This study developed an antifouling coating for polyethersulfone (PES) membranes by tuning the bandgap of TiO2 with Cu nanoparticles (NPs) via a polyacrylic acid (PAA)-plasma-grafted intermediate layer. Cu NPs were synthesized at different molar ratios and precipitated onto TiO2 using the sol-gel method. The resulting Cu@TiO2 photocatalysts were characterized using various techniques, showing reduced bandgap, particle size range of 100–200 nm, and generation of reactive free radicals under light irradiation. The 25% Cu@TiO2 photocatalyst displayed the highest catalytic efficiency for Acid Blue 260 (AB260) degradation, achieving 73% and 96% with and without H2O2, respectively. Photocatalytic membranes based on this catalyst achieved an AB260 degradation efficiency of 91% and remained stable over five cycles. Additionally, sodium alginate-fouled photocatalytic membranes fully recovered water permeability after undergoing photocatalytic degradation of foulants. The modified membrane displayed a higher surface roughness due to the presence of photocatalyst particles. This study demonstrates the potential application of Cu@TiO2/PAA/PES photocatalytic membranes for mitigating membrane fouling in practice.

  • Bandgap tuning of TiO2 with Cu NPs successfully enhanced photocatalytic performance.

  • Mechanism of photocatalytic decomposition of Acid Blue 260 attributed to •OH radicals.

  • PAA plasma-grafting improved binding between PES membrane surface and photocatalyst.

  • Cu@TiO2/PAA/PES membranes exhibited high water flux and FRR of 98%.

Membrane filtration has been studied and applied in practice for a long time in the filtration and separation of water and wastewater (Sójka-Ledakowicz et al. 1998). Membranes are used in one or more stages of the water treatment process and have proven to be effective in terms of economy, performance, environment, etc. (Kamali et al. 2019). However, an unavoidable limitation of membrane filtration is the phenomenon of membrane fouling during the long-term operation of membranes. This reduces the performance of membranes and forces operators to clean fouled membranes frequently or even replace them with new ones (Shi et al. 2014). Measures such as controlling the membrane operating mode or incorporating membrane pretreatment methods make significant changes to the water treatment system, requiring more space and cost, but have proved inefficient (Le-Clech et al. 2006). Another approach that seems reasonable and is being widely researched is changing the nature of the membrane surface. According to studies on the mechanism of membrane fouling, this phenomenon occurs due to the strong interaction of the similar hydrophobic nature of the hydrophobic foulants and hydrophobic membrane surfaces (Horseman et al. 2020). The organic membranes used in membrane filtration are mostly hydrophobic, so converting these surface properties to hydrophilic will increase their water permeability and decrease membrane fouling. In general, inorganic modifiers proved to be more effective in increasing the hydrophilicity of membrane surfaces than organic modifiers (Rana & Matsuura 2010). This is why inorganic catalysts are most often used to modify membrane surfaces. Therefore, catalytic membranes in general and photocatalytic membranes in particular are considered superior in mitigating membrane fouling because, in addition to enhancing the hydrophilicity of the membrane surface, it also provides catalytic activity to degrade the foulants under certain conditions, such as proper light irradiation (Yang et al. 2016).

The antifouling ability of photocatalytic membranes depends mainly on the photocatalytic performance and hydrophilicity of the photocatalyst applied (Wang et al. 2022). There have been many different photocatalysts incorporated into membranes, but in general, they are catalysts based on semiconductors such as TiO2, ZnO, g-C3N4, etc. (Acharya & Parida 2020). The TiO2 and ZnO catalysts themselves are UV-activated catalysts and were used in the initial studies. Their photocatalytic activity has been enhanced by doping or compositing stable metals or other semiconductors (Pelaez et al. 2012). However, the major drawback of ZnO is its sensitivity to large changes in the pH of the medium because it is an amphoteric oxide. g-C3N4 is a visible light-activated photocatalyst and has a particularly high photocatalytic performance when combined with several suitable semiconductors (Low et al. 2021). However, in terms of hydrophilicity, g-C3N4 does not have the desired effect because the composition is non-polar nitrogen carbons (Song et al. 2019). TiO2 proved to be superior overall when considering factors such as chemical stability, environmental friendliness, photocatalytic performance, hydrophilicity, commercial availability in large quantities, and so on (Dong et al. 2015). The antifouling activity of TiO2-based photocatalytic membranes will be significantly improved if the disadvantage of the high bandgap of TiO2 is overcome. There have been many published bandgap tuning of TiO2 but generally doping it by metallic or non-metallic elements or compositing it with other low bandgap semiconductors (Eddy et al. 2023). The doping of TiO2 with metal nanoparticles (NPs) shows the advantage of being easily performed by chemical reduction reactions, and the bandgap of the resulting catalyst is also significantly reduced (Moma & Baloyi 2019). The metals commonly used in published works are durable metals, such as platinum, gold, or silver. Although the efficiency of photocatalysts is high, from an economic point of view, they do not seem feasible (Ijaz & Zafar 2021). Another fairly durable metal that has received little attention is copper. In terms of the economy and commercial availability in large quantities, copper is a tough candidate to beat (Ma et al. 2014). Therefore, this is the first time in this study that copper metal is used to tune the bandgap of TiO2 for application in Cu@TiO2-based photocatalytic antifouling PES membranes.

