Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Cellulose nanofiber-scaffolded rGO/WO3 membrane for photocatalytic degradation of methylene blue under visible light irradiation

Medhanit Tefera Yifiraabc, Gebrehiwot Gebreslassieabd, Anteneh Kindu Mersha*abe and Kebede Nigussie Mekonnen*ab
aDepartment of Industrial Chemistry, Addis Ababa Science and Technology University, P.O. Box 16417, Addis Ababa, Ethiopia. E-mail: anteneh.kindu@aastu.edu.et; kebede.nigussie@aastu.edu.et
bNanotechnology Centers of Excellence, Addis Ababa Science and Technology University, P.O. Box 16417, Addis Ababa, Ethiopia
cDepartments of Chemistry, Wolkite University, P.O. Box 07, Wolkite, Ethiopia
dSchool of Mechanical and Electrical Engineering, University of Electronic Science and Technology of China, Chengdu 611731, China
eInstitute for Nanotechnology and Water Sustainability, College of Engineering, Science and Technology, University of South Africa, Florida Science Campus, Roodepoort 1709, South Africa

Received 7th May 2025 , Accepted 18th June 2025

First published on 2nd July 2025


Abstract

Although membrane-based water treatment has advanced significantly and many commercial products are currently available, its practical application remains limited, particularly in regions with inadequate electricity infrastructure. In this study, cellulose-based photocatalytic membranes incorporating visible light-active rGO/WO3 nanoparticles into a cellulose nanofiber (CNF) scaffold were fabricated using a simple casting method. The structure–property relationships of the nanostructures and the membranes were examined, and the membranes were tested for the photocatalytic degradation of methylene blue (MB). The CNF/rGO/WO3 composite membrane demonstrated significantly higher photodegradation capacity (89.3%) compared to pure WO3 (50.13%) for MB under visible light, with consistent performance for up to ten cycles. After ten cycles of repeated use, the membrane retained a degradation efficiency of 48%, which is comparable to that of WO3. This notable performance was attributed to improved charge separation, enhanced visible light absorption, larger surface area, and synergistic interactions among the components within the composite membrane. Besides enhancing photodegradation capacity through increased charge carrier separation, CNF played a vital role in immobilizing and stabilizing the photocatalyst. The principal active species identified in the photocatalytic material include h+, ˙O2, and ˙OH, with a slight predominance of h+. The simplicity of cellulose-mediated, visible-light active functional membrane design and fabrication creates opportunities for low-cost solutions to address water and energy challenges in environments with limited infrastructure. These visible light-responsive membrane materials hold great promise for applications in areas such as food packaging, surface disinfection, and the development of self-healing materials.


Introduction

Water pollution has emerged as a significant global environmental issue.1 It has become a worldwide environmental and public health concern, as the majority of areas across the world are facing a shortage of freshwater.2 Industrial discharges, including heavy metal ions and textile dyes, are among the primary contributors to water pollution.3 Many organic pollutants, such as textile dyes, exhibit high toxicity, and their high chemical stability and poor biodegradability lead to environmental persistence. Moreover, organic dye such as methylene blue (MB) is difficult to treat with traditional physicochemical or biochemical methods. Higher concentrations of organic dyes are associated with different health effects such as discoloration of urine, vomiting, tissue necrosis, abdominal pain, hypertension, and anemia.3 This calls for the development of efficient methods for remediating organic dyes in industrial effluents.

Water treatment methods such as membrane filtration4 and the use of ion exchange membranes5 and adsorption6–8 have been studied to reclaim dye-contaminated water. However, in addition to process inefficiency, these techniques involve the transition of contaminants from their liquid to solid states, creating secondary pollutants that could potentially harm the environment and complicate process technology.9 Photocatalysis, on the other hand, is recognized as a highly effective and promising technology because it can successfully remove organic pollutants from water without creating hazardous byproducts.10 Due to its unique qualities, such as cost-effectiveness, process simplicity, and response to visible light, photocatalytic technologies are advancing rapidly in the wastewater treatment industry.9

Semiconductor materials use solar energy to drive chemical reactions, creating electron–hole pairs. This process includes the formation of highly reactive species, such as hydroxyl radicals (˙OH) and superoxide radicals (˙O2), following a sequence of oxidation and reduction events. They serve as the primary oxidizing agents for the conversion of organic dyes into green products.11 Tungsten trioxide (WO3) is widely utilized as a semiconductor photocatalyst and it provides one of the most effective means to degrade organic dyes from water because it is a versatile metal oxide with outstanding chemical stability and good film-forming and coating properties.12 However, limitations such as its absorption in only the UV region, the rapid recombination of photogenerated electron/hole (e/h+) pairs, easy agglomeration, difficulty of recycling, and high energy costs, restrict its large-scale applications.13

To improve the photocatalytic properties of WO3 and, more importantly, facilitate the recycling and reuse of the catalyst, it is essential to load WO3 onto substrates such as reduced graphene oxide (rGO), molecular sieves, and natural minerals.14 Such substrate materials not only bind WO3 particles but also serve as excellent electron mediators to adjust electron transfer and improve electron–hole separation by restraining the recombination of photogenerated charges. These substrates also enhance the visible light absorption capacity of WO3.15 For example, Dhivyaprasath et al.16 incorporated WO3 into a rGO/WO3 composite thin film with 88% MB degradation performance after 120 min irradiation. Zhu et al.15 utilized rGO/WO3 composites for the removal of 98% sulfonamide under visible light irradiation. However, although these powdered composite materials exhibit excellent performance in pollutant degradation, recovering the photocatalyst from the treated wastewater and reusing it in further treatment cycles is challenging.15,17

Immobilizing photocatalysts such as rGO/WO3 nanocomposites in filtration membranes produce photocatalytic membranes (PCMs) that allow both photodegradation and catalyst recovery. This, in turn, imparts self-cleaning properties to the membrane, further increasing the lifetime of the photocatalyst and overall treatment performance. This class of membranes also minimizes biofouling, a critical challenge in membrane separation.2 Several researchers have reported photocatalytic nanocomposite membranes for water treatment applications, such as PVDF-MoS2/WO3-PVA membrane,17 rGO/TiO2-PPSU hybrid ultrafiltration membranes,18 CNF/ZnO–Ag membrane,19 and CNF/Ag2/Se composite films.20 In the present work, a cellulose nanofiber (CNF) scaffolded rGO/WO3 nanocomposite membrane was developed and tested for photodegradation of MB dye. The CNF-networks provide a platform for uniform distribution of rGO/WO3 nanomaterials,21 resulting in high surface area, improved stability of the resulting nanocomposite, and decreased electron–hole recombination rate, making CNF an ideal carrier for rGO/WO3 nanoparticles.

