Next Article in Journal
Characterizing the ZrC(111)/c-ZrO2(111) Hetero-Ceramic Interface: First Principles DFT and Atomistic Thermodynamic Modeling
Next Article in Special Issue
Designing Novel Strategy to Produce Active Nanohybrids in Sunlight for Purification of Water Based on Inorganic Nanolayers, Magnetic Nanocomposites and Organic Species
Previous Article in Journal
Correction: Rubnawaz et al. Polyphenol Rich Ajuga bracteosa Transgenic Regenerants Display Better Pharmacological Potential. Molecules 2021, 26, 4874
Previous Article in Special Issue
Photocatalytic Degradation of Paracetamol in Aqueous Medium Using TiO2 Prepared by the Sol–Gel Method
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Preparation and Photocatalytic Performance of TiO2 Nanowire-Based Self-Supported Hybrid Membranes

1
Faculty of Materials Science and Engineering, University of Miskolc, H-3515 Miskolc, Hungary
2
Polymers and Petrochemicals Engineering Department, Basrah University for Oil and Gas, Basrah 61004, Iraq
3
Advanced Materials and Intelligent Technologies Higher Education and Industrial Cooperation Centre, University of Miskolc, H-3515 Miskolc, Hungary
4
European Commission, Joint Research Centre (JRC), 21027 Ispra, Italy
5
MTA-ME Materials Science Research Group, ELKH, H-3515 Miskolc, Hungary
6
Institute of Mineralogy and Geology, University of Miskolc, H-3515 Miskolc, Hungary
7
Institute of Energy and Quality Affairs, University of Miskolc, H-3515 Miskolc, Hungary
8
Institute of Physical Metallurgy, Metal Forming and Nanotechnology, University of Miskolc, H-3515 Miskolc, Hungary
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(9), 2951; https://doi.org/10.3390/molecules27092951
Submission received: 31 March 2022 / Revised: 28 April 2022 / Accepted: 3 May 2022 / Published: 5 May 2022
(This article belongs to the Special Issue Thermal and Photocatalytic Analysis of Nanomaterials)

Abstract

:
Nowadays, the use of hybrid structures and multi-component materials is gaining ground in the fields of environmental protection, water treatment and removal of organic pollutants. This study describes promising, cheap and photoactive self-supported hybrid membranes as a possible solution for wastewater treatment applications. In the course of this research work, the photocatalytic performance of titania nanowire (TiO2 NW)-based hybrid membranes in the adsorption and degradation of methylene blue (MB) under UV irradiation was investigated. Characterization techniques such as scanning electron microscopy (SEM), transmission electron microscopy (TEM), energy-dispersive X-ray spectroscopy (EDS), X-ray powder diffractometry (XRD) were used to study the morphology and surface of the as-prepared hybrid membranes. We tested the photocatalytic efficiency of the as-prepared membranes in decomposing methylene blue (MB) under UV light irradiation. The hybrid membranes achieved the removal of MB with a degradation efficiency of 90% in 60 min. The high efficiency can be attributed to the presence of binary components in the membrane that enhanced both the adsorption capability and the photocatalytic ability of the membranes. The results obtained suggest that multicomponent hybrid membranes could be promising candidates for future photocatalysis-based water treatment technologies that also take into account the principles of circular economy.

