non-solvent-induced phase separationPES300 W Xe Lamp18093.4[196]
Methylene blue10RGO/PDA/TiO2/ Hummer methodCA/0.3–0.52 15080[197]
Phenol10O-g-C3N4PES/0.125–0.18830W UV lamp,12035.78[198]
RB530F e 3 + doped ZnO-artificial sunlight (D65, 72W)18098.34[199]
Methylene blue, RhodamineB, and methyl Orange20TiO2/ hydrothermalPPS/0.185300 W Xenon lamp90RhodamineB 99.56,
methylene blue 98.05,
methyl Orange 93.18
[200]
Methyl orange10meso-TiO2/PVDF25 W UV lamps720Higher than 90[201]
Methylene blue30PDA/RGO/Ag3PO4/PVDF200 W incandescent lamp 99.1[202]
Methyl orange7.8TiO2 nanoparticles/Dip-CoatingAl2O3/0.125300 W high-pressure mercury UV tube 61.2[203]
RhodamineB8MWCNTs/Ag3PO4/combining electrospinning with in situ
Ag3PO4 forming reaction
PAN300 W Xe arc lamp12096.9[204]
Methylene blue
4-CP
15reduced graphene oxide
(RGO)/poly(dopamine) (PDA)/Bi12O17Cl2/
CA/0.22500 W long-arc
Xe lamp,
160methylene blue 98,
4-CP 96
[205]
Methylene blue10TiO2 nanoparticles/Electrospraying TiO2
particles
Polyamide-6
nanofiber
300 W Osram Ultra-Vitalux lamp36099[206]
Methylene blue-TiO2/Magnetron sputteringPES/0.45 16070[207]
Congo red-Fe-doped ZnO/rGO/Sol-gelNFsolar radiation 87[208]
Methylene blue3.2Graphene oxidePVDF/0.1–0.7150 W xenon lamp 83.3[209]
Rhodamine 6G, Rhodamine B5 to 50Graphene oxide/direct heating of melaminePVDFLED lamp Rhodamine B 96,
Rhodamine 6G 94
[181]
Methylene blue1PdTFPPPVDF/0.24.6 W Green and white light emitting diode 83[210]
Rhodamine B10g-C3N4/RGO/photoreductionCA/~0.430Xe lamp9090[211]
Methylene blue50nitrogen-doped graphene/TiO2/nonsolvent-induced phase-separationPSF125W UV lamp,
100 W
fluorescent bulb
12094.6[212]
Remazol black B50g-PAA/ZnOPVDF/0.4515 W UV lamp30086[213]
Rhodamine B5Bi2O3/ZnS/CA/0.11–0.14200 W xenon light120~85[214]
Acid orange 750SrTiO3/TiO2/ hydrothermalCA/0.2UV lamp 100[215]
Methyl blue,
phenol
methyl blue 5
phenol 3.3
Phosphorus-doped g-C3N4/ thermal condensationAl2O3/0.18–0.2300 W Xe lamp Phenol 92,
methyl blue 90
[82]
Table 9. Review on Photocatalytic Membrane degradation of hydrocarbons.
Table 9. Review on Photocatalytic Membrane degradation of hydrocarbons.
PollutantPollutant Concentration (mg/L)Photocatalyst/Synthesis MethodMembrane/Pore Size (µm)Light SourceTime (min)Degradation (%)Ref.
Acenaphthene1–3TiO2PES16 W Hg UV-C lamp Batch process 95.1,
Continuous process 80
[216]
Crude oil100TiO2, BiVO4, WO3/hydrothermalPVDF/0.114.4W LED strip 89[217]
Roxarsone BiOC l 0.875 B r 0.125 PVDF 100[218]
propranolol2TiO2/rGO-TiF0.160–0.175UV lamp6035[219]
Bisphenol10Ag-doped TiO2/ Liquid impregnation—phase inversionPESf100 W Xe lamp27088[220]
Naphthalene5–25TiO2PES16 W Hg UV-C lamp180batch process 92.8, continuous process 93.1[183]
Table 10. Review on Photocatalytic Membrane degradation of other Pollutants.
Table 10. Review on Photocatalytic Membrane degradation of other Pollutants.
PollutantPollutant Concentration (mg/L)Photocatalyst/Synthesis MethodMembrane/Pore Size (µm)Setup ConfigurationLight SourceTime (min)Degradation (%)Ref.
BSA-SrTiO3–CrPVDF0.2–1 μmUVA light Higher than 98[221]
Oil in water emulsions-GO/MCU-C3N4PVDF0.22 30Higher than 96[222]
Humic acid20 Alumina-supported titania/0.657–0.425 mercury lamp with light emission of
1255 μW/cm2
180 nin92[223]
Oily wastewater500 and 1000TiO2Al2O3/Ceramic membrane-UV lamp 90[224]
Nitrate10LiNbO3/phase inversionPES-UV light18081.82[225]
Oily wastewater250–10,000TiO2-P25PVDF-8W UV-A lamp24080 TOC degradation[226]
BSA1000GO/TiO2/PVP/ solution casting and phase inversionPVDF-UVA irradiation12092.5[227]
Bentazone10 N–Ti O 2 /PMAA/PVDF/PAN/Loeb-Souriraja-UV-light
and solar irradiation
18090.1[228]
Hexavalent chromium (0Cr (VI))10Chitosan-sodium alginate/Fe-doped WO3PES-300 W Xe lamp24099.9[186]

6. Membrane Fouling in PMRs

Different components of feed, such as organic substances, photocatalysts, colloids, salts, and cells, can have varying effects on system performance [229]. The fouling of PMR membranes is caused by the deposition of feed components on the membrane. Photocatalytic oxidation partially eliminates foulants [230]. Foulant adhesion, pore blocking, the cake layer formation, and temporal and spatial changes of foulant structure during long-term filtration are among factors that promote fouling [230,231]. Photocatalyst nano particles may form microaggregate and deposit on the membrane during filtration. Organic pollutants accumulation in the vicinity of membrane may result in formation of a very thin layer leading to substantial pore clogging and flow rate reduction. Adsorption of organic contaminants on TiO2 particles and its composites such as Degussa P25 TiO2, Ca alginate polymer/TiO2 fibers, nano-structured TiO2/silica gel photocatalyst, and titanium tetraisopropoxide in the absence of effective UV absorption may result in formation of a dense cake layer on the membrane surface and further flow rate reduction [232].
Designing non-fouling membranes is highly sought. Zheng et al. [233] combined the cellulose nanocrystals (CNCs) onto Cu-MOF-74 by physically stirring, and then coated the composites on the membrane to enhance the antifouling efficiency of PVDF membranes. CNC/Cu-MOF-74 composite coating on PVDF membrane increased its hydrophilicity, which in turn considerably improved the membrane’s permeability and productivity. They also reported enhanced electrostatic repulsion based on the contact angle test and Zeta-potential measurement. Due to Cu-predominate MOF-74’s antibacterial activity, the CNC/Cu-MOF-74 modified membrane also demonstrated increased antibacterial performance. effective antibacterial performance of composite membrane was attributed to Cu2+ release and •OH production.

6.1. Reactor Design

Modulating photocatalytic reactions, exchanging catalysts, and degrading pollutants are all easier with slurry PMRs. Although catalyst separation is not required for IPMRs, the catalyst loading cannot be tailored to the feed’s specific composition; higher catalyst loading, larger membrane surface area, and higher-pressure drop are required which increases the reactor volume, energy consumption of pumps, and process cost. Exchanging the catalyst is also a challenging process especially in IPMRs [234,235].

6.2. Photocatalyst Loading

Membrane foulants can be reduced and photocatalytic degradation may be accelerated by increasing catalyst loading due to increase reaction surface area. [234]; however, photocatalyst loading has an optimum value for any specific process after which increased opacity of the reaction mixture hinders light absorption by photocatalyst [13,236]. Elevated photocatalyst loading results in reduced foulant degradation which enhances fouling rate on the membrane [237,238].

7. Conclusions and Future Perspectives

This review described various hybrid photocatalysis and membrane process designs for removing organic contaminants from water. The main advantage of PMRs is retention and reuse of photocatalyst. The advantages and disadvantages IPMRs and SPMRs, as well as important design/operation parameters were discussed. The performances of a photocatalytic reaction can be improved by utilizing a suspended photocatalyst as opposed to an immobilized one, owing to larger active surface and subsequently improved photocatalyst–substrate interaction. SPMR empowered by air bubbles and effluent flushing appears to be more suitable for treating water and wastewater. Visible light may be utilized as an irradiation source and efficient solar-driven photocatalytic conversion has emerged in recent years. The appropriate selection of the membrane is crucial, as it must have great permeability to the desired product and retain contaminants and photocatalyst to facilitate the rapid removal of the product from the reaction environment. When it comes to treating wastewater, TiO2-based PMRs excel because of their excellent separation efficiency and low maintenance requirements. When designing visible-light-operated photocatalysts, it is important to examine the option of employing the sun as a clean, low-cost light source to make the process more environmentally friendly. Advantages of using PMRs for the partial oxidation and reduction of organic matter include (a) extending the lifetime of polymeric membranes with the help of visible light as a source of radiation, and (b) enhancing photocatalyst recovery through the use of novel materials in the synthesis of photocatalyst composites and semiconductor coatings on optical fibers. PMRs utilization has become a mode viable wastewater treatment option upon developments in photovoltaic technology (solar energy conversion) and the use of LED lamps (UV and/or visible).

Author Contributions

M.B. and S.M.M. developed the idea and structure of the review article. S.A.H. and J.R. wrote the manuscript and collected the materials from databases. S.M.M. and C.W.L. and S.S., revised and improved the manuscript. M.B. and S.S. and C.W.L. supervised the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research work was financially supported by Fundamental Research Grant Scheme FRGS/1/2020/TK0/UM/02/8 (No. FP023-2020), and Global Collaborative Programme—SATU Joint Research Scheme (No. ST004-2021).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

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

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AFAcid fuchsineMRsMembrane reactors
ANAacenaphtheneMWCOmolecular weight cut-offs
AOPsadvanced oxidation processesMSmultiple sclerosis
ARSAlizarin red SNAPnaphthalene
BPABisphenol-ANFnanofiltration
BR51Basic Red 51PANIPolyaniline
BSAbovine serum albuminPCPPentachlorophenol
BTEXbenzene, toluene, ethylbenzene, and xylenePdTFPP(pentafluorophenyl)-21H, 23H-porphine palladium (II)
CAClofibric acidPEpolyethylene
CBConduction bondPECMphotoelectrocatalytic membrane filtration
CBZcarbamazepinePenicillin GBenzylpenicillin
CeFCerium fluoridePESPolyethersulfone
CFXCefiximePHEphenanthrene
CIPCiprofloxacinpHpzcpoint of zero charge
CNCscellulose nanocrystalsPMRsphotocatalytic membrane reactors
COScarbon oxidation statePPPolypropylene
DCFdiclofenacPSPhosphatidylserine
DCLDiclofenacPVDFpolyvinylidene fluoride
DXMDexamethasoneRB5Reactive Black 5
E a band-gap energyROReverse osmosis
e CB conduction band electronROSreactive oxygen species
HPCheterogeneous photocatalysisSPMRSuspended photocatalytic membrane reactors
hvirradiated photonsTMPtransmembrane pressure
h VB + valence band holeTPHstotal petroleum hydrocarbons
IBUibuprofenUFultrafiltration
IPMRimmobilized photocatalytic membrane reactorsUVvisible light
MBmethylene blueVBValance band
MFmicrofiltrationZnONStzinc oxide nano stars
MIL-53(Fe)Matériaux de l′Institut LavoisierZTPGZnO tetrapod-reduced graphene oxide
MNZMetronidazole
MOMethyl Orange
MOFMetal–organic framework
AFAcid fuchsine
ANAacenaphthene
AOPsadvanced oxidation processes
ARSAlizarin red S
BPABisphenol-A
BR51Basic Red 51
BSAbovine serum albumin
BTEXbenzene, toluene, ethylbenzene, and xylene
CAClofibric acid
CBConduction bond
CBZcarbamazepine
CeFCerium fluoride
CFXCefixime