Materials and methods

Materials

All chemicals and reagents used in this study were analytical-grade and used without further purification. HCl (35%), H2SO4 (98%), CH3COOH (98%), (CH3)2CO (98%), sodium hypochlorite (NaOCl, 5.0%), absolute ethanol (99.9%), and graphite powder (≥99%) were purchased from Fine Chemicals, India. NaCl (99%), NaNO3 (99%), KMnO4 (99%), H2O2 (30%), and KOH (98%) were purchased from Loba Chemicals, India. Sodium tungstate dihydrate (Na2WO4·2H2O, 99%) was obtained from Alpha Chemika, India. Toluene rectified (99.9%) was purchased from Pentokey Organy, India, and polyvinyl alcohol (PVA, 99%, MW 85,000–124,000 g mol−1) was obtained from Merck, India. Distilled water was used throughout the experiments.

Isolation and purification of CNFs from bamboo

The bamboo was dried in the open air, then the chopped cellulose fiber (10 g) was oven-dried and dewaxed by Soxhlet extraction with a 2[thin space (1/6-em)]:[thin space (1/6-em)]1 (v/v) mixture of toluene/ethanol at 70 °C for 7 h. The obtained extract was washed with distilled water and treated with NaOCl (3%) at pH 4, adjusted with glacial acetic acid (98%), while heating at 75 °C for 1 h. The bleaching process was repeated several times until the material became white, indicating the formation of holocellulose. The holocellulose was treated with 2 wt% KOH at 90 °C for 2 h to remove hemicelluloses, residual pectin, and starch. The purified cellulose powder was washed with distilled water until pH 7 and oven-dried at 105 °C to remove moisture.22 The CNF was prepared by chemical treatment of cellulose suspension with dilute H2SO4 under continuous mechanical stirring (500 rpm) at 35 °C. The obtained milky suspension was centrifuged, followed by ultrasonication (SJIA-250W, Ultrasonic Homogenizer, China) to obtain well-fibrillated nano-dimensional structures. The sonication process was performed for 30 min at an output power of 400 W. The transparent supernatant obtained after centrifugation was transferred to a glass Petri dish and oven-dried at 50 °C overnight to obtain CNF (Fig. 1).12,23
image file: d5ra03227j-f1.tif
Fig. 1 Schematic of the isolation and purification of CNF from bamboo.

Preparation of CNF/PVA blend

Varying concentrations of PVA solutions (0.5% and 2%) were slowly added to a previously prepared CNF dispersion (2%) while stirring at 80 °C for 2 h, followed by ultrasonication. The viscous CNF/PVA blend solution was either cast on a Petri dish to form a CNF/PVA reference film or used as a blended matrix for the preparation of composite membranes.

Synthesis of graphene oxide

Graphene oxide (GO) was synthesized from graphene powder using a modification of Hummer's method.24 Approximately 3 g of graphite powder and 1.5 g of NaNO3 were added to a beaker containing 75 mL of H2SO4 (98%) and placed in an ice bath with continuous stirring. Subsequently, KMnO4 (9 g) was gradually added, and the reaction was continued for 2 h. The solution was then removed from the ice bath and heated at 40 °C for 1 h, followed by the addition of 150 mL distilled water, and heating was continued for 1.5 h. The temperature was raised to 95 °C for 30 min then the mixture was diluted with 200 mL of distilled water under continuous stirring. H2O2 (30%, 30 mL) was then added to the solution while heating at 95 °C for 30 min. The mixture was then stirred for 12 h until effervescence ceased and a brilliant yellow suspension was observed. The suspension was centrifuged and the supernatant was decanted. The brown precipitates obtained were dispersed in HCl solution (10%) with vigorous stirring, followed by centrifugation. The obtained brown product was re-dispersed in distilled water, centrifuged, washed repeatedly, and dried overnight in a vacuum oven at approximately 80 °C to obtain GO.

Synthesis of rGO/WO3 nanocomposite

A one-step in situ solvothermal method was used to synthesize the rGO/WO3 nanocomposite.13,14,25 In brief, the synthesized GO (0.3 g) was dispersed in distilled water (40 mL) and kept under sonication. Then, Na2WO4·2H2O (7.35 g) and NaCl (1.4 g) were added to the GO dispersion while adjusting the pH to 3 with HCl (3 M). The solution was stirred using a magnetic stirrer, transferred to a 100 mL Teflon-lined stainless-steel autoclave, and kept in the hot air oven at 180 °C for 18 h for a hydrothermal reaction. The resultant residue was washed with ethanol and distilled water several times and dried at 60 °C in a vacuum oven. Similar steps were followed for the preparation of pristine rGO and WO3 nanomaterials, without the addition of WO3 precursors and GO, respectively.

Preparation of CNF/rGO/WO3 membrane

An optimum CNF2%/PVA0.5% blend solution was used as a matrix for the preparation of CNF/rGO/WO3 nanocomposite membranes. Different CNF/rGO/WO3 composite membranes were prepared using the solution casting method26 by varying the weight ratio of rGO/WO3 (0.5%, 1%, 3%, and 5%, m/v). The corresponding masses of rGO/WO3 were then slowly added, while stirring, to a previously prepared CNF2%/PVA0.5% dispersion. The resulting solutions were homogenized using ultrasonication, subsequently cast onto a Petri dish, and dried at room temperature for 72 h. The prepared CNF/rGO/WO3 membranes were designated as CNF/rGO/WO3-0.5, CNF/rGO/WO3-1, CNF/rGO/WO3-3, and CNF/rGO/WO3-5, based on the weight ratio of rGO/WO3 (0.5%, 1%, 3%, and 5%, m/v, respectively). It is noted that PVA is omitted in the final membrane designation as it acts only as a cross-linker with a low (0.5%) concentration; thus, we choose to simplify membrane designation. The same procedures were followed to prepare CNF/rGO, CNF/WO3, and CNF/PVA films.