Membrane fouling is caused by organic compounds in wastewater that accumulate on the membrane surface over a long period of operation (Wang & Li 2008). Depending on the source of the treated wastewater, these organic compounds have very different compositions. Therefore, using actual wastewater to assess the antifouling ability of membranes makes it difficult to accurately conclude the performance of membranes (Zhang 2022). Laboratory studies often use simulated foulants because of the ease of concentration control and analysis and the exclusion of undesirable byproducts from actual wastewater (Drews 2010). Because the primary antifouling mechanism of photocatalytic membranes is to catalyze the degradation of foulants, these model foulants must be carefully selected for ease of analysis and not too readily degraded. Photocatalyst studies often use methylene blue as the model pollutant; however, this dye is readily degraded and therefore unsuitable for model foulants (Khan et al. 2022). In this study, the azo dye was chosen because it is considered to be a persistent compound and is difficult to degrade. In addition, we found that the antifouling efficiency of the modified membrane was significantly enhanced by incorporating H2O2 activation into the photocatalysis. A very small amount of H2O2 helps to achieve equilibrium of reactive free radical-generating reactions and, thus, a faster and more efficient breakdown of foulants.

In addition to the performance of the photocatalyst itself, the performance of the photocatalytic membranes is highly dependent on how the photocatalyst is incorporated into the membranes (Parrino et al. 2018). The photocatalysts in photocatalytic membranes must ensure maximum exposure to pollutants and be persistently immobilized on the membranes (Subramaniam et al. 2022). In terms of the second aspect, the photocatalytic membrane fabricated by the blending method has outstanding advantages. However, the photocatalysts located deep inside the membrane cannot have photocatalytic activity and greatly affect the permeability of the membrane, while the number of photocatalysts on the membrane surface is not enough to work efficiently (Zhang et al. 2016). This is because coating methods have been studied more extensively in the fabrication of photocatalytic membranes. In this case, the photocatalysts do not affect the porous pore structure of the membrane but are abundantly concentrated on the membrane surface. The high concentration of the photocatalyst on the membrane surface ensures efficient light absorption and contact with the pollutant (Zango et al. 2023). The problem is how to make the photocatalyst stably bound to the membrane surface. This is a key issue in the field of photocatalytic membranes that has attracted considerable attention from researchers (Gnanasekaran et al.). Until now, the solution to this problem has been to introduce a cohesive intermediate layer between the photocatalyst and the membrane surface. This intermediate layer is usually polar polymers that can form strong bonds to the membrane surface and immobilize inorganic photocatalysts through polar functional groups such as hydroxyl, ammonium, sulphonate, or carboxylate (Barclay et al. 2017). Because these polymers are polar and membranes are generally hydrophobic, a strong chemical bond is required between these polymers and the membrane surface (Van der Bruggen 2009). To the best of our knowledge, a lack of investigation on photocatalytic antifouling Cu@TiO2/polyacrylic acid (PAA)/PES membranes has been performed previously. Therefore, the present study incorporates the use of oxygen plasma to facilitate covalent chemical bonding between the PAA and the PES membrane. Characteristic and photocatalysts of the prepared membranes were determined using Fourier-transform infrared (FTIR), scanning electron microscopy (SEM), transmission electron microscopy–energy-dispersive X-ray spectroscopy (TEM–EDS), ultraviolet–visible diffuse reflectance spectroscopy (UV)–Vis DRS, X-ray diffraction (XRD), X-ray photoelectron spectroscopy (XPS), and electron spin resonance spectroscopy (ESR). The catalytic performance of photocatalyst membrane was evaluated by the removal of Acid Blue 260 (AB260) from the aqueous solution.

Materials

Tisch Scientific Company (Ohio, USA) provided polyethersulfone (PES) membranes with a pore size of 0.22 μm and a water contact angle (WCA) of 0°. AB260 (C31H27N3O6SNa) was obtained from Alfa-Chemistry (New York, USA). Titanium dioxide (TiO2, 99%), copper sulfate pentahydrate (CuSO4.5H2O, 98%), and hydrogen peroxide (H2O2, > 30%) were supplied by Merck (Germany). L-ascorbic acid (98.5%), acrylic acid (CH2 = CHCOOH, 99%), Pluronic® F-127 ((C3H6O·C2H4O)x, non-ionic surfactant), KI (potassium iodide), IPA (isopropanol), HCOOH (formic acid), and NaN3 (sodium azide) and other chemicals were provided by Sigma Aldrich (Missouri, USA). The chemicals were used as received without further purification. The solutions and standards were diluted using deionized water.

Methods

Synthesis of Cu@TiO2 photocatalyst

The Cu@TiO2 photocatalyst were prepared as follows: First, 2 g of TiO2 NPs were added to a 500 mL three-necked flask containing 350 mL of deionised water (DI) water and fitted with a thermometer, condenser, and magnetic bar, and the flask was placed on a magnetic stirring heater. This mixture in a flask was vigorously stirred and subjected to ultrasonic treatment for 1 h to achieve good dispersion of TiO2 NPs in water. Next, a certain amount of CuSO4·5H2O (molar ratios of Cu to total moles of Cu and Ti of 1, 3, 5, 10, 15, and 25%) was transferred into the vigorously stirred reaction system. One gram (1 g) of Pluronic® F-127 was added to the reactor, and the temperature of the system was slowly raised to 80 °C. After the solids in the reaction system were dissolved, 100 mL of the ascorbic acid solution was slowly added to the mixture, while the system was vigorously stirred at 80 °C for 2 h. At the end of the reaction, the color of the mixture changed to the characteristic color of the Cu NPs dispersed in water. The solid product in the reaction was allowed to cool naturally and then vacuum-filtered. The solid product was washed several times with DI water and ethanol before being dried at 70 °C for 24 h. The resulting photocatalysts are denoted as x% Cu@TiO2, where x% is the molar ratio of the Cu mentioned above. The composite photocatalyst was then characterized by FTIR, SEM, TEM, EDS, UV–Vis DRS, XRD, XPS, and ESR.