References

  1. He, C.; Cheng, J.; Zhang, X.; Douthwaite, M.; Pattisson, S.; Hao, Z. Recent Advances in the Catalytic Oxidation of Volatile Organic Compounds: A Review Based on Pollutant Sorts and Sources. Chem. Rev. 2019, 119, 4471–4568. [Google Scholar] [CrossRef] [PubMed]
  2. Gligorovski, S.; Strekowski, R.; Barbati, S.; Vione, D. Environmental Implications of Hydroxyl Radicals (•OH). Chem. Rev. 2015, 115, 13051–13092. [Google Scholar] [CrossRef]
  3. Ahmed, S.F.; Mehejabin, F.; Momtahin, A.; Tasannum, N.; Faria, N.T.; Mofijur, M.; Hoang, A.T.; Vo, D.-V.N.; Mahlia, T.M.I. Strategies to improve membrane performance in wastewater treatment. Chemosphere 2022, 306, 135527. [Google Scholar] [CrossRef] [PubMed]
  4. Fereidooni, M.; Esmaeilzadeh, F.; Zandifar, A. Innovatively-synthesized CeO2/ZnO photocatalysts by sono-photochemical deposition: Catalyst characterization and effect of operational parameters on high efficient dye removal. J. Mater. Sci. 2022, 57, 16228–16244. [Google Scholar] [CrossRef]
  5. Reis, A.C.; Kolvenbach, B.A.; Nunes, O.C.; Corvini, P.F.X. Biodegradation of antibiotics: The new resistance determinants—Part II. New Biotechnol. 2020, 54, 13–27. [Google Scholar] [CrossRef]
  6. Tian, Y.; Li, J.; Tang, L.; Meng, J.; Li, J. Antibiotics removal from piggery wastewater by a novel aerobic-microaerobic system: Efficiency and mechanism. Chem. Eng. J. 2023, 454, 140265. [Google Scholar] [CrossRef]
  7. Barancheshme, F.; Munir, M. Strategies to Combat Antibiotic Resistance in the Wastewater Treatment Plants. Front. Microbiol. 2018, 8, 2603. [Google Scholar] [CrossRef]
  8. Aguilar-Ascón, E.; Solari-Godiño, A.; Cueva-Martínez, M.; Neyra-Ascón, W.; Albrecht-Ruíz, M. Characterization of Sludge Resulting from Chemical Coagulation and Electrocoagulation of Pumping Water from Fishmeal Factories. Processes 2023, 11, 567. [Google Scholar]
  9. Binazadeh, M.; Karimi, I.A.; Li, Z. Fast biodegradation of long chain n-alkanes and crude oil at high concentrations with Rhodococcus sp. Moj-3449. Enzym. Microb. Technol. 2009, 45, 195–202. [Google Scholar] [CrossRef]
  10. Andreyev, S.Y.; Lebedinskiy, K.V.; Stepanov, S. A novel technology for optimizing dissolved air flotation unit efficiency via secondary saturation of the flotation cell with air bubbles and thin-layer settling. Chem. Eng. Process.-Process Intensif. 2023, 184, 109292. [Google Scholar]
  11. Ahmad, A.; Priyadarshini, M.; Das, I.; Ghangrekar, M.M.; Surampalli, R.Y. Surfactant aided electrocoagulation/flotation using punched electrodes for the remediation of salicylic acid from wastewater. J. Environ. Chem. Eng. 2023, 11, 109049. [Google Scholar] [CrossRef]
  12. Wysocka, I. Absorption processes in reducing the odor nuisance of wastewater. MethodsX 2023, 10, 101996. [Google Scholar] [CrossRef] [PubMed]
  13. Rasouli, K.; Alamdari, A.; Sabbaghi, S. Ultrasonic-assisted synthesis of α-Fe2O3@ TiO2 photocatalyst: Optimization of effective factors in the fabrication of photocatalyst and removal of non-biodegradable cefixime via response surface methodology-central composite design. Sep. Purif. Technol. 2023, 307, 122799. [Google Scholar] [CrossRef]
  14. Kusworo, T.D.; Kumoro, A.C.; Aryanti, N.; Hasbullah, H.; Chaesarifa, D.R.S.; Fauzan, M.D.; Dalanta, F. Developing a robust photocatalytic and antifouling performance of PVDF membrane using spinel NiFe2O4/GO photocatalyst for efficient industrial dye wastewater treatment. J. Environ. Chem. Eng. 2023, 11, 109449. [Google Scholar] [CrossRef]
  15. Binazadeh, M.; Li, Z.; Karimi, I.A. Optimization of biodegradation of long chain n-Alkanes by Rhodococcus sp. Moj-3449 using response surface methodology. Phys. Chem. Res. 2020, 8, 45–59. [Google Scholar]
  16. Moradi, H.; Sabbaghi, S.; Mirbagheri, N.S.; Chen, P.; Rasouli, K.; Kamyab, H.; Chelliapan, S. Removal of chloride ion from drinking water using Ag NPs-Modified bentonite: Characterization and optimization of effective parameters by response surface methodology-central composite design. Environ. Res. 2023, 223, 115484. [Google Scholar] [CrossRef]
  17. Molinari, R.; Lavorato, C.; Argurio, P.; Szymański, K.; Darowna, D.; Mozia, S. Overview of photocatalytic membrane reactors in organic synthesis, energy storage and environmental applications. Catalysts 2019, 9, 239. [Google Scholar] [CrossRef]
  18. Pervez, M.N.; Talukder, M.E.; Mishu, M.R.; Buonerba, A.; Del Gaudio, P.; Stylios, G.K.; Hasan, S.W.; Zhao, Y.; Cai, Y.; Figoli, A.; et al. One-Step Fabrication of Novel Polyethersulfone-Based Composite Electrospun Nanofiber Membranes for Food Industry Wastewater Treatment. Membranes 2022, 12, 413. [Google Scholar] [CrossRef]
  19. Koe, W.S.; Lee, J.W.; Chong, W.C.; Pang, Y.L.; Sim, L.C. An overview of photocatalytic degradation: Photocatalysts, mechanisms, and development of photocatalytic membrane. Environ. Sci. Pollut. Res. 2020, 27, 2522–2565. [Google Scholar] [CrossRef]
  20. Khraibet, A.C.; Imran, N.J.; Majeed, H.M.; Ehmood, M.A.; Hasoon, G.S.; Nathim, Z.F.; Alwan, A.K.; Abd Alsada, A.S. Using titanium oxide membranes and ultraviolet (UV) light to remove pharmaceutical waste from hospitals wastewater. J. Genet. Environ. Resour. Conserv. 2022, 10, 41–48. [Google Scholar]
  21. Sim, S.I.; Teow, Y.H. Integrated Membrane-adsorption system as a sustainable development approach for semiconductor-industry wastewater treatment. Mater. Today Proc. 2023. [Google Scholar] [CrossRef]
  22. Vasishta, A.; Mahale, J.S.; Pandey, P.H.; Ukarde, T.M.; Shinde, P.; Pawar, H.S. Membrane Separation: An Advanced Tool for the Development of a Wastewater Treatment Process. In Membrane and Membrane-Based Processes for Wastewater Treatment; CRC Press: Boca Raton, FL, USA, 2023; pp. 17–34. [Google Scholar]
  23. Wu, J.; Wu, Y.; Hu, X.; Wu, C.; Ding, J. Water-bonding tubular membrane used in a 3D-printing dialyzer for diffusion dialysis. J. Membr. Sci. 2022, 664, 121078. [Google Scholar] [CrossRef]
  24. Xie, J.; Yang, Y.; Zhang, H.; Chen, S.; Lv, Z.; Zhou, Y.; Qi, J.; Sun, X.; Li, J. ZIF-67 derived Co/N carbon hollow fiber membrane with excellent decontamination performance. Chem. Eng. J. 2023, 451, 138403. [Google Scholar] [CrossRef]
  25. Dias, R.A.; Ferreira, R.S.B.; Medeiros, V.d.N.; Araujo, B.A.; Araújo, E.M.; Lira, H.d.L. Flat membranes of polyethersulfone/polysulfone blends in water/oil separation. Polym. Bull. 2022, 80, 4289–4305. [Google Scholar] [CrossRef]
  26. Khalili, M.; Sabbaghi, S.; Zerafat, M.M. Preparation of ceramic γ-Al2O3–TiO2 nanofiltration membranes for desalination. Chem. Pap. 2015, 69, 309–315. [Google Scholar] [CrossRef]
  27. Baker, R.W. Membrane Technology and Applications; John Wiley & Sons: Hoboken, NJ, USA, 2012. [Google Scholar]
  28. Pérez-Silva, I.; Páez-Hernández, M.; Ibarra, I.S.; Camacho-Mendoza, R.L. Evaluation of the Hybrid Membrane of ZnO Particles Supported in Cellulose Acetate for the Removal of Lead. Membranes 2023, 13, 123. [Google Scholar] [CrossRef] [PubMed]
  29. Kayvani Fard, A.; McKay, G.; Buekenhoudt, A.; Al Sulaiti, H.; Motmans, F.; Khraisheh, M.; Atieh, M. Inorganic membranes: Preparation and application for water treatment and desalination. Materials 2018, 11, 74. [Google Scholar] [CrossRef]
  30. Milovanovic, M.; Tabakoglu, F.; Saki, F.; Pohlkoetter, E.; Buga, D.; Brandt, V.; Tiller, J.C. Organic-inorganic double networks as highly permeable separation membranes with a chiral selector for organic solvents. J. Membr. Sci. 2023, 668, 121190. [Google Scholar] [CrossRef]
  31. Darowna, D.; Wróbel, R.; Morawski, A.W.; Mozia, S. The influence of feed composition on fouling and stability of a polyethersulfone ultrafiltration membrane in a photocatalytic membrane reactor. Chem. Eng. J. 2017, 310, 360–367. [Google Scholar] [CrossRef]
  32. Malekshahi, M.; Sabbaghi, S.; Rasouli, K. Preparation of α-alumina/γ-alumina/γ-alumina-titania ceramic composite membrane for chloride ion removal. Mater. Chem. Phys. 2022, 287, 126218. [Google Scholar] [CrossRef]
  33. Pabby, A.K.; Rizvi, S.S.; Requena, A.M.S. Handbook of Membrane Separations: Chemical, Pharmaceutical, Food, and Biotechnological Applications; CRC Press: Boca Raton, FL, USA, 2015. [Google Scholar]
  34. Mozia, S.; Rajakumaran, R.; Szymański, K.; Gryta, M. Removal of ketoprofen from surface water in a submerged photocatalytic membrane reactor utilizing membrane distillation: Effect of process parameters and evaluation of long-term performance. J. Chem. Technol. Biotechnol. 2023. [Google Scholar] [CrossRef]
  35. Feng, X.; Long, R.; Liu, C.; Liu, X. Novel dual-heterojunction photocatalytic membrane reactor based on Ag2S/NH2-MIL-88B (Fe)/poly (aryl ether nitrile) composite with enhanced photocatalytic performance for wastewater purification. Chem. Eng. J. 2023, 454, 139765. [Google Scholar] [CrossRef]
  36. Molinari, R.; Limonti, C.; Lavorato, C.; Siciliano, A.; Argurio, P. Upgrade of a slurry photocatalytic membrane reactor based on a vertical filter and an external membrane and testing in the photodegradation of a model pollutant in water. Chem. Eng. J. 2023, 451, 138577. [Google Scholar] [CrossRef]
  37. George, J.; Kumar, V.V. Designing a novel poly (methyl vinyl ether maleic anhydride) based polymeric membrane with enhanced antifouling performance for removal of pentachlorophenol from aqueous solution. Environ. Res. 2023, 223, 115404. [Google Scholar] [CrossRef] [PubMed]
  38. Tian, S.; He, Y.; Zhang, L.; Li, S.; Bai, Y.; Wang, Y.; Wu, J.; Yu, J.; Guo, X. CNTs/TiO2-loaded carbonized nanofibrous membrane with two-type self-cleaning performance for high efficiency oily wastewater remediation. Colloids Surf. A Physicochem. Eng. Asp. 2023, 656, 130306. [Google Scholar] [CrossRef]
  39. Sisi, A.J.; Fathinia, M.; Khataee, A.; Orooji, Y. Systematic activation of potassium peroxydisulfate with ZIF-8 via sono-assisted catalytic process: Mechanism and ecotoxicological analysis. J. Mol. Liq. 2020, 308, 113018. [Google Scholar] [CrossRef]
  40. Ghasemi, M.; Khataee, A.; Gholami, P.; Soltani, R.D.C.; Hassani, A.; Orooji, Y. In-situ electro-generation and activation of hydrogen peroxide using a CuFeNLDH-CNTs modified graphite cathode for degradation of cefazolin. J. Environ. Manag. 2020, 267, 110629. [Google Scholar] [CrossRef]
  41. Sharma, K.; Vaya, D.; Prasad, G.; Surolia, P. Photocatalytic process for oily wastewater treatment: A review. Int. J. Environ. Sci. Technol. 2022, 20, 4615–4634. [Google Scholar] [CrossRef]