Characterization of membrane materials

The surface morphology of nanomaterial and nanocomposite membranes was observed using a scanning electron microscope (SEM, Hitachi S-4800, USA) after carbon plating of the samples. The dimensions of CNF were determined using dynamic light scattering (DLS, V-770, JASCO International Co. Ltd, Japan). Functional group identification was achieved using a Fourier transform infrared spectrometer (FT-IR, iS50ABX Thermoscientific, Germany). The samples were prepared and measured in a frequency range of 4000 to 400 cm−1. The water contact angle of the membranes was measured using a Dataphysics OCA25 Contact Angle Measuring System (Dataphysics, Germany) by carefully dropping a 2 μL water droplet on ionogel surfaces using a syringe. The surface area was determined from the N2 adsorption isotherms collected at 77 K using the Brunauer–Emmett–Teller (BET) methods (SA-9600, USA). The residual MB concentration was determined using a UV/visible spectrophotometer (V-770, JASCO International Co. Ltd, Japan).

Photocatalytic degradation study

The pollutant degradation capacity of the as-prepared photocatalytic membranes was carried out under visible light irradiation. A known amount of the prepared membranes was placed into an MB aqueous solution in a beaker and stirred for 20 min in the dark to reach the adsorption–desorption equilibrium. The mixture solution was then irradiated with an LED lamp (300 W, >380 nm). After irradiation, the concentration of MB was measured using a UV/vis spectrophotometer at its λmax (664 nm) at different time intervals. The photoinduced removal efficiency (%R) and rate constant were calculated using eqn (1) and (2).27
 
image file: d5ra03227j-t1.tif(1)
 
image file: d5ra03227j-t2.tif(2)
where; C0 and C are the initial dye concentration and dye concentration at time t, respectively, and kap is the apparent pseudo-first-order rate constant.

Results and discussion

Dimensional analysis of CNF

The size distribution obtained from DLS for the milky white CNF suspension (0.015 %wt suspension) showed a single sharp peak with an effective hydrodynamic diameter of 38.37 nm (from 10 to 100 nm) with an observed unimodal peak. The Z average particle size was 7.513 nm, indicating the obtained cellulose suspension from acid hydrolysis of cellulose contains nanoscale particles. Similarly, Zhou et al. prepared a nanocellulose with a size of 115 nm from sisal fibers.28 Chattopadhyay and Patel isolated an approximately 348 nm CNF with a unimodal peak from rayon fibers.29 Beyan et al. reported a cellulose fiber in the nano range (66 nm) that was isolated from enset.30 The polydispersity index (PDI) was 0.857 (>0.1), confirming that the nanocellulose suspension had low uniformity.23

Optimization of membrane fabrication

The overall membrane fabrication process using solvent casting is depicted in Fig. 2. Parameters that tend to affect the morphology, integrity, and application performance of the final membranes were optimized during fabrication by varying the weight ratio of PVA and rGO/WO3 fillers.
image file: d5ra03227j-f2.tif
Fig. 2 Schematic representation of membrane fabrication by solution casting.

A lower concentration of PVA was assessed for its suitability for preparing a less dense CNF/PVA composite scaffold, and the corresponding effects on hydrophilicity, surface area, and porosity of the scaffold material and subsequent membranes were investigated. The average pore size, pore volume, and specific surface area of CNF/PVA, CNF/rGO/WO3 (0.5% PVA), and CNF/rGO/WO3 (2% PVA) membranes exhibited a typical type-I isotherm, indicating the presence of microporous structures (Table 1). With higher specific surface area and higher porosity, the CNF/rGO/WO3 (0.5% PVA) was perceived to be a more suitable membrane than CNF/rGO/WO3 (2% PVA), because both features can contribute to improved mass transfer and photodegradation performance. The higher surface area and porosity could be due to the lower volume proportion of PVA in the membrane. The CNF/rGO/WO3 membrane with 0.5% PVA loading also exhibited a higher surface area and porosity than the CNF/PVA scaffold material. This could be due to the ability of rGO/WO3 nanomaterials to act as spacers, thereby exposing the intercalated structure of CNF/rGO/WO3, which is characterized by an abundance of channel structures. The introduction of rGO/WO3 also prevented the folding behavior of the CNF/PVA scaffold that was observed when the film was immersed in water. The large surface areas could increase the number of active reaction sites, promoting the absorption and photodegradation of reactants.31

Table 1 Specific surface area, average pore size, and average pore volume of the CNF/PVA scaffold and the CNF/rGO/WO3 membranes
Membrane Surface area (m2 g−1) Average pore size (nm) Average pore volume (cm3 g−1)
CNF/PVA (0.5% PVA) 652.9 1.278 ± 0.406 0.106 ± 0.044
CNF/rGO/WO3 (0.5% PVA) 684.3 1.294 ± 0.406 0.110 ± 0.044
CNF/rGO/WO3 (2% PVA) 669.8 1.290 ± 0.396 0.106 ± 0.044


Morphological analysis

The SEM images of CNF, CNF/PVA, and CNF/rGO/WO3 are presented in Fig. 3a–c. CNF formed uniform net structures with no observable aggregates (Fig. 3a). The uniform continuous network structure was maintained after the addition of PVA (Fig. 3b), confirming the good dispersion of the CNF/PVA blend and the suitability of the film drying process. The entangled CNF network structure is still clearly visible in the denser CNF/PVA film, forming a dense and interconnected scaffold. The introduction of rGO/WO3 imparts denser and more uniform membrane morphology (Fig. 3c). The progressive changes towards denser morphology while maintaining uniformity ensure the uniform distribution of rGO/WO3 throughout the CNF/PVA matrix, demonstrating the successful fabrication of the CNF/rGO/WO3 composite membrane, further affirming the compatibility of membrane materials. The hydrogen bonding interactions between the hydroxyl-group-rich CNF and PVA, as well as the similar interactions with rGO/WO3 composite nanomaterials, could contribute to membrane integrity.32–34 This high integrity of the constituent materials also improves membrane stability and performance.
image file: d5ra03227j-f3.tif
Fig. 3 SEM images of (a) CNF, (b) CNF/PVA, (c) and CNF/rGO/WO3 membrane, and (d) water contact angles of the CNF/PVA, CNF/rGO/WO3 (0.5% PVA), and the CNF/rGO/WO3 (2% PVA) membrane.

The water contact angle (WCA) of CNF/PVA, CNF/rGO/WO3 (0.5% PVA), and CNF/rGO/WO3 (2% PVA) were found to be comparable, with values of 51.29°, 52.50° and 53.19°, respectively (Fig. 3d). These results indicate that the scaffold and composite membranes exhibit hydrophilic surfaces, as evidenced by their WCA of less than 90°. The hydrophilicity of the CNFs and PVA is attributable to the abundance of hydroxyl functional groups, even after membrane formation, as substantiated by the FTIR results (Fig. 4).35


image file: d5ra03227j-f4.tif
Fig. 4 FTIR spectra of the CNF, CNF/PVA, WO3, rGO, CNF/WO3, and the CNF/rGO/WO3 membrane.