Cu@TiO2/PES membrane fabrication

The Cu@TiO2/PES fabrication process undergoes three stages (Figure 1): (1) oxygen plasma treatment of the PES membrane surface, (2) PAA grafting onto the plasma-activated PES membrane, and (3) dip-coating the PAA/PES membrane into the suspension of Cu@TiO2 in water. The oxygen plasma treatment of the PES membrane was performed in a reaction chamber of 13.56 Mhz RF (radio frequency) plasma for 5 min with plasma operating parameters, such as plasma power of 100 W, oxygen flow rate of 20 sccm, and plasma pressure of 0.060 Torr. After plasma treatment, the PES membrane was further aged in an oxygen atmosphere with an oxygen flow rate of 50 sccm for 10 min. The plasma-activated PES membrane was then immediately transferred to 20% (v/v) acrylic acid (AA) solution at 80 °C for 2 h. The PAA/PES membrane was washed several times with DI water and allowed to dry naturally. A suspension of Cu@TiO2 in water was prepared by vigorous stirring and ultrasonic treatment of the mixture with a suitable composition of Cu@TiO2 photocatalyst and water. The process of dip-coating the PAA/PES membrane into a suspension of Cu@TiO2 and drying was repeated for several cycles before the Cu@TiO2/PAA/PES membrane was left to dry naturally in the air at room conditions.
Figure 1

Schematic illustrating the fabrication process of Cu@TiO2/PES membrane.

Figure 1

Schematic illustrating the fabrication process of Cu@TiO2/PES membrane.

Close modal
The PAA content on PES membrane surfaces (degree of PAA grafting) can be calculated by the following equation:
formula
(1)
where m0 is the mass of the pristine PES membrane; mg is the mass of the PAA-grafted PES membrane.

Photocatalysis of AB260

The photocatalyzed degradation of AB260 by Cu@TiO2 powder (20 mg) was carried out in 150 mL of AB260 solution with concentrations of 10 mg/L (without H2O2) and 20 mg/L (with H2O2). This dye degradation catalyzed by Cu@TiO2/PES membrane was carried out in 150 mL of AB260 solution at a concentration of 20 mg/L in the presence of hydrogen peroxide. The Cu@TiO2 powder was transferred to AB260 solution in a 250 mL beaker placed on a magnetic stirrer at 200 rpm. The light source is two ultraviolet-C (UVC) lamps with a power of 8 W, each placed 10 cm from the solution surface. All photocatalysis was performed at room temperature and atmospheric pressure. At each predetermined time, 5 mL of the reaction mixture was withdrawn, centrifuged at 2,500 rpm for 10 min to remove solids, and analyzed by UV–Vis spectroscopy at a characteristic wavelength of AB260. When using H2O2 (30%) in the photocatalytic decomposition of AB260, a very small amount (100 μL) of it was used. While in membrane's trapping tests approximately 0.0015 mole of scavengers, i.e. KI, IPA, HCOOH, and NaN3 were employed to evaluate the contributions of holes, hydroxyl radicals, free electrons, and superoxide anion radicals to the photocatalysis. The photocatalyst with the highest efficiency in catalyzing the degradation of AB260 was selected to fabricate the photocatalytic membrane. The degradation of AB260 catalyzed by the photocatalytic membrane was carried out in the same way as above but with the help of a membrane holder. The membrane holder helps keep the membrane in place and ensures light exposure of the membrane when the AB260 solution is stirred.

Membrane permeability and antifouling

The water permeabilities of pristine PES and modified membranes were evaluated based on a vacuum filtration system with a membrane active area of 9.08 cm2. This vacuum filtration system was equipped with a pressure gauge and can precisely control the speed of the vacuum pump. To easily compare membrane performance, the speed of the vacuum pump was kept constant and consistent across all membrane filtrations. The permeates were analyzed at certain times to calculate the permeate flux parameter using the following equation:
formula
(2)
where J is permeate flux (L·m−2·h−1), V is the volume of permeate liquid (L); A is the membrane's active area (m2), Δt is the filtration time (h).
For the experiment to evaluate the antifouling ability of the photocatalytic membrane, the membrane was fouled by filtering a solution of sodium alginate at a concentration of 0.5 g/L until the system pressure increased markedly compared to the case of water filtration. Then, the fouled membrane was transferred to the photocatalytic reaction system as described in Section 2.2.3 and performed similarly. The flux recovery ratio (FRR) was calculated using the following equation:
formula
(3)
where Jx is the water flux of the fouled photocatalytic membrane after being treated by photocatalysis, Jo is the the initial water flux of the photocatalytic membrane.

Characterization of Cu@TiO2 photocatalyst

Fourier-transform infrared spectroscopy

The functional groups and specific bonds of the Cu@TiO2 photocatalysts were investigated by FTIR spectroscopy, as shown in Figure 2. Because this photocatalyst is hydrophilic, adsorption of water vapor on its surface is unavoidable. Specifically, the characteristic wave number at 3,440 cm−1 is the stretching vibration of the –O–H functional group, while the wave numbers at 1,635 and 1,380 cm−1 are the distinct bending vibrations of this hydroxyl functional group (Liu et al. 2019). Interestingly, the appearance of a peak at a wave number of 1,058 cm−1 is the characteristic vibration of metal–metal bonds in the photocatalyst, which could be Cu–Ti bonds (Sharma et al. 2017). In addition, two high-intensity peaks in the range of wave numbers 400–800 cm−1 are characteristic vibrations of Ti–O bonding in TiO2 (Lal et al. 2021). We can see that when the content of TiO2 in the photocatalyst decreases, the intensity of these two peaks decreases significantly.
Figure 2

FTIR spectra of Cu@TiO2 composite photocatalysts.