  42. Chakachaka, V.; Tshangana, C.; Mahlangu, O.; Mamba, B.; Muleja, A. Interdependence of Kinetics and Fluid Dynamics in the Design of Photocatalytic Membrane Reactors. Membranes 2022, 12, 745. [Google Scholar] [CrossRef]
  43. Hou, C.; Yuan, X.; Niu, M.; Li, Y.; Wang, L.; Zhang, M. In situ composite of Co-MOF on a Ti-based material for visible light multiphase catalysis: Synthesis and the photocatalytic degradation mechanism. New J. Chem. 2022, 46, 11341–11349. [Google Scholar] [CrossRef]
  44. Wu, L.; Fu, C.; Huang, W. Surface chemistry of TiO2 connecting thermal catalysis and photocatalysis. Phys. Chem. Chem. Phys. 2020, 22, 9875–9909. [Google Scholar] [CrossRef]
  45. Zheng, X.; Shen, Z.-P.; Shi, L.; Cheng, R.; Yuan, D.-H. Photocatalytic membrane reactors (PMRs) in water treatment: Configurations and influencing factors. Catalysts 2017, 7, 224. [Google Scholar] [CrossRef]
  46. Ensie, B.; Samad, S. Removal of nitrate from drinking water using nano SiO2–FeOOH–Fe core–shell. Desalination 2014, 347, 1–9. [Google Scholar] [CrossRef]
  47. Molinari, R.; Argurio, P.; Bellardita, M.; Palmisano, L.; Bertoni, C. Photocatalytic processes in membrane reactors. In Comprehensive Membrane Science and Engineering, 2nd ed.; Elsevier: Oxford, UK, 2017. [Google Scholar]
  48. Vijayakumar, E.; Govinda Raj, M.; Narendran, M.G.; Preetha, R.; Mohankumar, R.; Neppolian, B.; John Bosco, A. Promoting Spatial Charge Transfer of ZrO2 Nanoparticles: Embedded on Layered MoS2/g-C3N4 Nanocomposites for Visible-Light-Induced Photocatalytic Removal of Tetracycline. ACS Omega 2022, 7, 5079–5095. [Google Scholar] [CrossRef]
  49. Hardikar, M.; Marquez, I.; Achilli, A. Emerging investigator series: Membrane distillation and high salinity: Analysis and implications. Environ. Sci. Water Res. Technol. 2020, 6, 1538–1552. [Google Scholar] [CrossRef]
  50. Aliyu, U.M.; Rathilal, S.; Isa, Y.M. Membrane desalination technologies in water treatment: A review. Water Pract. Technol. 2018, 13, 738–752. [Google Scholar] [CrossRef]
  51. Rao, L.; Tang, J.; Hu, S.; Shen, L.; Xu, Y.; Li, R.; Lin, H. Inkjet printing assisted electroless Ni plating to fabricate nickel coated polypropylene membrane with improved performance. J. Colloid Interface Sci. 2020, 565, 546–554. [Google Scholar] [CrossRef]
  52. Hassanzadeh, E.; Farhadian, M.; Razmjou, A.; Askari, N. An efficient wastewater treatment approach for a real woolen textile industry using a chemical assisted NF membrane process. Environ. Nanotechnol. Monit. Manag. 2017, 8, 92–96. [Google Scholar] [CrossRef]
  53. Wang, L.; Liang, W.; Chen, W.; Zhang, W.; Mo, J.; Liang, K.; Tang, B.; Zheng, Y.; Jiang, F. Integrated aerobic granular sludge and membrane process for enabling municipal wastewater treatment and reuse water production. Chem. Eng. J. 2018, 337, 300–311. [Google Scholar] [CrossRef]
  54. Alardhi, S.M.; Albayati, T.M.; Alrubaye, J.M. A hybrid adsorption membrane process for removal of dye from synthetic and actual wastewater. Chem. Eng. Process.-Process Intensif. 2020, 157, 108113. [Google Scholar] [CrossRef]
  55. Yun, T.; Chung, J.W.; Kwak, S.-Y. Recovery of sulfuric acid aqueous solution from copper-refining sulfuric acid wastewater using nanofiltration membrane process. J. Environ. Manag. 2018, 223, 652–657. [Google Scholar] [CrossRef] [PubMed]
  56. Salahi, A.; Noshadi, I.; Badrnezhad, R.; Kanjilal, B.; Mohammadi, T. Nano-porous membrane process for oily wastewater treatment: Optimization using response surface methodology. J. Environ. Chem. Eng. 2013, 1, 218–225. [Google Scholar] [CrossRef]
  57. Ren, Q.; Chen, X.; Yumminaga, Y.; Wang, N.; Yan, W.; Li, Y.; Liu, L.; Shi, J. Effect of operating conditions on the performance of multichannel ceramic ultrafiltration membranes for cattle wastewater treatment. J. Water Process Eng. 2021, 41, 102102. [Google Scholar] [CrossRef]
  58. Belibagli, P.; Isik, Z.; Özdemir, S.; Gonca, S.; Dizge, N.; Awasthi, M.K.; Balakrishnan, D. An integrated process for wet scrubber wastewater treatment using electrooxidation and pressure-driven membrane filtration. Chemosphere 2022, 308, 136216. [Google Scholar] [CrossRef] [PubMed]
  59. Sathya, U.; Nithya, M.; Balasubramanian, N. Evaluation of advanced oxidation processes (AOPs) integrated membrane bioreactor (MBR) for the real textile wastewater treatment. J. Environ. Manag. 2019, 246, 768–775. [Google Scholar] [CrossRef] [PubMed]
  60. Tang, Y.; Lin, Y.; Ford, D.M.; Qian, X.; Cervellere, M.R.; Millett, P.C.; Wang, X. A review on models and simulations of membrane formation via phase inversion processes. J. Membr. Sci. 2021, 640, 119810. [Google Scholar] [CrossRef]
  61. Wang, Y.; Li, T.; Zhu, J. Study on treatment of wastewater with low concentration of ammonia-nitrogen by vacuum plate membrane distillation technology. Water Sci. Technol. 2022, 86, 950–967. [Google Scholar] [CrossRef]
  62. Al-Juboori, R.A.; Naji, O.; Bowtell, L.; Alpatova, A.; Soukane, S.; Ghaffour, N. Power effect of ultrasonically vibrated spacers in air gap membrane distillation: Theoretical and experimental investigations. Sep. Purif. Technol. 2021, 262, 118319. [Google Scholar] [CrossRef]
  63. Kubo, M.; Kojima, M.; Mano, R.; Daiko, Y.; Honda, S.; Iwamoto, Y. A hydrostable mesoporous γ-Al2O3 membrane modified with Si–C–H organic-inorganic hybrid derived from polycarbosilane. J. Membr. Sci. 2020, 598, 117799. [Google Scholar] [CrossRef]
  64. Abd Jalil, S.N. Investigation of Vacuum-Assisted Preparation Methods of Inorganic Membranes. Master’s Thesis, School of Chemical Engineering, The University of Queensland, St Lucia, QLD, Australia, 2017. [Google Scholar]
  65. Mamivand, S.; Binazadeh, M.; Sohrabi, R. Applicability of membrane reactor technology in industrial hydrogen producing reactions: Current effort and future directions. J. Ind. Eng. Chem. 2021, 104, 212–230. [Google Scholar] [CrossRef]
  66. Khan, A.; Khan, S.J.; Miran, W.; Zaman, W.Q.; Aslam, A.; Shahzad, H.M.A. Feasibility Study of Anaerobic Baffled Reactor Coupled with Anaerobic Filter Followed by Membrane Filtration for Wastewater Treatment. Membranes 2023, 13, 79. [Google Scholar] [CrossRef] [PubMed]
  67. Jang, D.; Lee, J.; Jang, A. Impact of pre-coagulation on the ceramic membrane process during oil-water emulsion separation: Fouling behavior and mechanism. Chemosphere 2023, 313, 137596. [Google Scholar] [CrossRef] [PubMed]
  68. Ghalamchi, L.; Aber, S.; Vatanpour, V.; Kian, M. Comparison of NLDH and g-C3N4 nanoplates and formative Ag3PO4 nanoparticles in PES microfiltration membrane fouling: Applications in MBR. Chem. Eng. Res. Des. 2019, 147, 443–457. [Google Scholar] [CrossRef]
  69. Shahabi, S.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]
  70. Turgut, F.; Chong, C.Y.; Karaman, M.; Lau, W.J.; Gürsoy, M.; Ismail, A.F. Plasma surface modification of graphene oxide nanosheets for the synthesis of GO/PES nanocomposite ultrafiltration membrane for enhanced oily separation. J. Appl. Polym. Sci. 2023, 140, e53410. [Google Scholar] [CrossRef]
  71. Goh, P.S.; Samavati, Z.; Ismail, A.F.; Ng, B.C.; Abdullah, M.S.; Hilal, N. Modification of Liquid Separation Membranes Using Multidimensional Nanomaterials: Revealing the Roles of Dimension Based on Classical Titanium Dioxide. Nanomaterials 2023, 13, 448. [Google Scholar] [CrossRef]
  72. Bera, S.P.; Godhaniya, M.; Kothari, C. Emerging and advanced membrane technology for wastewater treatment: A review. J. Basic Microbiol. 2022, 62, 245–259. [Google Scholar] [CrossRef]
  73. Gupta, V.; Anandkumar, J. Membrane Processes. In Membrane and Membrane-Based Processes for Wastewater Treatment; CRC Press: Boca Raton, FL, USA, 2023; pp. 199–211. [Google Scholar]
  74. Hyeon, Y.; Kim, S.; Ok, E.; Park, C. A fluid imaging flow cytometry for rapid characterization and realistic evaluation of microplastic fiber transport in ceramic membranes for laundry wastewater treatment. Chem. Eng. J. 2023, 454, 140028. [Google Scholar] [CrossRef]
  75. Huang, S.; Wu, G.; Chen, S. Preparation of microporous poly (vinylidene fluoride) membranes via phase inversion in supercritical CO2. J. Membr. Sci. 2007, 293, 100–110. [Google Scholar] [CrossRef]
  76. Kim, N.; Kim, C.-S.; Lee, Y.-T. Preparation and characterization of polyethersulfone membranes with p-toluenesulfonic acid and polyvinylpyrrolidone additives. Desalination 2008, 233, 218–226. [Google Scholar] [CrossRef]
  77. Lee, A.; Elam, J.W.; Darling, S.B. Membrane materials for water purification: Design, development, and application. Environ. Sci. Water Res. Technol. 2016, 2, 17–42. [Google Scholar] [CrossRef]
  78. Zikalala, S.A.; Chabalala, M.B.; Gumbi, N.N.; Coville, N.J.; Mamba, B.B.; Mutuma, B.K.; Nxumalo, E.N. Microwave-assisted synthesis of titania–amorphous carbon nanotubes/amorphous nitrogen-doped carbon nanotubes nanohybrids for photocatalytic degradation of textile wastewater. RSC Adv. 2021, 11, 6748–6763. [Google Scholar] [CrossRef] [PubMed]
  79. Szymański, K.; Gryta, M.; Darowna, D.; Mozia, S. A new submerged photocatalytic membrane reactor based on membrane distillation for ketoprofen removal from various aqueous matrices. Chem. Eng. J. 2022, 435, 134872. [Google Scholar] [CrossRef]
  80. Gupta, S.; Gomaa, H.; Ray, M.B. Fouling control in a submerged membrane reactor: Aeration vs. membrane oscillations. Chem. Eng. J. 2022, 432, 134399. [Google Scholar] [CrossRef]
  81. Ganiyu, S.O.; Van Hullebusch, E.D.; Cretin, M.; Esposito, G.; Oturan, M.A. Coupling of membrane filtration and advanced oxidation processes for removal of pharmaceutical residues: A critical review. Sep. Purif. Technol. 2015, 156, 891–914. [Google Scholar] [CrossRef]
  82. Hu, C.; Wang, M.-S.; Chen, C.-H.; Chen, Y.-R.; Huang, P.-H.; Tung, K.-L. Phosphorus-doped g-C3N4 integrated photocatalytic membrane reactor for wastewater treatment. J. Membr. Sci. 2019, 580, 1–11. [Google Scholar] [CrossRef]
  83. Hu, C.; Yoshida, M.; Huang, P.-H.; Tsunekawa, S.; Hou, L.-B.; Chen, C.-H.; Tung, K.-L. MIL-88B(Fe)-coated photocatalytic membrane reactor with highly stable flux and phenol removal efficiency. Chem. Eng. J. 2021, 418, 129469. [Google Scholar] [CrossRef]
  84. Nguyen, V.-H.; Tran, Q.B.; Nguyen, X.C.; Hai, L.T.; Ho, T.T.T.; Shokouhimehr, M.; Vo, D.-V.N.; Lam, S.S.; Nguyen, H.P.; Hoang, C.T.; et al. Submerged photocatalytic membrane reactor with suspended and immobilized N-doped TiO2 under visible irradiation for diclofenac removal from wastewater. Process Saf. Environ. Prot. 2020, 142, 229–237. [Google Scholar] [CrossRef]