1. Introduction

Over the past decade, air and water pollution has been on an upward trend, causing concern in everyday life around the world. In general, the industrial sectors accumulate large quantities of wastewater during manufacturing processes, which contains various chemical components such as pharmaceutical residues, dyes or heavy metals and salts [1] that are costly to neutralize and remove. Semiconductors and their composites have been studied in depth to combat the problems associated with waste generation. They include TiO2 [2], ZnO [3], WO3 [4], Bi-based [5] and Ag-based [6] materials, etc. [7].
Titanium dioxide (TiO2) is one of the most widely used semiconductor for photocatalytic applications due to its advantageous properties (optical and electronic properties, chemical stability, low cost, lack of toxicity) [8]. However, its practical application is limited, as it can only be effectively excited under UV light due to its wide band gap. There have been numerous attempts to enhance the photocatalytic activity and excitability of titanium dioxide, for example, by synthesizing TiO2 with various morphologies [9], modifying it with noble metals [10,11], doping it with various elements [12,13] and preparing composites with graphene [14], graphene oxides [15] and other semiconductors [16].
Copper and copper oxide are widely used as dopants on titanium dioxide materials’ surfaces to enhance their photocatalytic performance by narrowing the bandgap and improving electron–hole separation with photoexcitation, as reported in [17,18]. Variations of iron oxides are used to modify the properties of titanium dioxide, due to their small band gap [19]. With the combination of iron oxide and titanium dioxide, shallow trap sites appear between the conduction band and the valance band that leads to a reduction in the band gap energy of TiO2 [20]. Furthermore, the radius of Fe2+ and Fe3+ ions is smaller than that of Ti4+; thus, iron ions can diffuse into the TiO2 lattice to substitute TiO2 [21]. In addition, iron oxides (magnetite, i.e., Fe3O4 and maghemite, i.e., Fe2O3) have special magnetic properties in the nanoscale range [22]. These ferromagnetic properties can be beneficial to improve the recyclability of photocatalysts and prevent particles from clumping together [23].
Many authors have reported that enhanced photocatalytic properties are expected from morphological modifications of TiO2-based photocatalytic nanomaterials, such as one-dimensional TiO2 nanostructures, including TiO2 nanowires (TiO2 NW) [19,24], TiO2 nanotubes (TiO2 NT) [25,26], TiO2 hollow spheres (TiO2 HS) [27] and TiO2 hollow fibers (TiO2 HF) [28]. These morphologies have a high specific surface area as well as other beneficial properties, as they can decrease the hole–electron recombination rate and increase the interfacial charge transfer rate [29,30]. TiO2 NW have excellent mechanical stability and chemical and physical properties. For the synthesis of TiO2 NW, mostly hydrothermal [31] and solvothermal methods [32] are used. Previous scientific results indicate that TiO2 nanowires have outstanding photocatalytic efficiency [31,32,33].
Photocatalytic membranes (PM) offer a great potential alternative for economical and eco-friendly treatments of wastewater, based on the combination of membrane filtration and photocatalysis [34,35]. Membrane filtration is widely used for drinking water treatment and wastewater treatment [36,37] due to its simple operation and effective removal of various types of pollutants. Using this technology, most organic and small amounts of inorganic substances can be decomposed using solar energy to reduce the harm of pollutants. There are various types of materials that can be used as membranes, such as natural polymers (cellulose [38,39]), green porous nano-membranes [40], graphene [41], carbon nanotube [42], ceramic [43] or synthetic polymer-based (polyvinyl alcohol (PVA), polyacrylonitrile (PAN), polyvinylidene fluoride (PVDF) etc.) [44] membranes [45]. TiO2 nanowire membranes have been investigated as mechanical microfilters and as photocatalysts to degrade pharmaceutical residues such as trimethoprim and other organic materials including dyes, phenol and humic acid under UV irradiation [46]. A. Hu et al. investigated various types of TiO2 nanostructured membranes, though the properties of CuO- and Fe2O3-modified TiO2 photocatalytic membranes were not explored.
Cellulose-based materials are promising substances that can partly or totally replace synthetic fibers as filters in masks [47] and can be used as membranes or support for other materials and to remove oil and heavy metal ions during water treatment [48]. Furthermore, cellulose has unique features in comparison to the usual supports and thus allows nanoparticles stability, reactivity, recyclability and prevents nanoparticle aggregation. Cellulose also represents a sustainable alternative to known methods with the aforementioned properties. CdS and TiO2 nanoparticle–nanocellulose hybrid composites have been used as photocatalysts for a model pollutant, i.e., methyl orange degradation. In the CdS case, 82% degradation efficiency was achieved after 90 min irradiation, and the material was reusable up to five times. [49]. A TiO2–cellulose hybrid composite was 20% more efficient in degrading methylene orange than pure TiO2 after 20 min of UV irradiation [50]. In a recent study, it was found that the photocatalytic degradation rate of TiO2–cellulose was much higher (99.72%) than that of bare TiO2 (69.18%) under UV irradiation for 30 min, because cellulose served as a support for TiO2 nanoparticles distribution and also promoted the adsorption of methyl orange molecules [51]. In this work, the aim was to develop a “green chemistry” solution to industrial wastewater effluents. Cellulose-based membranes were chosen over other materials due to their advantageous properties such as lack of toxicity, low cost, biodegradability and eco-friendliness. Other than this, cellulose membranes have excellent specific surface area, adjustable surface chemistry, hydrophilicity and mechanical strength [52].
In the current paper, we successfully developed titanium dioxide nanowire (TiO2 NW)-based hybrid membranes. The surface of TiO2 NW was decorated with iron oxide (Fe2O3) and copper oxide (CuO) nanoparticles to improve the photocatalytic performance of the as-prepared composites. Furthermore, cellulose fibers were applied as a reinforcement filler material to prepare self-supported hybrid membranes. The photocatalytic properties of the membranes were investigated against methylene blue decomposition under UV irradiation. The results showed that nanofiber-based hybrid membranes can provide an excellent alternative as environmentally friendly solutions for wastewater treatment requiring the degradation of organic pollutants.

2. Results

2.1. HRTEM and EDS Analysis of TiO2 NW@Fe2O3 and TiO2 NW@CuO Nanocomposites

Heat-treated nanocomposite samples were investigated by the HRTEM technique. Figure 1 shows HRTEM micrographs of the prepared TiO2 NW@Fe2O3 and TiO2 NW@CuO nanocomposites. These images revealed that the fabrication of both nanocomposites was successful, although different nanocomposite structures were observed during TEM. The HRTEM images showed that inorganic nanoparticles (Fe2O3 and CuO) were attached to the surface of TiO2 NW. Figure 1a,b show that Fe2O3 and CuO nanoparticles adhered on TiO2 NW, respectively, and segregated particles could not be observed.
Furthermore, from the analysis of the HRTEM images, the average particle sizes of Fe2O3 and CuO nanoparticles were calculated using the iTEM software (Olympus Soft Imaging Solutions). The average particle size of these components was determined by measuring 100 individual particles in both samples. Based on these calculations, it was found that Fe2O3 nanoparticles had a diameter in the range of 20–30 nm, as can be seen in Figure 1a, while the average diameter of CuO nanoparticles was 2–3 nm, as shown in Figure 1b.
EDS analysis was performed to determine the elements in the as-prepared nanocomposites. Figure 2a,b show the EDS spectra and confirmed that the most significant signals originated from carbon (C), oxygen (O), titanium (Ti), potassium (K), copper (Cu) and iron (Fe). The presence of TiO2 NW, Fe2O3 and CuO in the samples was confirmed by the Ti, Fe, Cu and O peaks, as shown in the spectra (Figure 2), while in the case of the TiO2 NW@Fe2O3 nanocomposite sample, the Cu peak originated from the sample holder (a lacey Cu grid), and the peak of K was due to the residual KOH that was used during the preparation of the TiO2 NW.