Figure 2

FTIR spectra of Cu@TiO2 composite photocatalysts.

Close modal

SEM and TEM

For solid materials, properties such as surface morphology, nanoparticle size, and particle size distribution are very important and determine the performance of the material. To better understand this aspect of the photocatalyst, methods such as SEM and TEM–EDS were applied in this study. Figure 3 shows the SEM images of the photocatalyst at different magnifications. The NPs of the photocatalyst had a narrow particle size distribution, which is clearly shown in Figure 3(a). The sizes of the particles range from 100 to 200 nm, and the NPs have the shape of TiO2 crystals. The TEM images of the photocatalyst (Figure 4(a)–4(c)) also clearly show the morphology and size of the photocatalyst particles. There was agreement on the particle size and morphology of the photocatalysts when analyzing the SEM and TEM images. In particular, the dark spots on the NPs in the TEM images could be layers of Cu NPs that obscure electron transmission (Xu et al. 2015). The presence of these Cu NPs is more clearly demonstrated by EDS analysis, as shown in Figure 4(d). It can be seen that only elements such as oxygen, titanium, and copper are present in this composite photocatalyst. The signal intensity of the Cu element is low because it is incorporated only in small amounts into the TiO2 photocatalyst.
Figure 3

SEM images of 10% Cu@TiO2 composite photocatalyst.

Figure 3

SEM images of 10% Cu@TiO2 composite photocatalyst.

Close modal
Figure 4

TEM images of 10% Cu@TiO2 composite photocatalyst (a–c) and its EDS survey (d).

Figure 4

TEM images of 10% Cu@TiO2 composite photocatalyst (a–c) and its EDS survey (d).

Close modal

Ultraviolet–visible diffuse reflectance spectroscopy

The purpose of incorporating Cu NPs into the TiO2 photocatalyst was to reduce the bandgap of TiO2. To assess the success of this process, it is necessary to use a method such as DRS to accurately determine the bandgap values of the photocatalysts before and after the combination. Figure 5(a) shows that the Cu@TiO2 photocatalyst has a light-reflective edge at a slightly longer wavelength than TiO2. However, because TiO2 is white and Cu@TiO2 is dark, TiO2 reflects light more strongly than Cu@TiO2 (Cheng et al. 2021). Calculating the bandgap from the DRS spectrum requires unit conversions and calculations to represent the Tauc plot (Figure 5(b)). It can be easily seen that at a molar ratio of Cu of only 10%, the bandgap of the modified photocatalyst is significantly reduced compared to that of the pristine TiO2. Specifically, the bandgap of pristine TiO2 is 3.55 eV, while the bandgap of Cu@TiO2 is reduced to 3.32 eV. Theoretically, the bandgap of the modified photocatalyst will decrease even more when increasing the Cu nanoparticle content in the photocatalyst, such as 15 and 25%. This increased the catalytic activity of the resulting composite photocatalyst.
Figure 5

The UV–Vis DRS of TiO2 and Cu@TiO2 (a), and their Tauc plots (b).

Figure 5

The UV–Vis DRS of TiO2 and Cu@TiO2 (a), and their Tauc plots (b).

Close modal

X-ray diffraction

From a materials science perspective, in many cases, the same constituent elements but the bond types and crystal forms are sometimes very different. Material performance is sometimes limited to a few types of crystals or bonds. Therefore, to determine the specific structure of photocatalysts, we used the XRD method, as shown in Figure 6. When the content of Cu NPs is very low (1%), the XRD pattern of the photocatalysts can be considered as that of TiO2 because this method cannot detect such small concentrations. Precisely, the characteristic peaks of TiO2 anatase at the diffraction angles of 25.2°, 37.7°, 48.0°, 53.9°, 55.0°, and 62.7° correspond to the lattice planes of (101), (004), (200), (105), (211), and (204), respectively (Li et al. 2020). However, when the content of Cu NPs was at 25%, the characteristic diffraction peaks of 25%Cu@TiO2 appeared three more peaks of Cu NPs at the diffraction angles of 43.3°, 50.4°, and 74.2° corresponding to the lattice planes of (111), (200), and (220), respectively (Mohl et al. 2011). The above proves that the crystal structure of Cu@TiO2 includes the crystal structure of TiO2 and Cu NPs.
Figure 6

XRD patterns of Cu@TiO2 and Cu NPs.

Figure 6

XRD patterns of Cu@TiO2 and Cu NPs.