  85. Wu, C.-J.; Valerie Maggay, I.; Chiang, C.-H.; Chen, W.; Chang, Y.; Hu, C.; Venault, A. Removal of tetracycline by a photocatalytic membrane reactor with MIL-53(Fe)/PVDF mixed-matrix membrane. Chem. Eng. J. 2023, 451, 138990. [Google Scholar] [CrossRef]
  86. Espíndola, J.C.; Cristóvão, R.O.; Mendes, A.; Boaventura, R.A.R.; Vilar, V.J.P. Photocatalytic membrane reactor performance towards oxytetracycline removal from synthetic and real matrices: Suspended vs. immobilized TiO2-P25. Chem. Eng. J. 2019, 378, 122114. [Google Scholar] [CrossRef]
  87. Salehian, S.; Heydari, H.; Khansanami, M.; Vatanpour, V.; Mousavi, S.A. Fabrication and performance of polysulfone/H2O2-g-C3N4 mixed matrix membrane in a photocatalytic membrane reactor under visible light irradiation for removal of natural organic matter. Sep. Purif. Technol. 2022, 285, 120291. [Google Scholar] [CrossRef]
  88. Ahmadi, A.; Sarrafzadeh, M.-H.; Hosseinian, A.; Ghaffari, S.-B. Foulant layer degradation of dye in Photocatalytic Membrane Reactor (PMR) containing immobilized and suspended NH2-MIL125(Ti) MOF led to water flux recovery. J. Environ. Chem. Eng. 2022, 10, 106999. [Google Scholar] [CrossRef]
  89. Rathna, T.; PonnanEttiyappan, J.; RubenSudhakar, D. Fabrication of visible-light assisted TiO2-WO3-PANI membrane for effective reduction of chromium (VI) in photocatalytic membrane reactor. Environ. Technol. Innov. 2021, 24, 102023. [Google Scholar] [CrossRef]
  90. Petsi, P.N.; Sarasidis, V.C.; Plakas, K.V.; Karabelas, A.J. Reduction of nitrates in a photocatalytic membrane reactor in the presence of organic acids. J. Environ. Manag. 2021, 298, 113526. [Google Scholar] [CrossRef] [PubMed]
  91. Azimifar, M.; Ghorbani, M.; Peyravi, M. Fabrication and evaluation of a photocatalytic membrane based on Sb2O3/CBO composite for improvement of dye removal efficiency. J. Mol. Struct. 2022, 1270, 133957. [Google Scholar] [CrossRef]
  92. Hindryawati, N.; Maniam, G.P.; Pratama, I.R.; Gunawan, R.; Koesnarpadi, S. Study of Sonocatalytic Activity ZnO-WO3 Composite on Degradation Phenol in Aqueous Solution. J. Bahan Alam Terbarukan 2022, 11, 50–57. [Google Scholar] [CrossRef]
  93. Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef]
  94. Zhang, J.; Tian, B.; Wang, L.; Xing, M.; Lei, J. Mechanism of photocatalysis. In Photocatalysis; Springer: Berlin/Heidelberg, Germany, 2018; pp. 1–15. [Google Scholar]
  95. Khalid, N.; Majid, A.; Tahir, M.B.; Niaz, N.; Khalid, S. Carbonaceous-TiO2 nanomaterials for photocatalytic degradation of pollutants: A review. Ceram. Int. 2017, 43, 14552–14571. [Google Scholar] [CrossRef]
  96. Nair, V.; Muñoz-Batista, M.J.; Fernández-García, M.; Luque, R.; Colmenares, J.C. Thermo-photocatalysis: Environmental and energy applications. ChemSusChem 2019, 12, 2098–2116. [Google Scholar] [CrossRef]
  97. Zyoud, A.H.; Zubi, A.; Hejjawi, S.; Zyoud, S.H.; Helal, M.H.; Zyoud, S.H.; Qamhieh, N.; Hajamohideen, A.; Hilal, H.S. Removal of acetaminophen from water by simulated solar light photodegradation with ZnO and TiO2 nanoparticles: Catalytic efficiency assessment for future prospects. J. Environ. Chem. Eng. 2020, 8, 104038. [Google Scholar] [CrossRef]
  98. Sabouni, R.; Gomaa, H. Photocatalytic degradation of pharmaceutical micro-pollutants using ZnO. Environ. Sci. Pollut. Res. 2019, 26, 5372–5380. [Google Scholar] [CrossRef] [PubMed]
  99. Abdel-Wahab, A.-M.; Al-Shirbini, A.-S.; Mohamed, O.; Nasr, O. Photocatalytic degradation of paracetamol over magnetic flower-like TiO2/Fe2O3 core-shell nanostructures. J. Photochem. Photobiol. A Chem. 2017, 347, 186–198. [Google Scholar] [CrossRef]
  100. Guo, Q.; Tang, G.; Zhu, W.; Luo, Y.; Gao, X. In situ construction of Z-scheme FeS2/Fe2O3 photocatalyst via structural transformation of pyrite for photocatalytic degradation of carbamazepine and the synergistic reduction of Cr (VI). J. Environ. Sci. 2021, 101, 351–360. [Google Scholar] [CrossRef] [PubMed]
  101. Berkani, M.; Smaali, A.; Kadmi, Y.; Almomani, F.; Vasseghian, Y.; Lakhdari, N.; Alyane, M. Photocatalytic degradation of Penicillin G in aqueous solutions: Kinetic, degradation pathway, and microbioassays assessment. J. Hazard. Mater. 2022, 421, 126719. [Google Scholar] [CrossRef]
  102. Arghavan, F.S.; Hossein Panahi, A.; Nasseh, N.; Ghadirian, M. Adsorption-photocatalytic processes for removal of pentachlorophenol contaminant using FeNi3/SiO2/ZnO magnetic nanocomposite under simulated solar light irradiation. Environ. Sci. Pollut. Res. 2021, 28, 7462–7475. [Google Scholar] [CrossRef]
  103. Eskandari, M.; Goudarzi, N.; Moussavi, S.G. Application of low-voltage UVC light and synthetic ZnO nanoparticles to photocatalytic degradation of ciprofloxacin in aqueous sample solutions. Water Environ. J. 2018, 32, 58–66. [Google Scholar] [CrossRef]
  104. Bohdziewicz, J.; Kudlek, E.; Dudziak, M. Influence of the catalyst type (TiO2 and ZnO) on the photocatalytic oxidation of pharmaceuticals in the aquatic environment. Desalination Water Treat. 2016, 57, 1552–1563. [Google Scholar] [CrossRef]
  105. Beheshti, F.; Tehrani, R.M.A.; Khadir, A. Sulfamethoxazole removal by photocatalytic degradation utilizing TiO2 and WO3 nanoparticles as catalysts: Analysis of various operational parameters. Int. J. Environ. Sci. Technol. 2019, 16, 7987–7996. [Google Scholar] [CrossRef]
  106. Ghenaatgar, A.; Tehrani, R.M.; Khadir, A. Photocatalytic degradation and mineralization of dexamethasone using WO3 and ZrO2 nanoparticles: Optimization of operational parameters and kinetic studies. J. Water Process Eng. 2019, 32, 100969. [Google Scholar] [CrossRef]
  107. Farzadkia, M.; Bazrafshan, E.; Esrafili, A.; Yang, J.-K.; Shirzad-Siboni, M. Photocatalytic degradation of Metronidazole with illuminated TiO2 nanoparticles. J. Environ. Health Sci. Eng. 2015, 13, 35. [Google Scholar] [CrossRef]
  108. Heidari, S.; Haghighi, M.; Shabani, M. Sunlight-activated BiOCl/BiOBr–Bi24O31Br10 photocatalyst for the removal of pharmaceutical compounds. J. Clean. Prod. 2020, 259, 120679. [Google Scholar] [CrossRef]
  109. Chinnaiyan, P.; Thampi, S.; Kumar, M.; Balachandran, M. Photocatalytic degradation of metformin and amoxicillin in synthetic hospital wastewater: Effect of classical parameters. Int. J. Environ. Sci. Technol. 2019, 16, 5463–5474. [Google Scholar] [CrossRef]
  110. Thi, V.H.-T.; Lee, B.-K. Effective photocatalytic degradation of paracetamol using La-doped ZnO photocatalyst under visible light irradiation. Mater. Res. Bull. 2017, 96, 171–182. [Google Scholar] [CrossRef]
  111. Salehia, F.; Sabbaghia, S.; Mirbagherib, N.S. Modification of graphitic carbon nitride photocatalyst by Pb-contaminated water for efficient removal of cefixime from aqueous media. Desalination Water Treat. 2021, 229, 331–342. [Google Scholar] [CrossRef]
  112. Cruz, D.; Ortiz-Oliveros, H.B.; Flores-Espinosa, R.M.; Ávila Pérez, P.; Ruiz-López, I.I.; Quiroz-Estrada, K.F. Synthesis of Ag/TiO2 composites by combustion modified and subsequent use in the photocatalytic degradation of dyes. J. King Saud Univ.-Sci. 2022, 34, 101966. [Google Scholar] [CrossRef]
  113. Sima, J.; Hasal, P. Photocatalytic degradation of textile dyes in a TiO2/UV system. Chem. Eng. Trans. 2013, 32, 79–84. [Google Scholar]
  114. Kumaresan, A.; Arun, A.; Kalpana, V.; Vinupritha, P.; Sundaravadivel, E. Polymer-supported NiWO4 nanocomposites for visible light degradation of toxic dyes. J. Mater. Sci. Mater. Electron. 2022, 33, 9660–9668. [Google Scholar] [CrossRef]
  115. Ye, Z.; Kong, L.; Chen, F.; Chen, Z.; Lin, Y.; Liu, C. A comparative study of photocatalytic activity of ZnS photocatalyst for degradation of various dyes. Optik 2018, 164, 345–354. [Google Scholar] [CrossRef]
  116. Shi, J.; Zheng, J.; Wu, P.; Ji, X. Immobilization of TiO2 films on activated carbon fiber and their photocatalytic degradation properties for dye compounds with different molecular size. Catal. Commun. 2008, 9, 1846–1850. [Google Scholar] [CrossRef]
  117. Kumar, S.; Kaushik, R.; Purohit, L. Novel ZnO tetrapod-reduced graphene oxide nanocomposites for enhanced photocatalytic degradation of phenolic compounds and MB dye. J. Mol. Liq. 2021, 327, 114814. [Google Scholar] [CrossRef]
  118. Yashni, G.; AlGheethi, A.; Mohamed, R.M.S.R.; Arifin, S.N.H.; Shanmugan, V.A.; Kassim, A.H.M. Photocatalytic degradation of basic red 51 dye in artificial bathroom greywater using zinc oxide nanoparticles. Mater. Today Proc. 2020, 31, 136–139. [Google Scholar] [CrossRef]
  119. Ameen, F.; Dawoud, T.; AlNadhari, S. Ecofriendly and low-cost synthesis of ZnO nanoparticles from Acremonium potronii for the photocatalytic degradation of azo dyes. Environ. Res. 2021, 202, 111700. [Google Scholar] [CrossRef] [PubMed]
  120. Arikal, D.; Kallingal, A. Photocatalytic degradation of azo and anthraquinone dye using TiO2/MgO nanocomposite immobilized chitosan hydrogels. Environ. Technol. 2021, 42, 2278–2291. [Google Scholar] [CrossRef]
  121. Zeng, Q.; Liu, Y.; Shen, L.; Lin, H.; Yu, W.; Xu, Y.; Li, R.; Huang, L. Facile preparation of recyclable magnetic Ni@ filter paper composite materials for efficient photocatalytic degradation of methyl orange. J. Colloid Interface Sci. 2021, 582, 291–300. [Google Scholar] [CrossRef] [PubMed]
  122. de Jesus Cubas, P.; Semkiw, A.W.; Monteiro, F.C.; Los Weinert, P.; Monteiro, J.F.H.L.; Fujiwara, S.T. Synthesis of CuCr2O4 by self-combustion method and photocatalytic activity in the degradation of Azo Dye with visible light. J. Photochem. Photobiol. A Chem. 2020, 401, 112797. [Google Scholar] [CrossRef]
  123. Aziz, A.; Ali, N.; Khan, A.; Bilal, M.; Malik, S.; Ali, N.; Khan, H. Chitosan-zinc sulfide nanoparticles, characterization and their photocatalytic degradation efficiency for azo dyes. Int. J. Biol. Macromol. 2020, 153, 502–512. [Google Scholar] [CrossRef]
  124. El Nahrawy, A.M.; Abou Hammad, A.B.; Bakr, A.M.; Hemdan, B.A.; Wassel, A.R. Decontamination of ubiquitous harmful microbial lineages in water using an innovative Zn2Ti0.8Fe0.2O4 nanostructure: Dielectric and terahertz properties. Heliyon 2019, 5, e02501. [Google Scholar] [CrossRef]