Fourier-transform infrared spectroscopy

The FTIR spectra of the WO3, CNF, CNF/PVA, CNF/rGO, CNF/WO3, and CNF/rGO/WO3 membranes are shown in Fig. 4. The spectrum of the CNF/PVA film demonstrates that the peak at 3323 cm−1 in the CNF spectrum shifted to 3330 cm−1 in the spectrum of the CNF/PVA film. The peaks of pure WO3 nanoparticles were observed at 3330, 2915, 1725, and 805 cm−1. The peak centered at 805 cm−1 could be attributed to the stretching vibrations of W–O–W bonding.36 A peak at 2915 cm−1 corresponded to C–H bonding.37 The peaks at 1725 and 3330 cm−1 were assigned to the –OH bending and stretching vibrations of water molecules.32 The peaks at around 1000–1300 cm−1 in CNF/rGO/WO3, due to C–OH stretching (1060 cm−1) and C–O–C bending vibrations (1243 cm−1), are weakened in comparison to the peaks in CNF/PVA due to the presence of rGO and WO3 nanoparticles.37

Optical properties

The UV-vis diffuse reflectance spectra of rGO, WO3, rGO/WO3, and CNF/rGO/WO3 membranes are shown in Fig. 5. The edge of the absorption spectra of the rGO is red-shifted compared to that of the pristine WO3 (Fig. 5a and b). While the pristine WO3 exhibited a lower degree of absorption, its absorption exhibited a slight increase when WO3 was composited with rGO in the WO3/rGO (Fig. 5c). Furthermore, the ternary nanocomposite membrane (CNF/rGO/WO3) absorbed light in the UV/vis region (300 to 800 nm) (Fig. 5d), which is consistent with the findings of a previous study.38 The substantial light absorbance enhancement and shift in the visible light region of the absorption spectrum for CNF/rGO/WO3 is primarily attributed to the synergistic effect of rGO and WO3, which facilitates light adsorption.37
image file: d5ra03227j-f5.tif
Fig. 5 UV/vis diffuse reflectance spectroscopy (DRS) of (a) rGO, (b) WO3, (c) rGO/WO3, and the (d) CNF/rGO/WO3 photocatalytic membrane material.

Tauc's equation (eqn. (3)) was used to extrapolate the optical band gap of the as-prepared pristine and membrane composite photocatalyst materials.

 
(αhv) = A(hvEg)n; α = 4πk/λ (3)
where α is the absorption coefficient, λ is the wavelength of the incident photon, is the incident photon energy, n = 1/2 and 2 for direct and indirect recombination, respectively, A is an arbitrary coefficient, and Eg is the optical band gap of the material.38

For the indirect bandgap metal oxide semiconductor, WO3 is presumed to have n = 2.39 By extrapolating the graph of (αhν)1/2 vs. hν, the optical band gaps of rGO, WO3, rGO/WO3, and CNF/rGO/WO3 membranes were determined to be 1.4, 2.4, 2.0, and 2.27 eV, respectively (Fig. 6a–d). The 2.27 eV band gap of the ternary nanocomposite membrane could be due to the strong absorbance of rGO over the wide UV/vis range.38 Therefore, the CNF-based WO3 and rGO compositing membranes reduce the optical band gap of WO3 and enhance its photocatalytic activity.


image file: d5ra03227j-f6.tif
Fig. 6 Tauc plots of (αhν)1/2 vs. hν of (a) rGO, (b) WO3, (c) rGO/WO3, and (d) the CNF/rGO/WO3 photocatalytic membrane material.

Mechanical properties

The mechanical properties of membranes are closely related to their durability and lifespan.40 The tensile strength and elongation at break (ε) of CNF/PVA, CNF/rGO, and CNF/WO3 composite membranes were evaluated using a universal testing machine (UTM); the results are shown in Fig. 7. According to the data, the CNF/rGO/WO3 composite membrane exhibited an increased tensile strength of 28.99 MPa, which is an improvement over CNF/rGO (25.9 MPa) and CNF/WO3 (15.99 MPa), due to synergistic effects between the rigid WO3 and rGO network. Thus, H-bonding and electrostatic interactions are favored between rGO/WO3 and negatively charged CNF surfaces.41 However, the elongation at break decreased significantly from 4.2% to 0.82%. This change could be attributed to the incorporation of the rGO/WO3 composite into the CNF matrix, which significantly enhanced the mechanical strength of the membrane. The presence of photocatalytic materials likely creates an interlocking chain structure between rGO/WO3 and the CNF matrix, resulting in a more compact and solid membrane structure.17 Additionally, the membrane coated with PVA exhibited the highest tensile strength due to the cross-linking effect of PVA on the CNF-based membrane surface, further reinforcing the mechanical integrity of the membrane.
image file: d5ra03227j-f7.tif
Fig. 7 Histogram showing the mechanical properties of CNF/PVA, CNF/WO3, CNF/rGO, and CNF/rGO/WO3 photocatalytic membrane materials.

Optimization of photocatalytic degradation parameters

Effect of catalyst dosage. A series of tests were performed to investigate the effect of the rGO/WO3 weight ratio in CNF/rGO/WO3-X on the degradation of MB. The experimental design involved the introduction of varying amounts of rGO/WO3 in the 0.45 g CNF/rGO/WO3-X (X = 0.5%, 1%, 3%, and 5%) while maintaining a constant MB solution concentration of 5 mg L−1 in a 100 mL volume. The photodegradation process was carried out under irradiation with a 300 W xenon lamp. Degradation efficiencies of 50.1%, 54.1%, 89.3%, and 69.7% were obtained with CNF/rGO/WO3-0.5, CNF/rGO/WO3-1, CNF/rGO/WO3-3, and CNF/rGO/WO3-5 membrane at 120 min, respectively (Fig. 8). As the catalyst dosage (rGO/WO3) increased, the photodegradation efficiency of the CNF/rGO/WO3 also improved in a certain range due to the increasing number of active sites in the catalyst (Fig. 8a). However, the CNF/rGO/WO3 membrane could not use the excess active sites provided by the CNF/rGO/WO3-X hybrid membrane at a high catalyst dose, resulting in a saturated degradation rate.43
image file: d5ra03227j-f8.tif
Fig. 8 (a) Plots of C/C0 vs. irradiation time for photodegradation of MB in the presence of CNF/rGO/WO3-0.5, CNF/rGO/WO3-1, CNF/rGO/WO3-3, and CNF/rGO/WO3-5 membranes under visible light; and (b) photodegradation rate of MB in the presence of CNF/rGO/WO3-0.5, CNF/rGO/WO3-1, CNF/rGO/WO3-3, and CNF/rGO/WO3-5 membranes under visible light.