2.2. SEM and EDS Analysis of the Hybrid Membranes

In order to gain information about the surface morphology of the as-prepared hybrid membranes, SEM analysis was performed. The surface nature and morphology of a neat cellulose membrane (Figure 3a), the TiO2 NW (Figure 3b) and the as-prepared TiO2 NWs@Fe2O3/cellulose (Figure 3c,d) and TiO2 NWs@CuO/cellulose hybrid membranes (Figure 3e,f) were characterized by SEM. The SEM images in Figure 3a show cellulose wires with an average diameter of 1–5 µm, while the TiO2 NW had an average diameter of 10–15 nm, as observed in Figure 3b. Figure 3c,e show the produced self-supported hybrid membranes, and the SEM images in Figure 3d,f demonstrate the microstructure of TiO2 NWs@Fe2O3/cellulose (d) and TiO2 NWs@CuO/cellulose (f) hybrid membranes. In the TiO2 NWs@Fe2O3/cellulose membrane, the Fe2O3 particles stuck together resulting in bigger agglomerates and increased pore size.
To determine the elemental composition and confirm the presence of Fe2O3 and CuO nanoparticles in the as-prepared hybrid membranes, EDS analysis was performed for each sample. The data of the EDS analysis in Table 1 revealed the atomic percentages (at%) of the detected elements in the samples. The most significant signals originated from carbon (C), oxygen (O) and titanium (Ti), confirming the presence of cellulose and TiO2 NW in the hybrid membranes. Furthermore, iron (Fe) and copper (Cu) signals were detected, which related to the Fe2O3 and CuO nanoparticles. Other elements were also observed, such as sodium (Na), potassium (K) and fluorine (F), which originatied from the preparation procedure of TiO2 NW. The results of EDS analysis from HRTEM and SEM showed good correlations.

2.3. XRD and Specific Surface Area Analysis of the Hybrid Membranes

In order to determine the degree of crystallization of the as-prepared nanocomposites and membranes and to identify and describe the crystal structure of these materials, XRD analysis was performed. As can be seen in Figure 4a, the diffraction peaks at the angles 2(θ°) of 24.1°, 33.2°, 35.6°, 40.8°, 49.5°, 54.1°, 57.3° and 62.6° correspond to the reflections from (012), (104), (110), (113), (024), (116), (212), respectively. All of these peaks originated from hematite (α-Fe2O3) (JCPDS 33-0664). The diffraction peaks at the angles 2(θ°) of 25.2°, 36.2°, 37.5°, 38.5°, 48.0°, 54.3° and 56,6° correspond to the (101), (103), (004), (112), (200), (105), (211) and (118) reflections, respectively, and relate to anatase phase of the TiO2 NW (JCPDS 21-1276). Furthermore, the diffraction peaks at 11.9°, 24.2°, 43.1°, 60.1°, correspond to the crystal phase (200), (002), (602), (610) of potassium titanium oxide K2Ti6O13 (PDF no. 40-0403). All these results are in agreement with previous reports [53]. The diffraction peaks of cellulose (PDF no. 03-0289) could be observed at angles 2(θ°) of 16.3°, 22.7° and 34.5°, corresponding to the crystal planes of cellulose at (110), (110), (200) and (400) reflections, respectively.
Figure 4b shows the XRD analysis of the TiO2 NWs@CuO/cellulose hybrid membrane. As shown, the diffraction peaks located at the angles 2(θ°) of 33.1°, 35.7°, 38.1°, 47.1°, 49.4°, 53.2°, 58.6°, 62.2°, 67.3 ° and 68.1° correspond to monoclinic crystal phases of the reflections (110), (002), (200), (112), (202), (020), (202), (113), (022), (220), related to copper oxide (CuO) (JCPDS card number 45-0937), which showed correlations with earlier published results [54].
The pure materials, the membranes and the hybrid membranes were also characterized by the N2 adsorption technique to determine their specific surface areas and pore diameters, as can be seen in Table 2. It was found that there was no significant difference between the surface areas of the as-prepared hybrid membranes. Both hybrid membranes had a specific surface area of approx. 120 m2/g.

2.4. Photocatalytic Efficiency of the Hybrid Membranes

The photodegradation efficiency of the synthesized membranes was tested using methylene blue dye as a model pollutant, under UV light irradiation. The results revealed the almost complete removal of MB. Approx. 90% of the initial MB dye was decomposed by the hybrid membrane containing a photocatalyst, i.e., the TiO2 NW@Fe2O3 and TiO2 NW@CuO nanocomposites. Figure 5 presents the removal efficiency of MB of the TiO2 NW@Fe2O3/cellulose and TiO2 NW@CuO/cellulose hybrid membranes under UV light.
It is important to note here that only the hybrid membranes are shown because it was reported that structure deformations appear in pure cellulose exposed to UV light. This is the reason why we applied a lower amount of cellulose in the hybrid membranes. The high removal of MB by the membranes reflects their potential to degrade organic dye molecules. The excellent degradation efficiency could be attributed to the synergistic effects of the TiO2 NW with nanoparticles (Fe2O3, CuO) on their surface and the adsorption of MB on cellulose. As the present study was carried out under UV light, significant differences could be observed under visible light; therefore, we will explore this possibility in a further study. In addition, since pure cellulose does not contain any photocatalyst, a pure cellulose membrane is not expected to possess photocatalytic activity. However, its adsorption capacity can be non-negligible; therefore, we performed adsorption studies for 2 h of this and other membranes, as shown in Figure 6. Since a small amount of cellulose was used in the hybrid membranes for the reason mentioned above and the adsorption capacities appeared to be similar for all materials, it can be concluded that titanate alone is capable of adsorbing MB molecules. In the literature, several works have reported enhanced adsorption of different organic pollutants onto Fe2O3 and CuO, making these compounds promising components of membranes for wastewater treatment [19,55].