  125. Guo, N.; Zeng, Y.; Li, H.; Xu, X.; Yu, H.; Han, X. Novel mesoporous TiO2@ g-C3N4 hollow core@ shell heterojunction with enhanced photocatalytic activity for water treatment and H2 production under simulated sunlight. J. Hazard. Mater. 2018, 353, 80–88. [Google Scholar] [CrossRef]
  126. Mohsenzadeh, M.; Mirbagheri, S.A.; Sabbaghi, S. Degradation of 1,2-dichloroethane by photocatalysis using immobilized PAni-TiO2 nano-photocatalyst. Environ. Sci. Pollut. Res. 2019, 26, 31328–31343. [Google Scholar] [CrossRef]
  127. Schnabel, T.; Jautzus, N.; Mehling, S.; Springer, C.; Londong, J. Photocatalytic degradation of hydrocarbons and methylene blue using floatable titanium dioxide catalysts in contaminated water. Water Reuse 2021, 11, 224–235. [Google Scholar] [CrossRef]
  128. Mousavi, S.M.; Hashemi, S.A.; Zarei, M.; Bahrani, S.; Savardashtaki, A.; Esmaeili, H.; Lai, C.W.; Mazraedoost, S.; Abassi, M.; Ramavandi, B. Data on cytotoxic and antibacterial activity of synthesized Fe3O4 nanoparticles using Malva sylvestris. Data Brief 2020, 28, 104929. [Google Scholar] [CrossRef] [PubMed]
  129. Rani, C.N.; Karthikeyan, S. Synergic effects on degradation of a mixture of polycyclic aromatic hydrocarbons in a UV slurry photocatalytic membrane reactor and its cost estimation. Chem. Eng. Process.-Process Intensif. 2021, 159, 108179. [Google Scholar] [CrossRef]
  130. Ul Haq, I.; Ahmad, W.; Ahmad, I.; Yaseen, M. Photocatalytic oxidative degradation of hydrocarbon pollutants in refinery wastewater using TiO2 as catalyst. Water Environ. Res. 2020, 92, 2086–2094. [Google Scholar] [CrossRef] [PubMed]
  131. Rahmani, E.; Rahmani, M.; Silab, H.R. TiO2: SiO2 thin film coated annular photoreactor for degradation of oily contamination from waste water. J. Water Process Eng. 2020, 37, 101374. [Google Scholar] [CrossRef]
  132. Cho, I.H.; Kim, L.H.; Zoh, K.D.; Park, J.H.; Kim, H.Y. Solar photocatalytic degradation of groundwater contaminated with petroleum hydrocarbons. Environ. Prog. 2006, 25, 99–109. [Google Scholar] [CrossRef]
  133. Mukwevho, N.; Gusain, R.; Fosso-Kankeu, E.; Kumar, N.; Waanders, F.; Ray, S.S. Removal of naphthalene from simulated wastewater through adsorption-photodegradation by ZnO/Ag/GO nanocomposite. J. Ind. Eng. Chem. 2020, 81, 393–404. [Google Scholar] [CrossRef]
  134. Sheikholeslami, Z.; Yousefi Kebria, D.; Qaderi, F. Investigation of photocatalytic degradation of BTEX in produced water using γ-Fe2O3 nanoparticle. J. Therm. Anal. Calorim. 2019, 135, 1617–1627. [Google Scholar] [CrossRef]
  135. Sekar, A.D.; Muthukumar, H.; Chandrasekaran, N.I.; Matheswaran, M. Photocatalytic degradation of naphthalene using calcined FeZnO/PVA nanofibers. Chemosphere 2018, 205, 610–617. [Google Scholar] [CrossRef]
  136. Rani, M.; Shanker, U. Enhanced photocatalytic degradation of chrysene by Fe2O3@ ZnHCF nanocubes. Chem. Eng. J. 2018, 348, 754–764. [Google Scholar]
  137. Tiburtius, E.R.L.; Peralta-Zamora, P.; Emmel, A. Treatment of gasoline-contaminated waters by advanced oxidation processes. J. Hazard. Mater. 2005, 126, 86–90. [Google Scholar] [CrossRef] [PubMed]
  138. Mohammadifard, Z.; Saboori, R.; Mirbagheri, N.S.; Sabbaghi, S. Heterogeneous photo-Fenton degradation of formaldehyde using MIL-100(Fe) under visible light irradiation. Environ. Pollut. 2019, 251, 783–791. [Google Scholar] [CrossRef] [PubMed]
  139. Veerakumar, P.; Sangili, A.; Saranya, K.; Pandikumar, A.; Lin, K.-C. Palladium and silver nanoparticles embedded on zinc oxide nanostars for photocatalytic degradation of pesticides and herbicides. Chem. Eng. J. 2021, 410, 128434. [Google Scholar] [CrossRef]
  140. Cavalcante, R.P.; de Oliveira, D.M.; da Silva, L.d.M.; Giménez, J.; Esplugas, S.; de Oliveira, S.C.; Dantas, R.F.; Sans, C.; Machulek, A. Evaluation of the main active species involved in the TiO2 photocatalytic degradation of ametryn herbicide and its by-products. J. Environ. Chem. Eng. 2021, 9, 105109. [Google Scholar] [CrossRef]
  141. Samsudin, M.F.R.; Jayabalan, P.J.; Ong, W.-J.; Ng, Y.H.; Sufian, S. Photocatalytic degradation of real industrial poultry wastewater via platinum decorated BiVO4/g-C3N4 photocatalyst under solar light irradiation. J. Photochem. Photobiol. A Chem. 2019, 378, 46–56. [Google Scholar] [CrossRef]
  142. Rani, M.; Shanker, U. Efficient photocatalytic degradation of Bisphenol A by metal ferrites nanoparticles under sunlight. Environ. Technol. Innov. 2020, 19, 100792. [Google Scholar] [CrossRef]
  143. Tsoumachidou, S.; Velegraki, T.; Poulios, I. TiO2 photocatalytic degradation of UV filter para-aminobenzoic acid under artificial and solar illumination. J. Chem. Technol. Biotechnol. 2016, 91, 1773–1781. [Google Scholar] [CrossRef]
  144. Kangralkar, M.V.; Manjanna, J.; Momin, N.; Rane, K.; Nayaka, G.; Kangralkar, V.A. Photocatalytic degradation of hexavalent chromium and different staining dyes by ZnO in aqueous medium under UV light. Environ. Nanotechnol. Monit. Manag. 2021, 16, 100508. [Google Scholar] [CrossRef]
  145. Santhosh, C.; Malathi, A.; Daneshvar, E.; Kollu, P.; Bhatnagar, A. Photocatalytic degradation of toxic aquatic pollutants by novel magnetic 3D-TiO2@ HPGA nanocomposite. Sci. Rep. 2018, 8, 15531. [Google Scholar] [CrossRef]
  146. Chen, P.; Wang, F.; Zhang, Q.; Su, Y.; Shen, L.; Yao, K.; Chen, Z.-F.; Liu, Y.; Cai, Z.; Lv, W. Photocatalytic degradation of clofibric acid by g-C3N4/P25 composites under simulated sunlight irradiation: The significant effects of reactive species. Chemosphere 2017, 172, 193–200. [Google Scholar] [CrossRef]
  147. Truc, N.T.T.; Duc, D.S.; Van Thuan, D.; Al Tahtamouni, T.; Pham, T.-D.; Hanh, N.T.; Tran, D.T.; Nguyen, M.V.; Dang, N.M.; Le Chi, N.T.P. The advanced photocatalytic degradation of atrazine by direct Z-scheme Cu doped ZnO/g-C3N4. Appl. Surf. Sci. 2019, 489, 875–882. [Google Scholar] [CrossRef]
  148. Aoudj, S.; Khelifa, A.; Drouiche, N.; Belkada, R.; Miroud, D. Simultaneous removal of chromium (VI) and fluoride by electrocoagulation–electroflotation: Application of a hybrid Fe-Al anode. Chem. Eng. J. 2015, 267, 153–162. [Google Scholar] [CrossRef]
  149. Kesarla, M.K.; Fuentez-Torres, M.O.; Alcudia-Ramos, M.A.; Ortiz-Chi, F.; Espinosa-González, C.G.; Aleman, M.; Torres-Torres, J.G.; Godavarthi, S. Synthesis of g-C3N4/N-doped CeO2 composite for photocatalytic degradation of an herbicide. J. Mater. Res. Technol. 2019, 8, 1628–1635. [Google Scholar] [CrossRef]
  150. Wongcharoen, S.; Panomsuwan, G. Easy synthesis of TiO2 hollow fibers using kapok as a biotemplate for photocatalytic degradation of the herbicide paraquat. Mater. Lett. 2018, 228, 482–485. [Google Scholar] [CrossRef]
  151. Mansourian, R.; Mousavi, S.M.; Alizadeh, S.; Sabbaghi, S. CeO2/TiO2/SiO2 nanocatalyst for the photocatalytic and sonophotocatalytic degradation of chlorpyrifos. Can. J. Chem. Eng. 2022, 100, 451–464. [Google Scholar] [CrossRef]
  152. Song, H.; Shao, J.; He, Y.; Liu, B.; Zhong, X. Natural organic matter removal and flux decline with PEG–TiO2-doped PVDF membranes by integration of ultrafiltration with photocatalysis. J. Membr. Sci. 2012, 405, 48–56. [Google Scholar] [CrossRef]
  153. Wang, F.; Chen, Z.; Zhu, Z.; Guo, J. Construction of visible light responsive ZnO/N-g-C3N4 composite membranes for antibiotics degradation. J. Mater. Res. Technol. 2022, 17, 1696–1706. [Google Scholar] [CrossRef]
  154. Athanasekou, C.P.; Moustakas, N.G.; Morales-Torres, S.; Pastrana-Martínez, L.M.; Figueiredo, J.L.; Faria, J.L.; Silva, A.M.; Dona-Rodriguez, J.M.; Romanos, G.E.; Falaras, P. Ceramic photocatalytic membranes for water filtration under UV and visible light. Appl. Catal. B Environ. 2015, 178, 12–19. [Google Scholar] [CrossRef]
  155. Chen, D.; Cheng, Y.; Zhou, N.; Chen, P.; Wang, Y.; Li, K.; Huo, S.; Cheng, P.; Peng, P.; Zhang, R. Photocatalytic degradation of organic pollutants using TiO2-based photocatalysts: A review. J. Clean. Prod. 2020, 268, 121725. [Google Scholar] [CrossRef]
  156. Ahmad, R.; Lee, C.S.; Kim, J.H.; Kim, J. Partially coated TiO2 on Al2O3 membrane for high water flux and photodegradation by novel filtration strategy in photocatalytic membrane reactors. Chem. Eng. Res. Des. 2020, 163, 138–148. [Google Scholar] [CrossRef]
  157. Wang, F.; Xu, J.; Wang, Z.; Lou, Y.; Pan, C.; Zhu, Y. Unprecedentedly efficient mineralization performance of photocatalysis-self-Fenton system towards organic pollutants over oxygen-doped porous g-C3N4 nanosheets. Appl. Catal. B Environ. 2022, 312, 121438. [Google Scholar] [CrossRef]
  158. Čizmić, M.; Ljubas, D.; Rožman, M.; Ašperger, D.; Ćurković, L.; Babić, S. Photocatalytic degradation of azithromycin by nanostructured TiO2 film: Kinetics, degradation products, and toxicity. Materials 2019, 12, 873. [Google Scholar] [CrossRef] [PubMed]
  159. Shang, C.; Bu, J.; Song, C. Preparation, Antimicrobial Properties under Different Light Sources, Mechanisms and Applications of TiO2: A Review. Materials 2022, 15, 5820. [Google Scholar] [CrossRef]
  160. Singh, A.; Ramachandran, S.K.; Gumpu, M.B.; Zsuzsanna, L.; Veréb, G.; Kertész, S.; Gangasalam, A. Titanium dioxide doped hydroxyapatite incorporated photocatalytic membranes for the degradation of chloramphenicol antibiotic in water. J. Chem. Technol. Biotechnol. 2020, 96, 1057–1066. [Google Scholar] [CrossRef]
  161. Zakeritabar, S.F.; Jahanshahi, M.; Peyravi, M.; Akhtari, J. Photocatalytic study of nanocomposite membrane modified by CeF3 catalyst for pharmaceutical wastewater treatment. J. Environ. Health Sci. Eng. 2020, 18, 1151–1161. [Google Scholar] [CrossRef] [PubMed]
  162. Shaku, K.; Dlamini, L.; Malinga, S. Highly efficient photocatalytic hyperbranched polyethyleneimine/bismuth vanadate membranes for the degradation of triclosan. Int. J. Environ. Sci. Technol. 2020, 17, 3297–3312. [Google Scholar] [CrossRef]
  163. Yu, S.; Wang, Y.; Sun, F.; Wang, R.; Zhou, Y. Novel mpg-C3N4/TiO2 nanocomposite photocatalytic membrane reactor for sulfamethoxazole photodegradation. Chem. Eng. J. 2018, 337, 183–192. [Google Scholar] [CrossRef]
  164. Sun, S.; Yao, H.; Fu, W.; Xue, S.; Zhang, W. Enhanced degradation of antibiotics by photo-fenton reactive membrane filtration. J. Hazard. Mater. 2020, 386, 121955. [Google Scholar] [CrossRef]
  165. Koe, W.S.; Chong, W.C.; Pang, Y.L.; Koo, C.H.; Ebrahim, M.; Mohammad, A.W. Novel nitrogen and sulphur co-doped carbon quantum dots/titanium oxide photocatalytic membrane for in-situ degradation and removal of pharmaceutical compound. J. Water Process Eng. 2020, 33, 101068. [Google Scholar] [CrossRef]