The surface area becomes a significant issue in the photocatalysis process because the photodegradation of an organic dye generally occurs on the photocatalyst surface. The results indicate that MB degraded more rapidly in the presence of CNF/rGO/WO3-3 (89.3%) than with CNF/rGO/WO3-0.5 (50.1%), CNF/rGO/WO3-1 (54.1%), and CNF/rGO/WO3-5 (69.7%) membranes under visible light. The optimal amount of rGO/WO3 plays a significant role because rGO significantly increases the surface area and enhances the photodegradation of MB. Although CNF itself does not exhibit photocatalytic activity, its supportive role in the composite membranes is crucial in determining the overall performance. In some cases, cellulose can impede light absorption, hinder electron generation, and diminish the degradation effect of the composite.14,19,44

To better illustrate the degradation rates, the rate constants (k) for the degradation of MB were calculated from the well-known first-order kinetics equation ln(C0/C) = kt (Fig. 8b). The degradation rate constant (k) was determined from the slope of the line. The CNF/rGO/WO3-3 membrane (0.05195 min−1) shows a higher degradation rate than CNF/rGO/WO3-0.5 (0.00923 min−1), CNF/rGO/WO3-1 (0.00942 min−1), and CNF/rGO/WO3-5 (0.01192 min−1) under visible irradiation, based on the result of ln(C0/C) versus irradiation time for the photocatalytic degradation of MB.

Effect of dye concentration. The effect of dye concentration on the degradation efficiency of the photocatalytic membranes was investigated at the optimized catalyst dosage (0.45 g of CNF/rGO/WO3-3) by varying the initial dye concentration (5–25 mg L−1) (Fig. 9). Maximum degradation was observed at the initial dye concentration of 5 mg L−1 (89.3%) at 120 min. When the dye concentration was increased from 5 mg L−1 to 25 mg L−1, the degradation efficiency of the catalyst decreased. This indicates that, as dye concentration increases, excess dye molecules can be adsorbed on the catalyst surface, thereby reducing the number of active sites of the catalysts.45 This suggests that the system's hydroxyl radical production is insufficient for efficient dye degradation at greater concentrations, similar to previous studies.46 Therefore, as a greater proportion of the catalyst surface is occupied by dye molecules, fewer hydroxyl radicals can be formed and have a short lifetime to react with dye molecules. Further increases in dye concentration cause the dye molecules to be adsorbed on the photocatalyst surface and more hydroxyl radicals are required for dye molecule degradation. On the other hand, the formation of hydroxyl radicals on the catalyst surface remains constant for a given light intensity, catalyst dose, and irradiation time. Hence, the number of hydroxyl radicals are insufficient for degrading dyes at high concentrations, diminishing the photodegradation efficiency. A similar result was reported for the dyes of MB, revealing the maximum degradation at an initial dye concentration of 5 mg L−1.47
image file: d5ra03227j-f9.tif
Fig. 9 Effect of the initial MB dye concentration (5, 10, 15, 20, and 25 mg L−1) on the photodegradation performance of the photocatalytic membranes under visible light irradiation.

Photocatalytic performance of the composite membrane

The photocatalytic activity of the membrane at the optimized catalyst dose and optimized initial dye concentration was evaluated by observing the degradation of MB (5 mg L−1) in the presence of the CNF/rGO/WO3-3 membrane (0.45 g) under visible light irradiation (Fig. 10). It can be observed that the absorbance intensity of the MB absorption peak decreased gradually. With increased irradiation time, the characteristic peak of MB absorption in the presence of CNF/rGO/WO3-3 membrane was gradually reduced without shifting its optimum absorption peak position (Fig. 10a). This proved that the MB dye was completely degraded into CO2 and water using the optimum catalyst dosage membrane (CNF/rGO/WO3-3) without producing organic intermediates during the irradiation time.
image file: d5ra03227j-f10.tif
Fig. 10 (a) Photodegradation of MB by the CNF/rGO/WO3-3 membrane after visible-light irradiation for 120 min (the inset shows an image of MB solutions before and after irradiation at time intervals), and (b) plots of C/C0 vs. irradiation time for the photocatalytic degradation of MB in the presence of WO3, CNF/WO3, and CNF/rGO/WO3-3 membranes under visible light.

The photocatalytic degradation MB solution at a concentration of 5 mg L−1 under visible light irradiation, in the presence of catalysts WO3, CNF/WO3, and CNF/rGO/WO3 is illustrated in Fig. 10b. The enhanced photocatalytic activity of WO3 nanoparticles is primarily responsible for the superior degradation observed using the CNF/rGO/WO3 membranes under visible light irradiation and was evaluated using MB dye solution after 120 min of visible light exposure. Based on the C/C0 values (Fig. 10b), pure WO3 exhibited a relatively low degradation efficiency of only 50.13%, attributed to its poor light absorption activity and rapid recombination of electron–hole pairs. The binary composite CNF/WO3 showed a higher degradation efficiency of 53.75%. However, under the same conditions of irradiation and time, the CNF/rGO/WO3-3 membrane composite exhibited a higher degradation efficiency of 89.3%, which could be due to a greater electron acceptor ability, enhanced visible light absorption, larger surface area, and synergistic interactions among the components within the composite membrane.