3. Materials and Methods

3.1. Materials

Titanium dioxide (TiO2) (P25), anatase nanopowder (with an average diameter of 21 nm), ammonium hydroxide (NH4OH, 25%) and sodium hydroxide (NaOH) were purchased from Sigma Aldrich (Budapest, Hungary). Hydrochloric acid (HCl, 37 %) and copper (II) acetate monohydrate (Cu(OOCCH3)2 × H2O) were obtained from VWR Chemicals (Debrecen, Hungary). Potassium hydroxide (KOH) was purchased from Thomasker Co., (Budapest, Hungary). Iron chloride hexahydrate (FeCl3 × 6H2O) was purchased from Scharlab, (Debrecen, Hungary). Cellulose originated from DIPA Ltd. (Miskolc, Hungary). A polyvinylidene (PVDF) filter membrane with pore size of 0.1 µm and diameter of 47 mm (Durapore-VVLP04700) was used for hybrid membranes’ preparation.

3.2. Synthesis of TiO2 Nanowires (TiO2 NW)

Recently, we showed the preparation of TiO2 NW using the so-called solvothermal process [56]. In brief, a homogeneous TiO2 suspension was transferred into a Teflon®-lined autoclave. The autoclave was kept in a dryer at 160 °C for 24 h. The as-prepared TiO2 NWs were washed with 0.1 M HCl and deionized water until a neutral pH was reached. The products were dried and calcined at 500 °C for 1 h.

3.3. Synthesis of TiO2 NW@Fe2O3/Cellulose Membranes

For the synthesis of the hybrid membrane, firstly, a TiO2 NW@Fe2O3 nanocomposite was prepared. The calculated amount of FeCl3 × 6H2O precursor was dissolved in 100 mL of distilled water to obtain a homogeneous solution. Then, 0.95 g of previously prepared TiO2 NWs was added to the solution and stirred for 1 h, then transferred to the autoclave for 9 h at 90 °C. The product obtained was washed with 0.1 M NaOH to adjust the pH to 7, dried for 12 h at 50 °C and then calcinated for 2 h at 500 °C using a static furnace. The load of the Fe2O3 nanoparticles in the final composition was 5 w/w %. In the next step, 0.2 g of the as-prepared TiO2 NW@Fe2O3 nanocomposite powder was dispersed in 100 mL of distilled water for 1 h, then 5 g of cellulose solution (1 w/w %) was added to the solution and stirred for 1 h. Finally, the preparation of the cellulose-based hybrid membranes was accomplished by vacuum filtration through a PVDF membrane (total mass of 250 mg/membrane), followed by drying in a furnace for 30 min at 40 °C.

3.4. Synthesis of TiO2 NW@CuO/Cellulose Membranes

In the same way, a calculated amount of (Cu(CH3COO)2 × H2O) was dissolved in 100 mL of EtOH and left under vigorous stirring for 30 min to ensure complete dissolution. Then, 0.95 g of TiO2 NW was added directly to the solution, which was kept under vigorous stirring for 1 h. The mixture was poured into an autoclave and placed in a static furnace at 150 °C for 12 h. The final product was collected and washed using vacuum filtration and calcinated for 2 h at 500 °C. The load of the CuO nanoparticles in the final composition was 5 w/w %. To prepare the TiO2 NW@CuO/cellulose membranes, 0.2 g of the above prepared composite was dipped into 100 mL of EtOH for 1 h, then 5 g of cellulose (1%) was added to the solution for another 1 h, and finally, the membrane was obtained by vacuum filtration using a PVDF membrane (total mass of 250 mg/membrane), followed by drying in a furnace for 30 min at 40 °C.