  166. Espındola, C.; Szymański, K.; Cristóvão, R.O.; Mendes, A.; Vilar, V.J.P.; Mozia, S. Performance of hybrid systems coupling advanced oxidation processes and ultrafiltration for oxytetracycline removal. Catal Today 2019, 328, 274–280. [Google Scholar] [CrossRef]
  167. Pastrana-Martinez, L.M.; Morales-Torres, S.; Figueiredo, J.L.; Faria, J.L.; Silva, A.M. Graphene oxide based ultrafiltration membranes for photocatalytic degradation of organic pollutants in salty water. Water Res. 2015, 77, 179–190. [Google Scholar] [CrossRef]
  168. Chakraborty, S.; Loutatidou, S.; Palmisano, G.; Kujawa, J.; Mavukkandy, M.O.; Al-Gharabli, S.; Curcio, E.; Arafat, H.A. Photocatalytic hollow fiber membranes for the degradation of pharmaceutical compounds in wastewater. J. Environ. Chem. Eng. 2017, 5, 5014–5024. [Google Scholar] [CrossRef]
  169. Sarasidis, V.C.; Plakas, K.V.; Patsios, S.I.; Karabelas, A.J. Investigation of diclofenac degradation in a continuous photo-catalytic membrane reactor. Influence of operating parameters. Chem. Eng. J. 2014, 239, 299–311. [Google Scholar] [CrossRef]
  170. Fischer, K.; Grimm, M.; Meyers, J.; Dietrich, C.; Gläser, R.; Schulze, A. Photoactive microfiltration membranes via directed synthesis of TiO2 nanoparticles on the polymer surface for removal of drugs from water. J. Membr. Sci. 2015, 478, 49–57. [Google Scholar] [CrossRef]
  171. Darowna, D.; Grondzewska, S.; Morawski, A.W.; Mozia, S. Removal of non-steroidal anti-inflammatory drugs from primary and secondary effluents in a photocatalytic membrane reactor. J. Chem. Technol. Biotechnol. 2014, 89, 1265–1273. [Google Scholar] [CrossRef]
  172. Molinari, R.; Pirillo, F.; Loddo, V.; Palmisano, L. Heterogeneous photocatalytic degradation of pharmaceuticals in water by using polycrystalline TiO2 and a nanofiltration membrane reactor. Catal. Today 2006, 118, 205–213. [Google Scholar] [CrossRef]
  173. Li, W.; Li, B.; Meng, M.; Cui, Y.; Wu, Y.; Zhang, Y.; Dong, H.; Feng, Y. Bimetallic Au/Ag decorated TiO2 nanocomposite membrane for enhanced photocatalytic degradation of tetracycline and bactericidal efficiency. Appl. Surf. Sci. 2019, 487, 1008–1017. [Google Scholar] [CrossRef]
  174. Plakas, K.V.; Sarasidis, V.C.; Patsios, S.I.; Lambropoulou, D.A.; Karabelas, A.J. Novel pilot scale continuous photocatalytic membrane reactor for removal of organic micropollutants from water. Chem. Eng. J. 2016, 304, 335–343. [Google Scholar] [CrossRef]
  175. Wang, C.; Wu, Y.; Lu, J.; Zhao, J.; Cui, J.; Wu, X.; Yan, Y.; Huo, P. Bioinspired synthesis of photocatalytic nanocomposite membranes based on synergy of Au-TiO2 and polydopamine for degradation of tetracycline under visible light. ACS Appl. Mater. Interfaces 2017, 9, 23687–23697. [Google Scholar] [CrossRef]
  176. Boopathy, G.; Gangasalam, A.; Mahalingam, A. Photocatalytic removal of organic pollutants and self-cleaning performance of PES membrane incorporated sulfonated graphene oxide/ZnO nanocomposite. J. Chem. Technol. Biotechnol. 2020, 95, 3012–3023. [Google Scholar] [CrossRef]
  177. Zakeritabar, S.F.; Jahanshahi, M.; Peyravi, M. Photocatalytic behavior of induced membrane by ZrO2–SnO2 nanocomposite for pharmaceutical wastewater treatment. Catal. Lett. 2018, 148, 882–893. [Google Scholar] [CrossRef]
  178. Gao, B.; Chen, W.; Liu, J.; An, J.; Wang, L.; Zhu, Y.; Sillanpää, M. Continuous removal of tetracycline in a photocatalytic membrane reactor (PMR) with ZnIn2S4 as adsorption and photocatalytic coating layer on PVDF membrane. J. Photochem. Photobiol. A Chem. 2018, 364, 732–739. [Google Scholar] [CrossRef]
  179. Horovitz, I.; Avisar, D.; Baker, M.A.; Grilli, R.; Lozzi, L.; Di Camillo, D.; Mamane, H. Carbamazepine degradation using a N-doped TiO2 coated photocatalytic membrane reactor: Influence of physical parameters. J. Hazard. Mater. 2016, 310, 98–107. [Google Scholar] [CrossRef] [PubMed]
  180. Dzinun, H.; Ichikawa, Y.; Mitsuhiro, H.; Zhang, Q. Efficient immobilised TiO2 in polyvinylidene fluoride (PVDF) membrane for photocatalytic degradation of methylene blue. J. Membr. Sci. Res. 2020, 6, 188–195. [Google Scholar]
  181. 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]
  182. Ma, N.; Zhang, Y.; Quan, X.; Fan, X.; Zhao, H. Performing a microfiltration integrated with photocatalysis using an Ag-TiO2/HAP/Al2O3 composite membrane for water treatment: Evaluating effectiveness for humic acid removal and anti-fouling properties. Water Res. 2010, 44, 6104–6114. [Google Scholar] [CrossRef] [PubMed]
  183. Rani, C.N.; Karthikeyan, S. Investigation of Naphthalene Removal from Aqueous Solutions in an Integrated Slurry Photocatalytic Membrane Reactor: Effect of Operating Parameters, Identification of Intermediates, and Response Surface Approach. Polycycl. Aromat. Compd. 2019, 41, 805–824. [Google Scholar] [CrossRef]
  184. Moslehyani, A.; Ismail, A.; Othman, M.; Matsuura, T. Hydrocarbon degradation and separation of bilge water via a novel TiO2-HNTs/PVDF-based photocatalytic membrane reactor (PMR). RSC Adv. 2015, 5, 14147–14155. [Google Scholar] [CrossRef]
  185. Goei, R.; Lim, T.-T. Ag-decorated TiO2 photocatalytic membrane with hierarchical architecture: Photocatalytic and anti-bacterial activities. Water Res. 2014, 59, 207–218. [Google Scholar] [CrossRef]
  186. Kazemi, M.; Jahanshahi, M.; Peyravi, M. Chitosan-sodium alginate multilayer membrane developed by Fe0@ WO3 nanoparticles: Photocatalytic removal of hexavalent chromium. Carbohydr. Polym. 2018, 198, 164–174. [Google Scholar] [CrossRef]
  187. Tran, M.L.; Fu, C.-C.; Chiang, L.-Y.; Hsieh, C.-T.; Liu, S.-H.; Juang, R.-S. Immobilization of TiO2 and TiO2-GO hybrids onto the surface of acrylic acid-grafted polymeric membranes for pollutant removal: Analysis of photocatalytic activity. J. Environ. Chem. Eng. 2020, 8, 104422. [Google Scholar] [CrossRef]
  188. Shi, Y.; Wan, D.; Huang, J.; Liu, Y.; Li, J. Stable LBL self-assembly coating porous membrane with 3D heterostructure for enhanced water treatment under visible light irradiation. Chemosphere 2020, 252, 126581. [Google Scholar] [CrossRef] [PubMed]
  189. Vatanpour, V.; Karami, A.; Sheydaei, M. Central composite design optimization of Rhodamine B degradation using TiO2 nanoparticles/UV/PVDF process in continuous submerged membrane photoreactor. Chem. Eng. Process. Process Intensif. 2017, 116, 68–75. [Google Scholar] [CrossRef]
  190. Kuvarega, A.T.; Khumalo, N.; Dlamini, D.; Mamba, B.B. Polysulfone/N, Pd co-doped TiO2 composite membranes for photocatalytic dye degradation. Sep. Purif. Technol. 2018, 191, 122–133. [Google Scholar] [CrossRef]
  191. Wang, X.; Wang, G.; Chen, S.; Fan, X.; Quan, X.; Yu, H. Integration of membrane filtration and photoelectrocatalysis on g-C3N4/CNTs/Al2O3 membrane with visible-light response for enhanced water treatment. J. Membr. Sci. 2017, 541, 153–161. [Google Scholar] [CrossRef]
  192. Berger, T.; Regmi, C.; Schäfer, A.; Richards, B. Photocatalytic degradation of organic dye via atomic layer deposited TiO2 on ceramic membranes in single-pass flow-through operation. J. Membr. Sci. 2020, 604, 118015. [Google Scholar] [CrossRef]
  193. Zhao, G.; Zou, J.; Chen, X.; Zhang, T.; Yu, J.; Zhou, S.; Li, C.; Jiao, F. Integration of Microfiltration and Visible-Light-Driven Photocatalysis on a ZnWO4 Nanoparticle/Nickel–Aluminum-Layered Double Hydroxide Membrane for Enhanced Water Purification. Ind. Eng. Chem. Res. 2020, 59, 6479–6487. [Google Scholar] [CrossRef]
  194. Gao, M.; Feng, J.; He, F.; Zeng, W.; Wang, X.; Ren, Y.; Wei, T. Carbon microspheres work as an electron bridge for degrading high concentration MB in CoFe2O4@ carbon microsphere/g-C3N4 with a hierarchical sandwich-structure. Appl. Surf. Sci. 2020, 507, 145167. [Google Scholar] [CrossRef]
  195. Gupta, A.; Pandey, O. NbC/C heterojunction for efficient photodegradation of methylene blue under visible irradiation. Sol. Energy 2019, 183, 398–409. [Google Scholar] [CrossRef]
  196. Wang, M.; Zhang, Y.; Yu, G.; Zhao, J.; Chen, X.; Yan, F.; Li, J.; Yin, Z.; He, B. Monolayer porphyrin assembled SPSf/PES membrane reactor for degradation of dyes under visible light irradiation coupling with continuous filtration. J. Taiwan Inst. Chem. Eng. 2020, 109, 62–70. [Google Scholar] [CrossRef]
  197. Yu, Z.; Zeng, H.; Min, X.; Zhu, X. High-performance composite photocatalytic membrane based on titanium dioxide nanowire/graphene oxide for water treatment. J. Appl. Polym. Sci. 2020, 137, 48488. [Google Scholar] [CrossRef]
  198. Salim, N.E.; Nor, N.; Jaafar, J.; Ismail, A.; Qtaishat, M.; Matsuura, T.; Othman, M.; Rahman, M.A.; Aziz, F.; Yusof, N. Effects of hydrophilic surface macromolecule modifier loading on PES/Og-C3N4 hybrid photocatalytic membrane for phenol removal. Appl. Surf. Sci. 2019, 465, 180–191. [Google Scholar] [CrossRef]
  199. Ashar, A.; Bhatti, I.A.; Ashraf, M.; Tahir, A.A.; Aziz, H.; Yousuf, M.; Ahmad, M.; Mohsin, M.; Bhutta, Z.A. Fe3+@ ZnO/polyester based solar photocatalytic membrane reactor for abatement of RB5 dye. J. Clean. Prod. 2020, 246, 119010. [Google Scholar] [CrossRef]
  200. Yang, C.; Han, N.; Zhang, W.; Wang, W.; Li, W.; Xia, B.; Han, C.; Cui, Z.; Zhang, X. Adhesive-free in situ synthesis of a coral-like TiO2@ PPS microporous membrane for visible-light photocatalysis. Chem. Eng. J. 2019, 374, 1382–1393. [Google Scholar] [CrossRef]
  201. Wang, M.; Yang, G.; Jin, P.; Tang, H.; Wang, H.; Chen, Y. Highly hydrophilic poly (vinylidene fluoride)/meso-titania hybrid mesoporous membrane for photocatalytic membrane reactor in water. Sci. Rep. 2016, 6, 19148. [Google Scholar] [CrossRef] [PubMed]
  202. Zhang, R.; Cai, Y.; Zhu, X.; Han, Q.; Zhang, T.; Liu, Y.; Li, Y.; Wang, A. A novel photocatalytic membrane decorated with PDA/RGO/Ag3PO4 for catalytic dye decomposition. Colloids Surf. A Physicochem. Eng. Asp. 2019, 563, 68–76. [Google Scholar] [CrossRef]
  203. Wang, X.; Shi, F.; Huang, W.; Fan, C. Synthesis of high quality TiO2 membranes on alumina supports and their photocatalytic activity. Thin Solid Film. 2012, 520, 2488–2492. [Google Scholar] [CrossRef]
  204. Wu, X.-Q.; Shen, J.-S.; Zhao, F.; Shao, Z.-D.; Zhong, L.-B.; Zheng, Y.-M. Flexible electrospun MWCNTs/Ag3PO4/PAN ternary composite fiber membranes with enhanced photocatalytic activity and stability under visible-light irradiation. J. Mater. Sci. 2018, 53, 10147–10159. [Google Scholar] [CrossRef]