Close modal

X-ray photoelectron spectroscopy

The chemical states and bonding types of the elements in the composite catalyst were further analyzed by the XPS method, as shown in Figure 7. The XPS survey (Figure 7(a)) indicates that the elemental composition of the photocatalyst material consists of oxygen (O 1s), copper (Cu 2p), and titanium (Ti 2p). This analysis only scans a thin layer on the material surface, so the atomic composition in Figure 7(a) indicates a high Cu content. This phenomenon may be attributed to the fact that Cu NPs were precipitated onto TiO2 rather than forming simultaneously with TiO2. The chemical state of oxygen (O 1s) (Figure 7(b)) shows that the obtained peak is the sum of the three-component peaks. These three-component peaks at binding energies of 529.2, 530.4, and 531.8 eV correspond to the state of oxygen in the lattice of Ti–O, the state of oxygen vacancy in the lattice, and the state of oxygen in water and CO2 molecules adsorbed on the material surface, respectively (Idriss 2021). Because the intensity of the oxygen vacancy is much higher than that of other states, it is essential for the high catalytic activity of this material. The state of copper (Cu 2p3/2) in the material (Figure 7(c)) is also shown to include two component peaks at binding energies of 932.2 and 934.0 eV. The peak with low binding energy and high intensity indicates that copper's state belongs to Cu0 (Li et al. 2021). Meanwhile, the peak with high binding energy and low intensity belongs to the state of Cu+. The Cu+ can be considered to be distributed in the corners, faces, and edges of Cu NPs. Similarly, the state of titanium (Ti 2p3/2) in the material (Figure 7(d)) is also analyzed into three-component peaks at binding energies of 457.8, 458.9, and 459.4 eV. These three peaks correspond to the titanium states of Ti2+, Ti3+, and Ti4+ in the material, respectively (Zhang et al. 1997). Notably, the peak of the Ti3+ state is the most intense, and according to the studies, this is the state that makes the TiO2-based catalyst highly active.
Figure 7

10%Cu@TiO2's XPS survey (a) and deconvoluted XPS of O 1s (b), Cu 2p (c), and Ti 2p(d).

Figure 7

10%Cu@TiO2's XPS survey (a) and deconvoluted XPS of O 1s (b), Cu 2p (c), and Ti 2p(d).

Close modal

ESR spectroscopy

Photocatalysis in the presence of photocatalysts is the generation of reactive free radicals in an aqueous medium, such as OH and . Therefore, the performance of the photocatalyst can be indirectly based on the presence of these active free radicals in the reaction solution. The ESR method was used to detect the presence of OH and free radicals in the solution applied in this study (Figure 8). In this method, the Cu@TiO2 photocatalyst was dispersed in an aqueous or methanol solvent in the presence of DMPO (5,5-Dimethyl-1-Pyrroline-N-Oxide). DMPOs are responsible for keeping these free radicals more stable and not easily recombined to form neutral molecules. Figure 8(a) shows that only in the case of UV light irradiation, the signal of OH is high and clear; in the remaining cases, there is almost no appearance of this free radical. The same conclusion was found for the case of the free radical (Figure 8(b)); that is, in the absence of UV light irradiation, the signal of this free radical was almost negligible. From these results, it can be seen that the Cu@TiO2 photocatalyst triggered the generation of active free radicals required for pollutant degradation under UV irradiation.
Figure 8

ESR spectra of OH/H2O (a) and /methanol (b) with the presence of Cu@TiO2.

Figure 8

ESR spectra of OH/H2O (a) and /methanol (b) with the presence of Cu@TiO2.

Close modal

Photocatalytic decomposition of AB260

The reason the photocatalyst can catalyze the decomposition of pollutants under the irradiation of suitable light is due to the transfer of electrons from the valence band to the conduction band to generate active electrons and holes. The free electrons in the conduction band are strong reducing agents, and the positively charged holes in the valence band are strong oxidizing agents. Because TiO2 has a high bandgap, the generation of free electrons and holes is more difficult and requires higher energy light than semiconductors with a lower bandgap. This is why Cu NPs were incorporated into TiO2 in this study to reduce the bandgap of the photocatalyst (Figure 9). Cu NPs introduce an intermediate energy level lower than that of the conduction band of TiO2, so the free electrons from the conduction band of TiO2 will shift to this energy level to achieve a more stable state and reduce recombination with the hole. The free electrons in the conduction band can reduce oxygen in solution to form reactive free radicals ; while holes in the valence band can oxidize water, OH, and H2O2 to form the active free radical OH (Raja et al. 2016). This is because the reduction potentials of these reactions are in the middle of the valence and conduction band energies in the composite photocatalyst. The photocatalytic mechanism of Cu@TiO2 composite photocatalyst is proposed in Equations (4)–(13). Accordingly, Equations (5), (6), (10), and (12) produce active free radicals OH, while Equation (6) produces active free radicals .
Figure 9

Illustration of the charge transfer mechanism in Cu@TiO2 photocatalysts.

Figure 9

Illustration of the charge transfer mechanism in Cu@TiO2 photocatalysts.

Close modal
formula
(4)
formula
(5)
formula
(6)
formula
(7)
formula
(8)
formula
(9)
formula
(10)
formula
(11)
formula
(12)
formula
(13)
The catalytic activity of Cu@TiO2 photocatalyst for the AB260 degradation reaction was investigated with and without using a small amount of H2O2 (Figure 10). The negligible AB260 adsorption of the photocatalysts in this study reached equilibrium after about 5 min. To simplify the illustration of the dye degradation efficiency, this adsorption was omitted in Figure 10. In the absence of H2O2, the degradation efficiency of AB260 increased markedly when the photocatalyst containing Cu NPs was compared with that of pristine TiO2 (more than 20% increase). In addition, Figure 10(a) shows that when increasing the content of Cu NPs in the photocatalyst, the degradation efficiency of AB260 also increases and reaches the highest at a molar ratio of 25% Cu@TiO2. Specifically, the degradation efficiency of AB260 when using photocatalysts at molar ratios of 0, 1, 3, 5, 10, 15, and 25% was 39, 56, 62, 67, 69, 71, and 73%, respectively. When using H2O2, the AB260 decomposition efficiency was significantly enhanced, even without the use of a photocatalyst (53%). This can be explained based on Equation (12) for the activation of H2O2 under UV light. Therefore, when using photocatalyst and H2O2 at the same time, the highest dye degradation efficiency was 96%, corresponding to the molar ratio of 25% Cu@TiO2 (Figure 10(b)). These degradation efficiencies were only markedly different in the presence and absence of Cu NPs in the photocatalyst, but when Cu NPs are present, this difference is not too obvious. Specifically, the dye degradation efficiency for molar ratios of 0, 1, 3, 5, 10, 15, and 25% is 80, 91, 93, 93, 94, and 96%, respectively.
Figure 10

Photocatalytic decompositions of AB260 catalyzed by Cu@TiO2 without H2O2 (a) and with H2O2 (b).