3.5. Characterization Techniques

For the qualitative characterization, high-resolution transmission electron microscopy (FEI Tecnai G2 F20 HRTEM, Hillsboro, OR, USA) was used to analyze the morphology of the synthetized TiO2 NW@Fe2O3 and TiO2 NW@CuO nanocomposites. To prepare the samples for HRTEM, the nanocomposites were dispersed in ethanol and sonicated for 5 min. On a Cu TEM-grid (300-mesh copper grids, lacey carbon, Ted Pella Inc., Redding, CA, USA), we placed a droplet of each suspension. The diameter of the materials was determined using the ImageJ software, utilizing the HRTEM images and the original scale bar. To determine the elemental composition of the TiO2 NW@Fe2O3 and the TiO2 NW@CuO nanocomposites, energy-dispersive X-ray spectroscopy (EDS; AMETEK Inc., Berwyn, PA, USA; active area 30 mm2) coupled to HRTEM was applied.
The crystal structure of the TiO2 NW@Fe2O3/cellulose and TiO2 NW@CuO/cellulose membranes was determined using X-ray powder diffraction (XRD) (Bruker D8 Advance diffractometer, Billerica, MA, USA) at (CuK α = 0.15418 nm; 40 kV and 40 mA) in parallel beam geometry (Göbel mirror) with a position-sensitive detector (Vantec1, Springfield, IA, USA; 1°opening). On top-loaded specimens in zero-background Si sample containers, measurements were made in the 2–80° (2 Theta) range using a 0.007° (2 Theta)/14 s goniometer speed.
The surface morphology of the as-prepared TiO2 NW@Fe2O3/cellulose and TiO2 NW@CuO/cellulose membranes was investigated by scanning electron microscopy. SEM and EDS spectroscopy was carried out in a Nova 600i Nanolab (Thermofisher, Eindhoven, The Netherlands) equipped with an EDS system for elemental analysis (EDAX Inc., Mahwah, NJ, USA). The EDS system mounted an Octane Elect Plus X-rays detector. Typical EDS maps and spectra were acquired using acceleration voltage values between 10 kV and 25 kV, with take-off angle of 35° and Dwell Time of 200 ms. For SEM analysis, the powders were deposited on a sticky carbon tape. Both powders and membranes were imaged directly without any conductive coating. The typical SEM working distance was 5 mm, and the acceleration voltage ranged from 2 kV up to 25 kV, depending on image quality and charging conditions.
To determine the surface area of the raw materials, the nanocomposites and the hybrid membranes, nitrogen adsorption–desorption experiments were carried out at 77 K to determine the Brunauer–Emmett–Teller [24] (BET) specific surface area using an ASAP 2020 instrument (Micromeritics Instrument Corp., Norcross, GA, USA).

3.6. Photocatalytic Experiments

The photocatalytic activity of the membranes was evaluated by studying the degradation of methylene blue as a model pollutant in an aqueous solution under UV light irradiation. The membranes were dipped into 100 mL of 0.03 mM MB solution and kept in the dark for two hours to attain the adsorption–desorption equilibrium. After this, the solution containing the pollutant and the photocatalyst (hybrid membranes) was exposed to UV-A lamps at a power between 300 and 500 W (Cosmedico N 400 R7S) for 60 min. The samples were withdrawn at regular time intervals and analyzed using a UV–Vis spectrophotometer (BEL UV-M51). The removal efficiency of the membranes was measured by recording the absorbance at 664 nm, and the degradation efficiency (% deg.) was calculated using the following formula:
%   d e g . = ( c 0 c t ) c 0 × 100
where, c0 is the initial concentration at time t = 0, ct is the concentration at time “t”, and % deg. is the photodegradation efficiency of the materials in relation to MB removal.

4. Conclusions

Herein, we presented the synthesis of two types of novel TiO2 nanowire-based hybrid membranes and studied their adsorption properties and photocatalytic performance in the decomposition of methylene blue under UV light. The as-prepared TiO2 NW@Fe2O3/cellulose and TiO2 NW@CuO/cellulose membranes were characterized by TEM, SEM, EDS and XRD. The results clearly demonstrated that not only the preparation of raw TiO2 NW, Fe2O3 and CuO decorated TiO2 NW, but also the production of hybrid membranes was successful.
Comparing the results of photocatalysis, it was found that both types of hybrid membranes showed outstanding performance in removing MB in only 60 min of UV irradiation. The photocatalytic degradation efficiency in MB removal of TiO2@Fe2O3/cellulose was 88%, while that of TiO2@CuO/cellulose membrane was up to 90%.
Since we applied UV light in this study, a significant degradation could presumably be observed under visible light irradiation; therefore, we intend to explore this possibility in a future study. Furthermore, we are planning to submit a further study in the near future regarding the microbiological and toxicological properties of the hybrid membranes here presented. It is believed that by exploiting the advantageous properties of the hybrid membrane-based water treatment technologies and solutions presented here, new and sustainable strategies could be implemented for photocatalyst-based water treatment technologies.

Author Contributions

Conceptualization, M.A.S. and N.S.; methodology, N.S. and Z.N.; validation, M.A.S., T.K., N.S. and A.K.L.; investigation, T.K., A.V., G.K., F.K. and N.S.; writing—original draft preparation, M.A.S., N.S., L.N. and Z.N.; writing—review and editing, A.V., A.K.L., K.H. and Z.N.; visualization, K.H. and Z.N.; supervision, K.H. and Z.N.; project administration, A.K.L. and Z.N.; funding acquisition, Z.N. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Research, Development and Innovation Office (NRDI Fund), grant number 2020-2.1.1-ED-2020-00029, Hungary; the experimental data used in this research were generated through access to the JRC Nanobiotechnology Laboratory under the Framework of access to the Joint Research Centre Physical Research Infrastructures of the European Commission (Virimmo project, Research Infrastructure Access Agreement N° 36171/11).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