Detection of reactive species

Reactive species detection tests were conducted to identify the principal active species in the photocatalytic process. Active trapping agents, including ascorbic acid (AA), isopropanol (IPA), and formic acid (FA), were utilized as superoxide radical (˙O2), hydroxyl radical (˙OH), and hole (h+) scavengers, respectively. All scavengers were employed at an equivalent concentration of 0.5 mmol L−1.48,49 The photodegradation of MB was slightly decreased with the addition of FA and IPA (Fig. 11). At 120 min reaction time, ∼78.8% and ∼61.74% decomposition of MB occurred in the presence of IPA and AA, respectively. However, ∼89% photodegradation of MB occurred when no scavenger was present. The rate of MB degradation was significantly slower in the presence of AA, with only 47.2% of MB decomposed within 120 min. These results indicate that h+ and ˙O2, play a critical role in MB photodegradation.
 
image file: d5ra03227j-t3.tif(4)
 
EVB = ECB + Eg (5)
where ECB is the conduction band edge potential, EVB is the valence band edge potential, and X is the absolute electronegativity of the photocatalyst (the X values of WO3, rGO, rGO/WO3, and CNF/rGO/WO3 are 6.57 eV, 5.85 eV, 9.08 eV, and 6.05 eV, respectively50), E is the hydrogen scale's free electron energy (4.5 eV), and Eg is the band-gap energy. The values of EVB and ECB for WO3, rGO, rGO/WO3, and CNF/rGO/WO3 were calculated using eqn (4) and (5), respectively. The results are shown in Table 2.

image file: d5ra03227j-f11.tif
Fig. 11 Effect of different scavengers (ascorbic acid (AA), isopropanol (IPA), and formic acid (FA)) on the photodegradation reaction of MB over the CNF/rGO/WO3-3 composite membrane. The Eg values for WO3, rGO, rGO/WO3, and CNF/rGO/WO3 are 2.4, 1.4, 2.0, and 2.27 eV, respectively, and the respective EVB and ECB band edge positions are +3.27, +2.1, +2.15, and +2.69 eV vs. NHE and −0.87, −0.6, −0.65, and −0.42 eV vs. NHE by employing Milliken's electronegativity equations (eqn (4) and (5)).
Table 2 Electronegativity (X), band gap (Eg), conduction band (ECB) position and valence band (EVB) of the photocatalysts on NHE
Photocatalyst X (ev) Eg (ev) ECB (ev) EVB (ev)
WO3 6.57 2.40 0.87 3.27
rGO 5.85 1.40 0.60 2.10
rGO/WO3 6.20 2.00 0.65 2.15
CNF/rGO/WO3 6.05 2.27 0.42 2.69


The enhanced degradation of MB dye in the presence of a CNF/rGO/WO3 composite membrane could be due to the delayed recombination rate of the electron (e) and holes (h+) due to the electron transfer mediator nature of rGO in CNF/rGO/WO3 membrane, as shown in Fig. 12. The following reactions could be suggested to occur at the interface between the prepared composite membrane and the MB dye solution.

 
CNF/rGO/WO3 + hveCB + hVB (6)
 
eCB + rGO/WO3 → rGO(eCB) (7)
 
rGO(eCB) + O2 → −O2˙ (8)
 
WO3(eCB) + H2O → OH˙ (9)
 
CNF/rGO/WO3(hVB) + ˙OH + −O2˙ + MB → CO2 + H2O + other byproducts (10)


image file: d5ra03227j-f12.tif
Fig. 12 Proposed mechanism for the photodegradation of MB over the CNF/rGO/WO3 composite membrane.

Upon exposure of the CNF/rGO/WO3 membrane to visible light, photo-induced electrons are excited from the VB to the CB of the WO3, generating photo-induced holes in the VB of WO3 (eqn (6)). Then, the photoexcited electrons in the CB of the WO3 are transferred into the rGO sheet, which leads to complete separation of the electron–hole pairs (eqn (7)). Thereafter, the transferred electrons on the surface of the rGO combine with O2 and form ˙O2 species (eqn. (8)). Furthermore, the produced ˙O2 reacts with the MB pollutant, degrading it into less toxic by-products, such as CO2 and H2O. On the other hand; holes (h+) in the VB of WO3 allow H2O2 to react with water to produce HO˙ radicals. The HO˙ then mineralizes the MB dye into H2O and CO2 molecules (eqn (9)).51 In summary, the rGO sheet in the composite membrane served as an electron transporter and collector by decreasing the recombination rate and increasing the photogenerated charge carrier lifetime, leading to enhanced photocatalytic performance of the CNF/rGO/WO3 membrane (eqn. (10)).

Reusability test

The reusability of the photocatalyst membrane is another important aspect of photodegradation. The long-term stability of the CNF/rGO/WO3 membrane was evaluated by its repeated use for the removal of MB under visible light within 120 min. At the end of each test, the samples were washed and recovered with distilled water and reused in ten cycles of photodegradation. The recycled CNF/rGO/WO3 photocatalyst membrane demonstrated a photodegradation efficiency for MB that was similar to the original, exhibiting a slight decrease in efficiency when utilized up to six times (Fig. 13a).
image file: d5ra03227j-f13.tif
Fig. 13 (a) Reusability study of the CNF/rGO/WO3 membrane for the degradation of MB under the optimal conditions. (b) FTIR spectra of the CNF/WO3, CNF/rGO, and CNF/rGO/WO3 membranes after photodegradation.

Notably, the photocatalyst membrane achieved 48% degradation of MB after ten cycles, suggesting its durability and effectiveness under prolonged use. Furthermore, the molecular characteristics of the CNF/rGO/WO3 membrane after the photocatalytic process were studied using FTIR techniques. Compared with the FTIR spectra before photodegradation (Fig. 4), the spectra after photodegradation (Fig. 13b) show some changes in peak positions and intensities, indicating surface modifications and possible fouling interactions. The O–H stretching peak becomes slightly broader and shifts to 3278.4 cm−1, while the C–H stretching peak shifts to 2915 cm−1. The C[double bond, length as m-dash]O stretching peak appears at 1727 cm−1 and the C–O–C stretching peak shifts to 1033.6 cm−1, both with reduced intensity. These changes suggest possible adsorption of foulants on the membrane surface and minor chemical modifications during the photodegradation process. The strong W–O–W peak, which appears around 583.9 cm−1 before photodegradation in the WO3, CNF/WO3, and CNF/rGO/WO3 membranes, remains present after photodegradation but becomes slightly sharper. This indicates that WO3 remains chemically stable throughout the photodegradation process, maintaining its structural integrity.17,42

The CNF/rGO/WO3 membrane is effective for the photodegradation of MB. Its characteristics, including preparation method, irradiation time, target concentration, weight dose, and photocatalytic degradation efficiency are compared to those reported in other studies in Table 3. The CNF/rGO/WO3 membrane is prepared using the solution casting method, which is similar to the blend casting techniques used for the membrane preparation. Various studies reported irradiation times ranging from a maximum of 7 days to a minimum of 90 min. Our results demonstrate a significant photodegradation efficiency of 89.3% at 120 min irradiation under visible light, which is comparable to the time needed for the complete degradation of various molecules under similar conditions, demonstrating the excellent photocatalytic reactivity of the system with organic pollutants. Notably, while the highest photodegradation reported was 100% for the CNF–ZnO–Ag membrane, that system required 300 min of irradiation time. This underlines the efficiency of the current study in achieving effective MB degradation within a shorter irradiation time.