Figure 10

Photocatalytic decompositions of AB260 catalyzed by Cu@TiO2 without H2O2 (a) and with H2O2 (b).

Close modal
A membrane trapping test was performed (Figure 11(a)) to better understand the main mechanism of the photocatalytic membranes in this study. The scavenging effects of KI, IPA, HCOOH, and NaN3 to active agents, i.e. h+, OH, e, and O2, respectively were evaluated in experiments.
Figure 11

Membrane's trapping tests (a), dye decomposition catalyzed by photocatalyst and membrane (b), and membrane's stability cycle test (c,d).

Figure 11

Membrane's trapping tests (a), dye decomposition catalyzed by photocatalyst and membrane (b), and membrane's stability cycle test (c,d).

Close modal

The scavengers will capture these active agents and make the degradation reaction caused by them virtually nonexistent, which greatly reduces the degradation reaction efficiency. Figure 11(a) shows the dye degradation efficiency of 30, 60, 68, and 85% using IPA, NaN3, HCOOH, and KI scavengers, respectively. Therefore, it can be concluded that the main mechanism causing the dye degradation catalyzed by this photocatalytic membrane is related to the free radical OH. When comparing the performance of the photocatalyst and the photocatalytic membrane (Figure 11(b)), it can be seen that the degradation reaction rate at the initial time when using the photocatalyst is higher than when using the photocatalytic membrane. However, at the later stage of the reaction, this difference in performance was not very significant (about 7%). In addition, when the photocatalytic membrane is used with and without H2O2, the high difference in dye degradation efficiency is about 35%. Therefore, it proves the role of H2O2 activation in the antifouling activity of this photocatalytic membrane. The stability of the antifouling performance of the photocatalytic membrane was also evaluated through cycle tests (Figure 11(c) and 11(d)). The dye degradation efficiency decreased slightly with each cycle but was not significant and can be considered stable. In detail, the dye degradation efficiency over five consecutive cycles was 91, 89, 86, 82, and 82%, respectively; i.e. this efficiency decreases by 9% after five photocatalysis cycles.

Membrane's permeability and antifouling activity

One of the important parameters for membranes applied in wastewater filtration is the membrane's water permeability. Therefore, it is necessary to evaluate the effect of the modification process on the water permeability of the photocatalytic membrane. The PES membrane used in this study is hydrophilic, with a WCA of 0°, so its water permeability was quite high around 2,013 L·m−2·h−1. In addition, the pure water flux of the pristine PES membrane was very stable for 300 min of water filtration. Because of the influence of a photocatalytic coating and an intermediate PAA layer on the membrane surface, the pure water flux of the photocatalytic membrane was lower than that of the pristine PES membrane (1,868 L·m−2·h−1). However, when compared with other hydrophobic membranes of the same parameter, such as polyvinylidene fluoride (PVDF), polytetrafluoroethylene (PTFE), or polysulfone (PSf), the water permeability of this photocatalytic membrane is still very high and acceptable. Specifically, after 300 min of water filtration, the permeability of the photocatalytic membrane was only about 10% lower than that of the pristine PES membrane and was also relatively stable. Figure 12(b) shows that when the photocatalytic membrane was fouled by sodium alginate, the water permeability of the fouled membrane was only about 220 L·m−2·h−1, which was about a 90% reduction compared to the original membrane. However, after being treated by photocatalysis, the water permeability of the fouled photocatalytic membrane recovered to the value of 1,832 L·m−2·h−1. Thus, the FRR of the photocatalytic membrane, in this case, was about 98%. Therefore, it proves that the modified membrane in this study has a high antifouling ability under the investigated conditions.
Figure 12

The pure water fluxes of pristine and photocatalytic membranes (a) and antifouling performance of photocatalytic membrane (b).

Figure 12

The pure water fluxes of pristine and photocatalytic membranes (a) and antifouling performance of photocatalytic membrane (b).

Close modal
Atomic force microscopy (AFM) analysis was used to understand the surface roughness of the pristine PES membrane and the photocatalyst-coated membrane in this study (Figure 13). Figure 13(a) and 13(b) show the surface of a pristine PES membrane with pore sizes of about 1.5 μm. However, because the density of these pores is not too high, the overall surface roughness Ra parameter is only 71.3 nm. For coated membranes, it is easy to see the morphology of the photocatalyst NPs on the membrane surface. Because this photocatalyst layer is quite thick, the surface roughness parameter Ra is quite high at about 180 nm. Although theoretically high surface roughness could somewhat reduce the antifouling activity of membranes, because these coated photocatalytic membranes can strongly catalyze the degradation of foulants, the surface roughness effect is not significant in this case.
Figure 13

The membrane surface roughness analysis by AFM: pristine PES membrane (a,b) and photocatalytic membrane (c,d).