A. K. Leskó was supported by the ÚNKP-21-4 New National Excellence Program of the Ministry for Innovation and Technology from the source of the National Research, Development and Innovation Fund. Z. Németh would like to thank the Hungarian Academy of Sciences Bolyai János Research Scholarship Program.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Siddique, K.; Rizwan, M.; Shahid, M.J.; Ali, S.; Ahmad, R.; Rizvi, H. Textile wastewater treatment options: A critical review. In Enhancing Cleanup of Environmental Pollutants; Springer International Publishing: Cham/Basel, Switzerland, 2017; pp. 183–207. [Google Scholar]
  2. di Paola, A.; García-López, E.; Marcì, G.; Palmisano, L. A survey of photocatalytic materials for environmental remediation. J. Hazard. Mater. 2012, 211–212, 3–29. [Google Scholar] [CrossRef] [PubMed]
  3. Duo, S.; Li, Y.; Liu, Z.; Zhong, R.; Liu, T.; Xu, H. Preparation of ZnO from 2D nanosheets to diverse 1D nanorods and their structure, surface area, photocurrent, optical and photocatalytic properties by simple hydrothermal synthesis. J. Alloys Compd. 2017, 695, 2563–2579. [Google Scholar] [CrossRef]
  4. Zheng, H.; Ou, J.Z.; Strano, M.S.; Kaner, R.B.; Mitchell, A.; Kalantar-zadeh, K. Nanostructured tungsten oxide-properties, synthesis, and applications. Adv. Funct. Mater. 2011, 21, 2175–2196. [Google Scholar] [CrossRef]
  5. He, R.A.; Cao, S.W.; Yu, J.G. Recent advances in morphology control and surface modification of Bi-based photocatalysts. Acta Phys.-Chim. Sin. 2016, 32, 2841–2870. [Google Scholar] [CrossRef]
  6. Dong, H.; Sun, J.; Chen, G.; Li, C.; Hu, Y.; Lv, C. An advanced Ag-based photocatalyst Ag2Ta4O11 with outstanding activity, durability and universality for removing organic dyes. Phys. Chem. Chem. Phys. 2014, 16, 23915–23921. [Google Scholar] [CrossRef]
  7. Ani, I.J.; Akpan, U.G.; Olutoye, M.A.; Hameed, B.H. Photocatalytic degradation of pollutants in petroleum refinery wastewater by TiO2- and ZnO-based photocatalysts: Recent development. J. Clean. Prod. 2018, 205, 930–954. [Google Scholar] [CrossRef]
  8. Dharma, H.N.C.; Jaafar, J.; Widiastuti, N.; Matsuyama, H.; Rajabsadeh, S.; Othman, M.H.D.; Rahman, M.A.; Jafri, N.N.M.; Suhaimin, N.S.; Nasir, A.M.; et al. A Review of Titanium Dioxide (TiO2)-Based Photocatalyst for Oilfield-Produced Water Treatment. Membranes 2022, 12, 345. [Google Scholar] [CrossRef]
  9. Xiang, L.; Zhao, X. Wet-chemical preparation of TiO2-based composites with different morphologies and photocatalytic properties. Nanomaterials 2017, 7, 310. [Google Scholar] [CrossRef] [Green Version]
  10. Mandal, R.K.; Kundu, S.; Sain, S.; Pradhan, S.K. Enhanced photocatalytic performance of V2O5-TiO2 nanocomposites synthesized by mechanical alloying with morphological hierarchy. New J. Chem. 2019, 43, 2804–2816. [Google Scholar] [CrossRef]
  11. Chakhtouna, H.; Benzeid, H.; Zari, N.; el kacem Qaiss, A.; Bouhfid, R. Recent progress on Ag/TiO2 photocatalysts: Photocatalytic and bactericidal behaviors. Environ. Sci. Pollut. Res. 2021, 28, 44638–44666. [Google Scholar] [CrossRef]
  12. Wang, L.; Zhao, J.; Liu, H.; Huang, J. Design, modification and application of semiconductor photocatalysts. J. Taiwan Inst. Chem. Eng. 2018, 93, 590–602. [Google Scholar] [CrossRef]
  13. Samokhvalov, A. Hydrogen by photocatalysis with nitrogen codoped titanium dioxide. Renew. Sustain. Energy Rev. 2017, 72, 981–1000. [Google Scholar] [CrossRef]
  14. Akhavan, O.; Ghaderi, E.; Rahimi, K. Adverse effects of graphene incorporated in TiO2 photocatalyst on minuscule animals under solar light irradiation. J. Mater. Chem. 2012, 22, 23260–23266. [Google Scholar] [CrossRef]
  15. Akhavan, O.; Ghaderi, E. Photocatalytic reduction of graphene oxide nanosheets on TiO2 thin film for photoinactivation of bacteria in solar light irradiation. J. Phys. Chem. C 2009, 113, 20214–20220. [Google Scholar] [CrossRef]
  16. Fua, J.; Wanga, X.; Mac, Z.; Wenmingd, H.; Lie, J.; Wanga, Z.; Wanga, L. Photocatalytic ultrafiltration membranes based on visible light responsive photocatalyst: A review. Desalin. Water Treat. 2019, 168, 42–55. [Google Scholar] [CrossRef]
  17. Assadi, M.H.N.; Hanaor, D.A.H. The effects of copper doping on photocatalytic activity at (101) planes of anatase TiO2: A theoretical study. Appl. Surf. Sci. 2016, 387, 682–689. [Google Scholar] [CrossRef] [Green Version]