Table 3 Comparative MB photodegradation performance of the CNF/rGO/WO3 membrane with other related work
No Membrane Method of preparation Weight Target concentration (mg L−1) Irradiation time Light source Photocatalytic degradation efficiency Ref.
1 2D-WO3/CA Scraper method 478.13 mg 10 7 days Sunlight 75% 52
2 CA-ZnO-NP Spin-coating techniques 2000 mg 5 120 min Sunlight 74% 53
3 CNF–ZnO–Ag Drop-drying method 40 mg 10 300 min Visible light 100% 19
4 Cellulose-TiO2 Coating solution 5 90 min UV irradiation 90% 54
5 WO2.72@Fe3O4@CNF Freeze-drying 10 120 min Solar irradiation 85% 55
6 CNF/rGO/WO3 Blend casting technique 450 mg 5 120 min Visible light 89% This study


Conclusion

The environmentally friendly ternary composite CNF/rGO/WO3 nanocomposite membrane was prepared using the blend solution casting technique. The structural properties of the nanostructures and the membranes were investigated. The morphological results showed that the rGO/WO3 nanoparticles could be uniformly distributed along the CNF/PVA with heating, increasing the surface area, the mechanical properties of the membrane, and the number of active sites available for the removal of dye molecules. The CNF/rGO/WO3 membrane can effectively stimulate the separation of electron–hole pairs generated by rGO/WO3 in the nanocomposite and enhance the photocatalytic capacities of the CNF-based membrane under visible-light irradiation. The CNF/rGO/WO3 composite membrane exhibits a higher photodegradation capacity than pure WO3 MB at 120 min. The design provides a green approach with excellent photodegradation performance for the remediation of organic pollutants, and can enable further applications for photocatalysis and wastewater treatment. Compared with WO3 and CNF/WO3, the CNF/rGO/WO3 membrane exhibits faster photocatalytic degradation rates and efficient reusability. The CNF-supported photocatalytic membrane features a simple design that is economical, requires low energy consumption, and is environmentally friendly.

Data availability

The data supporting this article have been included as part of the ESI.

Author contributions

MTY: investigation, material preparation, and writing original draft; GG: conceptualization, supervision, methodology, writing: review and editing; AKM: conceptualization, supervision, methodology, writing: review and editing; KNM: conceptualization, supervision, methodology, writing: review and editing.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

The work was partially supported by the Addis Ababa Science and Technology University under Grant Number: EA-552/4-1/20.