Figure 13

The membrane surface roughness analysis by AFM: pristine PES membrane (a,b) and photocatalytic membrane (c,d).

Close modal

In this investigation, Cu NPs were employed to adjust the high bandgap energy of TiO2, resulting in the production of a composite photocatalyst that was coated onto a PES membrane with a plasma-grafted PAA layer. The catalytic activity of both powdered photocatalysts and coated photocatalytic membranes was fully examined. Various characterizations were utilized, including FTIR, SEM, TEM–EDS, UV–Vis DRS, XRD, XPS, and ESR, to demonstrate the successful synthesis of the composite photocatalyst, which has a reduced bandgap (3.32 eV) and a small and uniform particle size (100–200 nm), capable of generating reactive free radicals under light irradiation. The highest catalytic efficiency (73%) was observed in the powdered photocatalyst at a molar ratio of 25%Cu@TiO2. The activation of H2O2 contributed greatly to the AB260 dye degradation efficiency catalyzed by the composite photocatalyst (96%). The coated photocatalytic 25%Cu@TiO2/PAA/PES membrane catalyzed the dye degradation with an efficiency of 91% and was quite stable over five continuous cycles. The water permeability of the coated photocatalytic membrane was still quite high (1,868 L·m−2·h−1), stable, and only reduced by about 9% compared to the pristine PES membrane. The FRR of the coated photocatalytic membrane after photocatalytic treatment was very high, at about 98%. Although coated photocatalytic membranes have higher surface roughness than pristine PES membranes, the antifouling ability may not be significantly affected because of the high photocatalytic activity of modified membranes. This study shows that Cu@TiO2/PAA/PES photocatalytic membrane has the potential to be researched and applied in real textile wastewater or other contaminant in industry wastewater.

This research work was partly supported by the International Atomic Energy Agency (IAEA) under Research Contract No. 24715.

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

All relevant data are included in the paper or its Supplementary Information.

The authors declare there is no conflict.