  18. Endo-Kimura, M.; Karabiyik, B.; Wang, K.; Wei, Z.; Ohtani, B.; Markowska-Szczupak, A.; Kowalska, E. Vis-responsive copper-modified titania for decomposition of organic compounds and microorganisms. Catalysts 2020, 10, 1194. [Google Scholar] [CrossRef]
  19. Zhang, Q.; Rao, G.; Rogers, J.; Zhao, C.; Liu, L.; Li, Y. Novel anti-fouling Fe2O3/TiO2 nanowire membranes for humic acid removal from water. Chem. Eng. J. 2015, 271, 180–187. [Google Scholar] [CrossRef]
  20. Saber, O.; Kotb, H.M.; Osama, M.; Khater, H.A. An effective photocatalytic degradation of industrial pollutants through converting titanium oxide to magnetic nanotubes and hollow nanorods by kirkendall effect. Nanomaterials 2022, 12, 440. [Google Scholar] [CrossRef]
  21. Akhavan, O. Thickness dependent activity of nanostructured TiO2/α-Fe2O3 photocatalyst thin films. Appl. Surf. Sci. 2010, 257, 1724–1728. [Google Scholar] [CrossRef]
  22. Koe, W.S.; Lee, J.W.; Chong, W.C.; Pang, Y.L.; Sim, L.C. An overview of photocatalytic degradation: Photocatalysts, mechanisms, and development of photocatalytic membrane. Environ. Sci. Pollut. Res. 2020, 27, 2522–2565. [Google Scholar] [CrossRef] [PubMed]
  23. Tetteh, E.K.; Rathilal, S.; Asante-Sackey, D.; Chollom, M.N. Prospects of synthesized Magnetic TiO2-based membranes for wastewater treatment: A review. Materials 2021, 14, 3524. [Google Scholar] [CrossRef] [PubMed]
  24. Horváth, E.; Rossi, L.; Mercier, C.; Lehmann, C.; Sienkiewicz, A.; Forró, L. Photocatalytic nanowires-based air filter: Towards reusable protective masks. Adv. Funct. Mater. 2020, 30, 2004615. [Google Scholar] [CrossRef] [PubMed]
  25. Roy, P.; Berger, S.; Schmuki, P. TiO2 nanotubes: Synthesis applications. Angew. Chem. Int. Ed. 2011, 50, 2904–2939. [Google Scholar] [CrossRef] [PubMed]
  26. Gao, B.; Sun, M.; Ding, W.; Ding, Z.; Liu, W. Decoration of γ-graphyne on TiO2 nanotube arrays: Improved photoelectrochemical and photoelectrocatalytic properties. Appl. Catal. B Environ. 2021, 281, 119492. [Google Scholar] [CrossRef]
  27. Gyulavári, T.; Kovács, K.; Kovács, Z.; Bárdos, E.; Kovács, G.; Baán, K.; Magyari, K.; Veréb, G.; Pap, Z.; Hernadi, K. Preparation and characterization of noble metal modified titanium dioxide hollow spheres new insights concerning the light trapping efficiency. Appl. Surf. Sci. 2020, 534, 147327. [Google Scholar] [CrossRef]
  28. Chakraborty, S.; Loutatidou, S.; Palmisano, G.; Kujawa, J.; Mavukkandy, M.O.; Al-Gharabli, S.; Arafat, H.A. Photocatalytic hollow fiber membranes for the degradation of pharmaceutical compounds in wastewater. J. Environ. Chem. Eng. 2017, 5, 5014–5024. [Google Scholar] [CrossRef]
  29. Rosales, M.; Zoltan, T.; Yadarola, C.; Mosquera, E.; Gracia, F.; García, A. The influence of the morphology of 1D TiO2 nanostructures on photogeneration of reactive oxygen species and enhanced photocatalytic activity. J. Mol. Liq. 2019, 281, 59–69. [Google Scholar] [CrossRef]
  30. Nakata, K.; Fujishima, A. TiO2 photocatalysis: Design and applications. J. Photochem. Photobiol. C Photochem. Rev. 2012, 13, 169–189. [Google Scholar] [CrossRef]
  31. Jitputti, J.; Suzuki, Y.; Yoshikawa, S. Synthesis of TiO2 nanowires and their photocatalytic activity for hydrogen evolution. Catal. Commun. 2008, 9, 1265–1271. [Google Scholar] [CrossRef]
  32. Mamaghani, A.H.; Haghighat, F.; Lee, C.-S. Hydrothermal/solvothermal synthesis and treatment of TiO2 for photocatalytic degradation of air pollutants: Preparation, characterization, properties, and performance. Chemosphere 2019, 219, 804–825. [Google Scholar] [CrossRef]
  33. Al-Hajji, L.A.; Ismail, A.A.; Al-Hazza, A.; Ahmed, S.A.; Alsaidi, M.; Almutawa, F.; Bumajdad, A. Impact of calcination of hydrothermally synthesized TiO2 nanowires on their photocatalytic efficiency. J. Mol. Struct. 2020, 1200, 127153. [Google Scholar] [CrossRef]
  34. Shi, Y.; Huang, J.; Zeng, G.; Cheng, W.; Hu, J. Photocatalytic membrane in water purification: Is it stepping closer to be driven by visible light? J. Membr. Sci. 2019, 584, 364–392. [Google Scholar] [CrossRef]
  35. Leong, S.; Razmjou, A.; Wang, K.; Hapgood, K.; Zhang, X.; Wang, H. TiO2 based photocatalytic membranes: A. review. J. Membr. Sci. 2014, 472, 167–184. [Google Scholar] [CrossRef]
  36. Madhura, L.; Kanchi, S.; Sabela, M.I.; Singh, S.; Bisetty, K. Inamuddin. Membrane technology for water purification. Environ. Chem. Lett. 2018, 16, 2343–2365. [Google Scholar] [CrossRef]