  205. Yu, Z.; Min, X.; Li, F.; Yin, D.; Peng, Y.; Zeng, G. A mussel-inspired method to fabricate a novel reduced graphene oxide/Bi12O17Cl2 composites membrane for catalytic degradation and oil/water separation. Polym. Adv. Technol. 2019, 30, 101–109. [Google Scholar] [CrossRef]
  206. Daels, N.; Radoicic, M.; Radetic, M.; Van Hulle, S.W.; De Clerck, K. Functionalisation of electrospun polymer nanofibre membranes with TiO2 nanoparticles in view of dissolved organic matter photodegradation. Sep. Purif. Technol. 2014, 133, 282–290. [Google Scholar] [CrossRef]
  207. Fischer, K.; Gläser, R.; Schulze, A. Nanoneedle and nanotubular titanium dioxide–PES mixed matrix membrane for photocatalysis. Appl. Catal. B Environ. 2014, 160, 456–464. [Google Scholar] [CrossRef]
  208. Ong, C.B.; Mohammad, A.W.; Ng, L.Y. Integrated adsorption-solar photocatalytic membrane reactor for degradation of hazardous Congo red using Fe-doped ZnO and Fe-doped ZnO/rGO nanocomposites. Environ. Sci. Pollut. Res. 2019, 26, 33856–33869. [Google Scholar] [CrossRef] [PubMed]
  209. Alyarnezhad, S.; Marino, T.; Parsa, J.B.; Galiano, F.; Ursino, C.; Garcìa, H.; Puche, M.; Figoli, A. Polyvinylidene fluoride-graphene oxide membranes for dye removal under visible light irradiation. Polymers 2020, 12, 1509. [Google Scholar] [CrossRef]
  210. Lyubimenko, R.; Busko, D.; Richards, B.S.; Schäfer, A.I.; Turshatov, A. Efficient photocatalytic removal of methylene blue using a metalloporphyrin–poly (vinylidene fluoride) hybrid membrane in a flow-through reactor. ACS Appl. Mater. Interfaces 2019, 11, 31763–31776. [Google Scholar] [CrossRef]
  211. Zhao, H.; Chen, S.; Quan, X.; Yu, H.; Zhao, H. Integration of microfiltration and visible-light-driven photocatalysis on g-C3N4 nanosheet/reduced graphene oxide membrane for enhanced water treatment. Appl. Catal. B Environ. 2016, 194, 134–140. [Google Scholar] [CrossRef]
  212. Xu, H.; Ding, M.; Chen, W.; Li, Y.; Wang, K. Nitrogen–doped GO/TiO2 nanocomposite ultrafiltration membranes for improved photocatalytic performance. Sep. Purif. Technol. 2018, 195, 70–82. [Google Scholar] [CrossRef]
  213. Laohaprapanon, S.; Vanderlipe, A.D.; Doma, B.T., Jr.; You, S.-J. Self-cleaning and antifouling properties of plasma-grafted poly (vinylidene fluoride) membrane coated with ZnO for water treatment. J. Taiwan Inst. Chem. Eng. 2017, 70, 15–22. [Google Scholar] [CrossRef]
  214. Li, B.; Chu, J.; Li, Y.; Meng, M.; Cui, Y.; Li, Q.; Feng, Y. Preparation and Performance of Visible-Light-Driven Bi2O3/ZnS Heterojunction Functionalized Porous CA Membranes for Effective Degradation of Rhodamine B. Phys. Status Solidi (A) 2018, 215, 1701061. [Google Scholar] [CrossRef]
  215. Bai, H.; Zan, X.; Juay, J.; Sun, D.D. Hierarchical heteroarchitectures functionalized membrane for high efficient water purification. J. Membr. Sci. 2015, 475, 245–251. [Google Scholar] [CrossRef]
  216. Rani, C.N.; Karthikeyan, S. Feasibility study of acenaphthene degradation in a novel slurry UV photocatalytic membrane reactor: Effect of operating parameters and optimization using response surface modeling. Chem. Eng. Process.-Process Intensif. 2020, 155, 108051. [Google Scholar] [CrossRef]
  217. Nascimben Santos, E.; Agoston, A.; Kertész, S.; Hodúr, C.; László, Z.; Pap, Z.; Kása, Z.; Alapi, T.; Krishnan, S.G.; Arthanareeswaran, G. Investigation of the applicability of TiO2, BiVO4, and WO3 nanomaterials for advanced photocatalytic membranes used for oil-in-water emulsion separation. Asia-Pac. J. Chem. Eng. 2020, 15, e2549. [Google Scholar] [CrossRef]
  218. Zhou, J.; Zhao, Z.; Wang, Y.; Ding, Z.; Xu, X.; Peng, W.; Fan, J.; Zhou, X.; Liu, J. BiOCl0.875Br0.125/polydopamine functionalized PVDF membrane for highly efficient visible-light-driven photocatalytic degradation of roxarsone and simultaneous arsenic immobilization. Chem. Eng. J. 2020, 402, 126048. [Google Scholar] [CrossRef]
  219. Gao, Y.; Yan, N.; Jiang, C.; Xu, C.; Yu, S.; Liang, P.; Zhang, X.; Liang, S.; Huang, X. Filtration-enhanced highly efficient photocatalytic degradation with a novel electrospun rGO@ TiO2 nanofibrous membrane: Implication for improving photocatalytic efficiency. Appl. Catal. B Environ. 2020, 268, 118737. [Google Scholar] [CrossRef]
  220. Shareef, U.; Othman, M.H.D.; Ismail, A.F.; Jilani, A. Facile removal of bisphenol A from water through novel Ag-doped TiO2 photocatalytic hollow fiber ceramic membrane. J. Aust. Ceram. Soc. 2020, 56, 29–39. [Google Scholar] [CrossRef]
  221. Lin, Y.-C.; Wang, D.K.; Liu, J.-Y.; Niaei, A.; Tseng, H.-H. Low band-gap energy photocatalytic membrane based on SrTiO3–Cr and PVDF substrate: BSA protein degradation and separation application. J. Membr. Sci. 2019, 586, 326–337. [Google Scholar] [CrossRef]
  222. Shi, Y.; Huang, J.; Zeng, G.; Cheng, W.; Hu, J.; Shi, L.; Yi, K. 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]
  223. Leong, S.; Low, Z.X.; Liu, Q.; Hapgood, K.; Zhang, X.; Wang, H. Preparation of supported photocatalytic membrane from mesoporous titania spheres for humic acid removal from wastewater. Asia-Pac. J. Chem. Eng. 2016, 11, 611–619. [Google Scholar] [CrossRef]
  224. Golshenas, A.; Sadeghian, Z.; Ashrafizadeh, S.N. Performance evaluation of a ceramic-based photocatalytic membrane reactor for treatment of oily wastewater. J. Water Process Eng. 2020, 36, 101186. [Google Scholar] [CrossRef]
  225. Xu, H.; Li, Y.; Ding, M.; Chen, W.; Wang, K.; Lu, C. Engineered photocatalytic material membrane assemblies for removing nitrate from water. ACS Sustain. Chem. Eng. 2018, 6, 7042–7051. [Google Scholar] [CrossRef]
  226. Ong, C.; Lau, W.; Goh, P.; Ng, B.; Ismail, A. Investigation of submerged membrane photocatalytic reactor (sMPR) operating parameters during oily wastewater treatment process. Desalination 2014, 353, 48–56. [Google Scholar] [CrossRef]
  227. Xu, Z.; Wu, T.; Shi, J.; Teng, K.; Wang, W.; Ma, M.; Li, J.; Qian, X.; Li, C.; Fan, J. Photocatalytic antifouling PVDF ultrafiltration membranes based on synergy of graphene oxide and TiO2 for water treatment. J. Membr. Sci. 2016, 520, 281–293. [Google Scholar] [CrossRef]
  228. Mungondori, H.H.; Tichagwa, L.; Katwire, D.M.; Aoyi, O. Preparation of photo-catalytic copolymer grafted asymmetric membranes (N-TiO2-PMAA-g-PVDF/PAN) and their application on the degradation of bentazon in water. Iran. Polym. J. 2016, 25, 135–144. [Google Scholar] [CrossRef]
  229. Lin, H.; Zhang, M.; Wang, F.; Meng, F.; Liao, B.-Q.; Hong, H.; Chen, J.; Gao, W. A critical review of extracellular polymeric substances (EPSs) in membrane bioreactors: Characteristics, roles in membrane fouling and control strategies. J. Membr. Sci. 2014, 460, 110–125. [Google Scholar] [CrossRef]
  230. Zhang, W.; Hao, T. Insights into the role of concentration polarization on the membrane fouling and cleaning during the aerobic granular sludge filtration process. Sci. Total Environ. 2022, 813, 151871. [Google Scholar] [CrossRef] [PubMed]
  231. Augugliaro, V.; Litter, M.; Palmisano, L.; Soria, J. The combination of heterogeneous photocatalysis with chemical and physical operations: A tool for improving the photoprocess performance. J. Photochem. Photobiol. C Photochem. Rev. 2006, 7, 127–144. [Google Scholar] [CrossRef]
  232. Gao, W.; Liang, H.; Ma, J.; Han, M.; Chen, Z.-l.; Han, Z.-s.; Li, G.-b. Membrane fouling control in ultrafiltration technology for drinking water production: A review. Desalination 2011, 272, 1–8. [Google Scholar] [CrossRef]
  233. Zheng, H.; Zhu, M.; Wang, D.; Zhou, Y.; Sun, X.; Jiang, S.; Li, M.; Xiao, C.; Zhang, D.; Zhang, L. Surface modification of PVDF membrane by CNC/Cu-MOF-74 for enhancing antifouling property. Sep. Purif. Technol. 2023, 306, 122599. [Google Scholar] [CrossRef]
  234. Mozia, S. Photocatalytic membrane reactors (PMRs) in water and wastewater treatment. A review. Sep. Purif. Technol. 2010, 73, 71–91. [Google Scholar] [CrossRef]
  235. Mendret, J.; Hatat-Fraile, M.; Rivallin, M.; Brosillon, S. Hydrophilic composite membranes for simultaneous separation and photocatalytic degradation of organic pollutants. Sep. Purif. Technol. 2013, 111, 9–19. [Google Scholar] [CrossRef]
  236. Wang, P.; Fane, A.G.; Lim, T.-T. Evaluation of a submerged membrane vis-LED photoreactor (sMPR) for carbamazepine degradation and TiO2 separation. Chem. Eng. J. 2013, 215, 240–251. [Google Scholar] [CrossRef]
  237. Zhang, J.; Wang, L.; Zhang, G.; Wang, Z.; Xu, L.; Fan, Z. Influence of azo dye-TiO2 interactions on the filtration performance in a hybrid photocatalysis/ultrafiltration process. J. Colloid Interface Sci. 2013, 389, 273–283. [Google Scholar] [CrossRef]
  238. Chin, J.Y.; Ahmad, A.L.; Low, S.C. Evolution of photocatalytic membrane for antibiotics degradation: Perspectives and insights for sustainable environmental remediation. J. Water Process Eng. 2023, 51, 103342. [Google Scholar] [CrossRef]
Figure 1. The number of articles published annually on the topic of photocatalytic membrane reactors based on Scopus (searched “Photocatalytic membrane reactor”).
Figure 1. The number of articles published annually on the topic of photocatalytic membrane reactors based on Scopus (searched “Photocatalytic membrane reactor”).
Materials 16 03526 g001
Figure 2. (a) Mechanism of photocatalytic degradation of pollutants, and (b) mechanism of cefixime photocatalytic degradation via Fe2O3@TiO2 photocatalyst [13]. Adapted with permission from Elsevier. Copyright 2023.
Figure 2. (a) Mechanism of photocatalytic degradation of pollutants, and (b) mechanism of cefixime photocatalytic degradation via Fe2O3@TiO2 photocatalyst [13]. Adapted with permission from Elsevier. Copyright 2023.
Materials 16 03526 g002
Figure 4. (a) Cross-section SEM images of PES membrane [76]. Adapted with permission from Elsevier. Copyright 2008, (b) top-view SEM images of PVDF membranes with irregular porous surface [75]. Adapted with permission from Elsevier. Copyright 2007, and (c) TEM images showing components of a hierarchical layer of a TiO2 nanowire membrane [77]. Adapted with permission from water and research technology. Copyright 2015.
Figure 4. (a) Cross-section SEM images of PES membrane [76]. Adapted with permission from Elsevier. Copyright 2008, (b) top-view SEM images of PVDF membranes with irregular porous surface [75]. Adapted with permission from Elsevier. Copyright 2007, and (c) TEM images showing components of a hierarchical layer of a TiO2 nanowire membrane [77]. Adapted with permission from water and research technology. Copyright 2015.