Barclay
T. G.
,
Hegab
H. M.
,
Clarke
S. R.
&
Ginic‐Markovic
M.
2017
Versatile surface modification using polydopamine and related polycatecholamines: chemistry, structure, and applications
.
Advanced Materials Interfaces
4
(
19
),
1601192
.
Eddy
D. R.
,
Permana
M. D.
,
Sakti
L. K.
,
Sheha
G. A. N.
,
Hidayat
S.
,
Takei
T.
,
Kumada
N.
&
Rahayu
I.
2023
Heterophase polymorph of TiO2 (Anatase, Rutile, Brookite, TiO2 (B)) for efficient photocatalyst: fabrication and activity
.
Nanomaterials
13
(
4
),
704
.
Gnanasekaran
L.
,
Santhamoorthy
M.
,
Naushad
M.
,
ALOthman
Z. A.
,
Soto-Moscoso
M.
,
Show
P. L.
&
Khoo
K. S.
2022
Photocatalytic removal of food colorant using NiO/CuO heterojunction nanomaterials
.
Food and Chemical Toxicology
167
,
113277
.
Horseman
T.
,
Yin
Y.
,
Christie
K. S.
,
Wang
Z.
,
Tong
T.
&
Lin
S.
2020
Wetting, scaling, and fouling in membrane distillation: state-of-the-art insights on fundamental mechanisms and mitigation strategies
.
ACS ES&T Engineering
1
(
1
),
117
140
.
Ijaz
M.
&
Zafar
M.
2021
Titanium dioxide nanostructures as efficient photocatalyst: progress, challenges and perspective
.
International Journal of Energy Research
45
(
3
),
3569
3589
.
Kamali
M.
,
Suhas
D.
,
Costa
M. E.
,
Capela
I.
&
Aminabhavi
T. M.
2019
Sustainability considerations in membrane-based technologies for industrial effluents treatment
.
Chemical Engineering Journal
368
,
474
494
.
Khan
I.
,
Saeed
K.
,
Zekker
I.
,
Zhang
B.
,
Hendi
A. H.
,
Ahmad
A.
,
Ahmad
S.
,
Zada
N.
,
Ahmad
H.
&
Shah
L. A.
2022
Review on methylene blue: its properties, uses, toxicity and photodegradation
.
Water
14
(
2
),
242
.
Le-Clech
P.
,
Chen
V.
&
Fane
T. A.
2006
Fouling in membrane bioreactors used in wastewater treatment
.
Journal of Membrane Science
284
(
1–2
),
17
53
.
Li
Z.
,
Xing
X.
,
Zhang
J.
,
Li
M.
&
Zhang
Q.
2020
Highly ordered hierarchically macroporous-mesoporous TiO2 for thiol-ene polymer design by photoclick chemistry
.
Microporous and Mesoporous Materials
291
,
109696
.
Li
G.
,
Kou
M.
,
Tu
J.
,
Luo
Y.
,
Wang
M.
&
Jiao
S.
2021
Coordination interaction boosts energy storage in rechargeable Al battery with a positive electrode material of CuSe
.
Chemical Engineering Journal
421
,
127792
.
Liu
Z.
,
Rios-Carvajal
T.
,
Ceccato
M.
&
Hassenkam
T.
2019
Nanoscale chemical mapping of oxygen functional groups on graphene oxide using atomic force microscopy-coupled infrared spectroscopy
.
Journal of Colle and Interface Science
556
,
458
465
.
Ma
Y.
,
Wang
X.
,
Jia
Y.
,
Chen
X.
,
Han
H.
&
Li
C.
2014
Titanium dioxide-based nanomaterials for photocatalytic fuel generations
.
Chemical Reviews
114
(
19
),
9987
10043
.
Mohl
M.
,
Dobo
D.
,
Kukovecz
A.
,
Konya
Z.
,
Kordas
K.
,
Wei
J.
,
Vajtai
R.
&
Ajayan
P. M.
2011
Formation of CuPd and CuPt bimetallic nanotubes by galvanic replacement reaction
.
Journal of Physical Chemistry C
115
(
19
),
9403
9409
.
Moma
J.
&
Baloyi
J.
2019
Modified titanium dioxide for photocatalytic applications
.
Photocatalysts – Applications and Attributes
18
,
10
5772
.
Parrino
F.
,
Bellardita
M.
,
García-López
E.
,
Marcì
G.
,
Loddo
V.
&
Palmisano
L.
2018
Heterogeneous photocatalysis for selective formation of high-value-added molecules: some chemical and engineering aspects
.
ACS Catalysis
8
(
12
),
11191
11225
.
Pelaez
M.
,
Nolan
N. T.
,
Pillai
S. C.
,
Seery
M. K.
,
Falaras
P.
,
Kontos
A. G.
,
Dunlop
P. S.
,
Hamilton
J. W.
,
Byrne
J. A.
&
O'Shea
K.
2012
A review on the visible light active titanium dioxide photocatalysts for environmental applications
.
Applied Catalysis B: Environmental
125
,
331
349
.
Raja
V.
,
Shiamala
L.
,
Alamelu
K.
&
Ali
B. J.
2016
A study on the free radical generation and photocatalytic yield in extended surfaces of visible light active TiO2 compounds
.
Solar Energy Materials and Solar Cells
152
,
125
132
.
Rana
D.
&
Matsuura
T.
2010
Surface modifications for antifouling membranes
.
Chemical Reviews
110
(
4
),
2448
2471
.
Shi
X.
,
Tal
G.
,
Hankins
N. P.
&
Gitis
V.
2014
Fouling and cleaning of ultrafiltration membranes: a review
.
Journal of Water Process Engineering
1
,
121
138
.
Sójka-Ledakowicz
J.
,
Koprowski
T.
,
Machnowski
W.
&
Knudsen
H. H.
1998
Membrane filtration of textile dyehouse wastewater for technological water reuse
.
Desalination
119
(
1–3
),
1
9
.
Song
B.
,
Wang
T.
,
Wang
L.
,
Liu
H.
,
Mai
X.
,
Wang
X.
,
Wang
N.
,
Huang
Y.
,
Ma
Y.
&
Lu
Y.
2019
Interfacially reinforced carbon fiber/epoxy composite laminates via in-situ synthesized graphitic carbon nitride (g-C3N4)
.
Composites Part B: Engineering
158
,
259
268
.
Subramaniam
M. N.
,
Goh
P. S.
,
Kanakaraju
D.
,
Lim
J. W.
,
Lau
W. J.
&
Ismail
A. F.
2022
Photocatalytic membranes: a new perspective for persistent organic pollutants removal
.
Environmental Science and Pollution Research
29
(
9
),
12506
12530
.
Van der Bruggen
B.
2009
Chemical modification of polyethersulfone nanofiltration membranes: a review
.
Journal of Applied Polymer Science
114
(
1
),
630
642
.
Wang
X.
,
Li
S.
,
Chen
P.
,
Li
F.
,
Hu
X.
&
Hua
T.
2022
Photocatalytic and antifouling properties of TiO2-based photocatalytic membranes
.
Materials Today Chemistry
23
,
100650
.
Xu
S.
,
Man
B.
,
Jiang
S.
,
Wang
J.
,
Wei
J.
,
Xu
S.
,
Liu
H.
,
Gao
S.
,
Liu
H.
&
Li
Z.
2015
Graphene/Cu nanoparticle hybrids fabricated by chemical vapor deposition as surface-enhanced Raman scattering substrate for label-free detection of adenosine
.
ACS Applied Materials & Interfaces
7
(
20
),
10977
10987
.
Yang
H.-C.
,
Hou
J.
,
Chen
V.
&
Xu
Z. K.
2016
Surface and interface engineering for organic–inorganic composite membranes
.
Journal of Materials Chemistry A
4
(
25
),
9716
9729
.
Zango
Z. U.
,
Khoo
K. S.
,
Garba
A.
,
Kadir
H. A.
,
Usman
F.
,
Zango
M. U.
,
Oh
W. D.
&
Lim
J. W.
2023
A review on superior advanced oxidations and photocatalytic degradation techniques for perfluorooctanoic acid (PFOA) elimination from wastewater
.
Environmental Research
221,
115326
.
Zhang
F.
,
Jin
S.
,
Mao
Y.
,
Zheng
Z.
,
Chen
Y.
&
Liu
X.
1997
Surface characterization of titanium oxide films synthesized by ion beam enhanced deposition
.
Thin Solid Films
310
(
1–2
),
29
33
.
Zhang
W.
,
Ding
L.
,
Luo
J.
,
Jaffrin
M. Y.
&
Tang
B.
2016
Membrane fouling in photocatalytic membrane reactors (PMRs) for water and wastewater treatment: a critical review
.
Chemical Engineering Journal
302
,
446
458
.
This is an Open Access article distributed under the terms of the Creative Commons Attribution Licence (CC BY-NC-ND 4.0), which permits copying and redistribution for non-commercial purposes with no derivatives, provided the original work is properly cited (http://creativecommons.org/licenses/by-nc-nd/4.0/).