References

  1. K. K. Ibraheem, O. Y. T. Al-Janabi, E. T. B. Al-Tikrity, I. S. Ali, B. M. Ahmed and P. J. S. Foot, Inorg. Chem. Commun., 2025, 173, 113844 CrossRef CAS.
  2. M. T. Yifira, A. K. Mersha, G. Gebreslassie and K. N. Mekonnen, Carbohydr. Polym. Technol. Appl., 2024, 8, 100589 CAS.
  3. D. Singh, F. Khan, V. K. Jain and S. Bhattacharya, J. Indian Chem. Soc., 2024, 101, 101356 CrossRef CAS.
  4. H. Fan, Y. Quan, X. Zhao, H. Chen, S. Yu, Q. Zhang and Y. Zhang, J. Appl. Polym. Sci., 2017, 134, 1–9 CrossRef.
  5. Y. Ibrahim, E. Abdulkarem, V. Naddeo, F. Banat and S. W. Hasan, Sci. Total Environ., 2019, 690, 167–180 CrossRef CAS PubMed.
  6. C. Vilela, C. Moreirinha, A. Almeida, A. J. D. Silvestre and C. S. R. Freire, Materials, 2019, 12, 1404 CrossRef CAS PubMed.
  7. W. Wang, Q. Bai, T. Liang, H. Bai and X. Liu, Polymers, 2017, 9, 455 CrossRef PubMed.
  8. C. Zhu, P. Liu and A. P. Mathew, ACS Appl. Mater. Interfaces, 2017, 9, 21048–21058 CrossRef CAS PubMed.
  9. G. Huang, D. Zeng, M. Wei, Y. Chen and A. Ma, Inorg. Chem. Commun., 2025, 172, 1–10 CrossRef.
  10. B. Parasuraman, B. Kandasamy, V. Vasudevan and P. Thangavelu, Environ. Sci. Pollut. Res., 2024, 31, 46591–46601 CrossRef CAS PubMed.
  11. A. T. Sheikhzadeh and A. Mouradzadegun, Photochem. Photobiol. Sci., 2023, 22, 837–855 CrossRef PubMed.
  12. C. Yadav, A. Saini and P. K. Maji, Carbohydr. Polym., 2017, 165, 276–284 CrossRef CAS PubMed.
  13. V. Ramar and K. Balasubramanian, ACS Appl. Nano Mater., 2021, 4, 5512–5521 CrossRef CAS.
  14. S. S. Mehta, D. Y. Nadargi, M. S. Tamboli, T. Alshahrani, V. R. M Reddy, E. S. Kim, I. S. Mulla, C. Park and S. S. Suryavanshi, Sci. Rep., 2021, 11, 1–17 CrossRef PubMed.
  15. W. Zhu, F. Sun, R. Goei and Y. Zhou, Appl. Catal. B Environ., 2017, 207, 93–102 CrossRef CAS.
  16. K. Dhivyaprasath, B. R. Wiston, M. Preetham, S. Harinee and M. Ashok, Optik, 2021, 244, 167593 CrossRef CAS.
  17. T. D. Kusworo, M. Yulfarida, A. C. Kumoro, S. Sumardiono, M. Djaeni, T. A. Kurniawan, M. H. D. Othman and B. Budiyono, J. Environ. Chem. Eng., 2023, 11, 109583 CrossRef CAS.
  18. F. Dai, S. Zhang, Q. Wang, H. Chen, C. Chen, G. Qian and Y. Yu, Front. Chem., 2021, 9, 1–13 Search PubMed.
  19. J. Han, N. B. Tri Pham, K. Oh and H. K. Choi, ACS Omega, 2024, 9, 7143–7153 CrossRef CAS PubMed.
  20. T. A. Hameed, A. Salama and R. A. Nasr, J. Polym. Environ., 2024, 32, 4440–4455 CrossRef CAS.
  21. L. A. Worku, R. K. Bachheti and M. G. Tadesse, Int. J. Polym. Sci., 2022, 1, 155552 Search PubMed.
  22. Y. Hu, L. Tang, Q. Lu, S. Wang, X. Chen and B. Huang, Cellulose, 2014, 21, 1611–1618 CrossRef CAS.
  23. E. G. Bacha and H. D. Demsash, Extraction and Characterization of Nanocellulose from Eragrostis Teff Straw, 2021, Preprint (Version 1), available at Research Square,  DOI:10.21203/rs.3.rs-296990/v1.
  24. L. Shahriary and A. A. Athawale, Int. J. Renew. Energy Environ. Eng., 2014, 02, 58–63 Search PubMed.
  25. V. Ramar and K. Balasubramanian, Appl. Phys. A Mater. Sci. Process., 2018, 124, 1–11 CrossRef.
  26. U. Siemann, Prog. Colloid Polym. Sci., 2005, 130, 1–14 CAS.
  27. Z. Yina, Z. Li, Y. Deng, M. Xueb, Y. Chen, J. Ou, Y. Xie, Y. Luo, C. Xie and Z. Hong, Ind. Crops Prod., 2023, 197, 116672 CrossRef.
  28. Y. M. Zhou, S. Y. Fu, L. M. Zheng and H. Y. Zhan, Express Polym. Lett., 2012, 6, 794–804 CrossRef CAS.
  29. D. P. Chattopadhyay and B. H. Patel, J. Text. Sci. Eng., 2016, 6(2), 1–8 Search PubMed.
  30. S. M. Beyan, T. A. Amibo, S. V. Prabhu and A. G. Ayalew, J. Nanomater., 2021, 7492532 CAS.
  31. C. Zhao, F. Ran, L. Dai, C. Li, C. Zheng and C. Si, Carbohydr. Polym., 2021, 117343 CrossRef CAS PubMed.
  32. M. Kaur, S. Singh, S. K. Mehta and S. K. Kansal, Molecules, 2022, 27, 6956 CrossRef CAS PubMed.
  33. M. Wohlert, T. Benselfelt, L. Wågberg, I. Furó, L. A. Berglund and J. Wohlert, Cellulose, 2022, 29, 1–23 CrossRef CAS.
  34. P. Song, Z. Xu, Y. Lu and Q. Guo, Macromolecules, 2015, 48, 3957–3964 CrossRef CAS.
  35. A. Autissier, C. Le Visage, C. Pouzet, F. Chaubet and D. Letourneur, Acta Biomater., 2010, 6, 3640–3648 CrossRef CAS PubMed.
  36. L. Tie, C. Yu, Y. Zhao, H. Chen, S. Yang, J. Sun, S. Dong and J. Sun, J. Alloys Compd., 2018, 769, 83–91 CrossRef CAS.
  37. L. Yang, C. Chen, Y. Hu, F. Wei, J. Cui, Y. Zhao, X. Xu, X. Chen and D. Sun, J. Colloid Interface Sci., 2020, 562, 21–28 CrossRef CAS PubMed.
  38. V. A. Tran, T. P. Nguyen, V. T. Le, I. T. Kim, S.-W. Lee and C. T. Nguyen, J. Sci. Adv. Mater. Devices, 2021, 6, 108–117 CrossRef CAS.
  39. C. T. Nguyen, T. P. Pham, T. L. A. Luu, X. S. Nguyen, T. T. Nguyen, H. L. Nguyen and D. C. Nguyen, Ceram. Int., 2020, 46, 8711–8718 CrossRef CAS.
  40. Y. Zhang, Q. Zhang, Y. Li and L. Chen, Struct. Eng. Mech., 2014, 49(6), 745–762 CrossRef.
  41. K. Fernández, A. Llanquileo, M. Bustos, V. Aedo, I. Ruiz, S. Carrasco, M. Tapia, M. Pereira, M. F. Meléndrez, C. Aguayo and L. I. Atanase, Polymers, 2023, 15(12), 2752 CrossRef PubMed.
  42. I. Koyuncu, B. Eryildiz, R. Kaya, Y. Karakus, F. Zakeri, A. Khataee and V. Vatanpour, J. Environ. Manage., 2023, 326, 116758 CrossRef CAS PubMed.
  43. W. Zhu, M. Han, D. Kim, J. Park, H. Choi, G. Kwon, J. You, S. Li, T. Park and J. Kim, J. Water Process Eng., 2023, 53, 103620 CrossRef.
  44. S. Elbakry, M. E. A. Ali, M. Abouelfadl, N. A. Badway and K. M. M. Salam, J. Photochem. Photobiol., A, 2022, 430, 113957 CrossRef CAS.
  45. I. K. Konstantinou and T. A. Albanis, Appl. Catal., B, 2004, 49, 1–14 CrossRef CAS.
  46. P. V. Nidheesh and R. Rajan, RSC Adv., 2016, 6, 5330–5340 RSC.
  47. S. D. Khairnar, M. R. Patil and V. S. Shrivastava, Iran. J. Catal., 2018, 8, 143–150 CAS.
  48. G. Gebreslassie, M. Gebrezgiabher, B. Lin, M. Thomas and W. Linert, Inorganics, 2023, 11(3), 119 CrossRef CAS.
  49. H. D. Weldekirstos, T. Mengist, N. Belachew and M. L. Mekonnen, Results Chem., 2024, 7, 101306 CrossRef CAS.
  50. A. Sarkar, A. Mandal, S. Mandal, S. K. Sen, D. Banerjee, S. Ganguly and K. Kargupta, J. Chem. Sci., 2024, 136(2), 1–11 Search PubMed.
  51. H. A. Alburaih, M. Aadil, S. Mubeen, W. Hassan, S. R. Ejaz, A. Anwar, S. Aman and I. A. Alsafari, FlatChem, 2022, 34, 100380 CrossRef CAS.
  52. B. Qi, T. Chen, Z. Liu, J. Qi, Q. Wang, W. Zhang and X. Li, J. Water Process Eng., 2022, 47, 102739 CrossRef.
  53. M. A. Abu-Dalo, S. A. Al-Rosan and B. A. Albiss, Polymers, 2021, 13(19), 3451 CrossRef CAS PubMed.
  54. H. K. M. Ng and C. P. Leo, Prog. Org. Coat., 2019, 132, 70–75 CrossRef CAS.
  55. S. Chaudhuri, C. M. Wu and K. G. Motora, J. Photochem. Photobiol., A, 2023, 438, 114525 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5ra03227j

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.