Materials 16 03526 g004
Figure 5. Two different configurations of PMRs (a) photocatalyst immobilized in/on the membrane, (b) photocatalyst in suspension media.
Figure 5. Two different configurations of PMRs (a) photocatalyst immobilized in/on the membrane, (b) photocatalyst in suspension media.
Materials 16 03526 g005
Figure 6. (a) Submerged membrane in a slurry reactor, (b) Slurry reactor followed by a membrane filtration unit, (c) Photocatalytic membrane, and (d) Submerged membrane in photocatalytic-coated reactor.
Figure 6. (a) Submerged membrane in a slurry reactor, (b) Slurry reactor followed by a membrane filtration unit, (c) Photocatalytic membrane, and (d) Submerged membrane in photocatalytic-coated reactor.
Materials 16 03526 g006
Figure 7. (a) FTIR spectra of iron glycolate, pure Fe2O3 microflowers and 50%TiO2/Fe2O3 core-shell and (b) XRD patterns for iron glycolate, pure Fe2O3 microflowers and TiO2/Fe2O3 core-shells [99]. Adapted with permission from Elsevier. Copyright 2017.
Figure 7. (a) FTIR spectra of iron glycolate, pure Fe2O3 microflowers and 50%TiO2/Fe2O3 core-shell and (b) XRD patterns for iron glycolate, pure Fe2O3 microflowers and TiO2/Fe2O3 core-shells [99]. Adapted with permission from Elsevier. Copyright 2017.
Materials 16 03526 g007
Table 1. Membranes for water decontamination.
Table 1. Membranes for water decontamination.
MembraneMaterialManufacturerThickness (mm)Pore Size (μm)P (bar)T (K)PollutantMembrane Separation (%)Ref.
Commercial spiral wound polyamide nano filterTFC---8-COD98[52]
AGS reactor, UP and NFPES, PESH, and PA-- 0.5293.15COD51.33, 90.48 and 99.26[53]
Hollow fiber UF membrane PVC-0.1850.01–0.11298.15methyl green dye (MG)94.79[54]
NF membranepiperazine based polyamide, proprietary (cross-linked modified polyacrylonitrile), polyethersulfoneToray chemical, Koch membrane systems, Nadir-0.130-copper-refining sulfuric acid wastewater95[55]
A sheet nano-porous membranePANSepro Membranes of USA.Top layer 0.1–0.5 and
Sublayer 100–150
0.014318.15TSS, TDS, content of oil and grease, COD and BOD5100, 44.4, 99.9, 80.3 and 76.9[56]
Multichannel tubularCeramic membraneJiangsu Jiuwu HiTech Co. Ltd., Nanjing, China 0.052233.15suspended solid, turbidity, and total phosphorus100, 99.20, and 80.21,[57]
SW30 membrane
BW30
NF270
PA---69
41
41
318.15COD
and
TPh
(95.18, 91.15, 80.11) and (98.02, 96.06, 27.08)[58]
Difluoride hollow fiber membrane modulePVDFM/s. TECHNIC, India-0.10.5–1-TOC64[59]
Table 3. Review on Photocatalytic degradation of pharmaceutical compounds.
Table 3. Review on Photocatalytic degradation of pharmaceutical compounds.
PollutantPollutant Concentration (mg·L−1)PhotocatalystLight SourceTime (min)Degradation (%)Ref.
Carbamazepine (CBZ)2.5Z-scheme FeS2/Fe2O3 /hexavalent chromiumsimulated visible light3065[100]
Penicillin G (PG)5TiO2/P25UV-A sunlight (365 nm)15072.72[101]
Pentachlorophenol (PCP)10FeNi3/SiO2/ZnOIrradiation
of solar light
18092.47[102]
Ciprofloxacin (CIP)10Nano-ZnOUV3096[103]
Diclofenac (DCL)1ZnOLP Hg (150 W)3068[104]
Sulfamethoxazole25TiO2UV15061.28[105]
Sulfamethoxazole25WO3UV15043.3[105]
Dexamethasone (DXM)5ZrO2/WO3halogen80100[106]
Metronidazole (MNZ)80illuminated TiO2UV (125 W)18096.55[107]
Levofloxacin50BiOCl(25)/BiOBr-Bi24O31Br10(75) type-II nano heterojunctionHalogen (400 W)18080.2[108]
Ofloxacin50BiOCl(25)/BiOBr-Bi24O31Br10(75) type-II nano heterojunctionHalogen (400 W)18078.3[108]
Amoxicillin
metformin
10TiO2125 W low-pressure mercury vapor lamp15090
98
[109]
Paracetamol100La-doped ZnOfluorescent lamps (20 W)18099[110]
Cefixime20.5α-Fe2O3@TiO2Visible10398[13]
Cefixime47g C3N4/TiO2Visible11384[111]
Table 4. Review on Photocatalytic degradation of dye compounds.
Table 4. Review on Photocatalytic degradation of dye compounds.
PollutantPollutant Concentration (mg/L)PhotocatalystSynthesis MethodLight SourceTime (min)Degradation (%)Ref.
Acid fuchsine (AF)120TiO2/ACFSol–gel-adsorptionHg lamp (500 W)3077[116]
4-chlorophenol (4-CP)15ZTPGHigh-temperature refluxingUV18094.8[117]
Methylene blue (MB)20ZTPGHigh-temperature refluxingUV9098.05[117]
Basic Red 51 (BR51)1ZnO-sunlight irradiation33089.01[118]
Methylene blue (MB)9.56ZnO-UV3093[119]
Methyl Orange (MO)5TiO2/MgO/ChitosanHydrogelUV (125 W)9082.4[120]
Alizarin Red S (ARS)5TiO2/MgO/ChitosanHydrogelUV (125 W)9041.8[120]
Methyl Orange (MO)15Ni@FP + NaBH4 (10 mg)-UV593.40[121]
Tartrazine50CuCr2O4self-combustionHg lamp (125 W)12099.6[122]
Acid Brown 9810CS-ZnS-NPsco-precipitationUV (254 nm)16592.6[123]
Acid Black 23410CS-ZnS-NPsco-precipitationUV (254 nm)10096.7[123]
Methylene blue (MB)3.2MnTiO3/TiO2sol-gelsunlight24075[124]
Tartrazine10TiO2–Chitosan-Sunlight18099.37[125]
1,2-Dichloroethane250PAni-TiO2deposition oxidative polymerizationvisible24088.84[126]
Table 5. Review on Photocatalytic degradation of hydrocarbons.
Table 5. Review on Photocatalytic degradation of hydrocarbons.
PollutantPollutant Concentration (mg/L)PhotocatalystSynthesis MethodLight SourceTime (min)Degradation (%)Ref.
Benzene
Toluene
Phenol
Naphthalene
65TiO2-UV (400 W)9092
98.8
91.5
93
[130]
Paraffin500TiO2/SiO2 thin filmsol-gelUV18085[131]
BTEX
TPHs
60.8TiO2-solar light240>70[132]
Naphthalene50ZnO/Ag/GO nanocomposit-Xe lamp (250 W)2080[133]
BTEX600 γ -Fe2O3 nanoparticle-UV light (100 W)9097[134]
Naphthalene40Calcinated Fe-doped ZnO/PVA nanofibers-UV light (16 W)36096[135]
Chrysene2Fe2O3@ZnHCF nanocubes-sunlight144092[136]
Benzene, toluene and xylenes (BTX) and gasoline-contaminated waters20TiO2-Fenton system-medium-pressure mercury vapor lamp (125 W)9075[137]
Formaldehyde700MIL-100(Fe)solvothermalVisible11993[138]
Table 6. Review on Photocatalytic degradation of other pollutants.
Table 6. Review on Photocatalytic degradation of other pollutants.
PollutantPollutant Concentration (mg/L)PhotocatalystSynthesis MethodLight SourceTime (min)Degradation (%)Ref.
Bisphenol-A (BPA)50ZnFe2O4/leaf extract of Azarachita indica-sunlight72092[142]
Bisphenol-A (BPA)50CoFe2O4/leaf extract of Azarachita indica-sunlight72089[142]
Bisphenol-A (BPA)50Fe2O3/leaf extract of Azarachita indica-sunlight72070[142]
Bisphenol-A (BPA)50ZnO/leaf extract of Azarachita indica-sunlight72068[142]
Bisphenol-A (BPA)50Co3O4/leaf extract of Azarachita indica-sunlight72054[142]
Para-aminobenzoic acid20TiO2 P25-UV (9 W)120>80[143]
toluidine blue o, safranin o, falcon carboxylic acid, Hexavalent chromium Cr5ZnO-UV light (250 W)190, 310, 260, 30094, 87, 92, 68[144]
Hexavalent chromium Cr (VI) bisphenol A (BPA)10magnetic 3D-TiO2@HPGAsolvothermal processlow pressure mercury vapor lamps (8 W)140
240
100
90
[145]
Clofibric acid (CA)2g-C3N4-Xe-lamp (350 W)<5046.8[146]
Clofibric acid (CA)2P25-Xe-lamp (350 W)<5056.8[146]
Clofibric acid (CA)2g-C3N4/P25 (8 wt%)-Xe-lamp (350 W)<5085.4[146]
Atrazine100Cu-ZnO/g-C3N4 Z-direct scheme-UV12090[147]
Herbicide glyphosate100BiOBr/Fe3O4 nanocompositesChemical co-precipitation
method
Xe lamp (500 W)6097[148]
Diuron herbicide25g-C3N4/N-doped
CeO2 composite
-Xe lamp (1500 W)12046[149]
Gramoxone herbicide10TiO2 hollow fibers-UV lamp (6 W)480<50[150]
Chlorpyrifos2CeO2/TiO2/SiO2sonophotocatalyticVisible15090.8[151]
Table 7. Review on Photocatalytic Membrane degradation of pharmaceutical compounds.
Table 7. Review on Photocatalytic Membrane degradation of pharmaceutical compounds.
PollutantPollutant Concentration (mg/L)Photocatalyst/Synthesis MethodMembrane/Pore Size (µm)Light SourceTime (min)Degradation (%)Ref.
Chloramphenicol50TiO2 doped hydroxyapatite/hydrothermalpolysulfone (PSF)/0.003UV Light12061.59[160]
Pharmaceutical industry
wastewater
TDS 4740
COD 17360
Cerium fluoride (CeF)/Phase inversion by immersion precipitation techniquepolysulfone (PSF)24 W UV lamp 97[161]
Triclosan10bismuth vanadate/polyethyleneimine blended in polyethersulphone 300 W Xenon
Lamp
86[162]
Sulfamethoxazole10mesoporous graphitic carbon nitride/TiO2 300 W ozone
free xenon lamp
49[163]
Sulfadiazine12Goethite (α-FeOOH)/precipitation0.14UVL214W lamp 70 (no H2O2) and 99 (with H2O2)[164]
Diclofenac10N, S-CQDs/TiO2polysulfone (PSF)light (40 W), visible light (12 W)15062.3[165]
Oxytetracycline20TiO2UF membraneUVC lamp (16 W) 52 DOC removal[166]
Diphenhydramine-graphene oxide-TiO2Mixed cellulose ester (MCE)UV/Vis and visible light irradiation 73[167]
Chlorhexidine Digluconate-TiO2PES and PVC-PAN1500 W solar simulator 40[168]
Diclofenac2Ti O 2 PVDF/0.044 × 24 W
black Light
99.5[169]
Diclofenac and ibuprofenDiclofenac sodium salt 25, ibuprofen sodium salt 100TiO2PES and PVDF/0.227.6 mW/cm2 UVA lamp120For PES 68,
For PVDF 55
[170]
Diclofenac (DCF), ibuprofen (IBU) and naproxen (NAP)100 μg d m 3 TiO2Polypropylene (PP)/0.216 W UVC germicidal lamp DCF100, IBU 73,
NAP 90
[171]
furosemide,
ranitidine, ofloxacine, phenazone,
naproxen, carbamazepine
10TiO2different NF membranes (PES, PSF, PAN)125 W medium pressure Hg lamp120furosemide 80,
ranitidine 50,
Naproxen 90
[172]
tetracycline5Au/Ag/Ti O 2 CA/0.213Xe lamp 90[173]
Diclofenac0.05TiO2PVDF/0.0352 W UV-C power 100[174]
Tetracycline10Au-TiO2PDA-PVDF/0.22Xenon lamp12090[175]
Diclofenac50 for batch
20 for continuous
N-doped TiO2MF membrane 0.1, RO membrane 0.0001–0.0015 × visible 250 W lamps150–18097.66[84]
Ciprofloxacin10sulfonated graphene oxide/ZnOPES/0.011150 W UV24095.1[176]
Pharmaceutical industry
wastewater
TDS 4740
COD 17360
ZrO2
SnO2/sol-gel
PS24 W UV lamp 90[177]
Tetracycline10ZnI n 2 S 4 PVDF/0.3150 W
halogen tungsten lamp
50[178]
Carbamazepine1α-A l 2 O 3 coated with N-doped Ti O 2 /Sol-gel0.2–0.8300 W ozone-free xenon arc lamp 90[179]
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

Binazadeh, M.; Rasouli, J.; Sabbaghi, S.; Mousavi, S.M.; Hashemi, S.A.; Lai, C.W. An Overview of Photocatalytic Membrane Degradation Development. Materials 2023, 16, 3526. https://doi.org/10.3390/ma16093526

AMA Style

Binazadeh M, Rasouli J, Sabbaghi S, Mousavi SM, Hashemi SA, Lai CW. An Overview of Photocatalytic Membrane Degradation Development. Materials. 2023; 16(9):3526. https://doi.org/10.3390/ma16093526

Chicago/Turabian Style

Binazadeh, Mojtaba, Jamal Rasouli, Samad Sabbaghi, Seyyed Mojtaba Mousavi, Seyyed Alireza Hashemi, and Chin Wei Lai. 2023. "An Overview of Photocatalytic Membrane Degradation Development" Materials 16, no. 9: 3526. https://doi.org/10.3390/ma16093526

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