  37. Ezugbe, E.O.; Rathilal, S. Membrane technologies in wastewater treatment: A Review. Membranes 2020, 10, 89. [Google Scholar] [CrossRef] [PubMed]
  38. Lu, W.; Duan, C.; Liu, C.; Zhang, Y.; Meng, X.; Dai, L.; Wang, W.; Yu, H.; Ni, Y. A self-cleaning and photocatalytic cellulose-fiber- supported “Ag@AgCl@MOF- cloth’’ membrane for complex wastewater remediation. Carbohydr. Polym. 2020, 247, 116691. [Google Scholar] [CrossRef]
  39. Azmi, A.; Lau, K.S.; Chin, S.X.; Zakaria, S.; Chia, C.H. Regenerated cellulose membrane incorporating photocatalytic zinc oxide as a bifunctional membrane for decoloration of methylene blue. Polym. Adv. Technol. 2022, 33, 3843–3850. [Google Scholar] [CrossRef]
  40. Rabiee, N.; Fatahi, Y.; Asadnia, M.; Daneshgar, H.; Kiani, M.; Ghadiri, A.M.; Amin, M.A.; Mashhadzadeh, H.; Akhavan, O.; Bagherzadeh, M.; et al. Green porous benzamide-like nanomembranes for hazardous cations detection, separation, and concentration adjustment. J. Hazard. Mater. 2022, 423, 127130. [Google Scholar] [CrossRef]
  41. Xu, C.; Xu, Y.; Zhu, J. Photocatalytic antifouling graphene oxide-mediated hierarchical filtration membranes with potential applications on water purification. ACS Appl. Mater. Interfaces 2014, 6, 16117–16123. [Google Scholar] [CrossRef]
  42. Bai, H.; Zan, X.; Zhang, L.; Sun, D.D. Multi-functional CNT/ZnO/TiO2 nanocomposite membrane for concurrent filtration and photocatalytic degradation. Sep. Purif. Technol. 2015, 156, 922–930. [Google Scholar] [CrossRef]
  43. Athanasekou, C.P.; Moustakas, N.G.; Morales-Torres, S.; Pastrana-Martínez, L.M.; Figueiredo, J.L.; Faria, J.L.; Silva, A.; Rodríguez, J.M.D.; Romanos, G.E.; Falaras, P. Ceramic photocatalytic membranes for water filtration under UV and visible light. Appl. Catal. B Environ. 2015, 178, 12–19. [Google Scholar] [CrossRef] [Green Version]
  44. Jamshidifard, S.; Koushkbaghi, S.; Hosseini, S.; Rezaei, S.; Karamipour, A.; Irani, M. Incorporation of UiO-66-NH2 MOF into the PAN/chitosan nanofibers for adsorption and membrane filtration of Pb(II), Cd(II) and Cr(VI) ions from aqueous solutions. J. Hazard. Mater. 2019, 368, 10–20. [Google Scholar] [CrossRef] [PubMed]
  45. Santos, É.N.; László, Z.; Hodúr, C.; Arthanareeswaran, G.; Veréb, G. Photocatalytic membrane filtration and its advantages over conventional approaches in the treatment of oily wastewater: A review. Asia-Pac. J. Chem. Eng. 2020, 15, 5. [Google Scholar] [CrossRef]
  46. Hu, A.; Zhang, X.; Oakes, K.D.; Pemg, P.; Zhou, Y.N.; Servos, M.R. Hydrothermal growth of free standing TiO2 nanowire membranes for photocatalytic degradation of pharmaceuticals. J. Hazard. Mater. 2011, 189, 278–285. [Google Scholar] [CrossRef]
  47. Garcia, R.A.; Stevanovic, T.; Berthier, J.; Njamen, G.; Tolnai, B.; Achim, A. Cellulose, nanocellulose, and antimicrobial materials for the manufacture of disposable face masks: A review. BioResources 2021, 16, 4321–4353. [Google Scholar] [CrossRef]
  48. Peng, B.; Yao, Z.; Wang, X.; Crombeen, M.; Sweeney, D.G.; Tam, K.C. Cellulose-based materials in wastewater treatment of petroleum industry. Green Energy Environ. 2020, 5, 137–149. [Google Scholar] [CrossRef]
  49. Yang, J.; Yu, J.; Fan, J.; Sun, D.; Tang, W.; Yang, X. Biotemplated preparation of CdS nanoparticles/bacterial cellulose hybrid nanofibers for photocatalysis application. J. Hazard. Mater. 2011, 189, 377–383. [Google Scholar] [CrossRef]
  50. Sun, D.; Yang, J.; Wang, X. Bacterial cellulose/TiO2 hybrid nanofibers prepared by the surface hydrolysis method with molecular precision. Nanoscale 2010, 2, 287–292. [Google Scholar] [CrossRef]
  51. Liu, G.Q.; Pan, X.J.; Li, J.; Li, C.; Ji, C.L. Facile preparation and characterization of anatase TiO2/nanocellulose composite for photocatalytic degradation of methyl orange. J. Saudi Chem. Soc. 2021, 25, 101383. [Google Scholar] [CrossRef]
  52. Cao, X.; Huang, M.; Ding, B.; Yu, J.; Sun, G. Robust polyacrylonitrile nanofibrous membrane reinforced with jute cellulose nanowhiskers for water purification. Desalination 2013, 316, 120–126. [Google Scholar] [CrossRef]
  53. Li, Y.; Yu, H.; Yang, Y.; Zheng, F.; Ni, H.; Zhang, M.; Guo, M. Synthesis of potassium hexatitanate whiskers with high thermal stability from Ti-bearing electric arc furnace molten slag. Ceram. Int. 2016, 42, 11294–11302. [Google Scholar] [CrossRef]
  54. Suresh, S.; Karthikeyan, S.; Jayamoorthy, K. FTIR and multivariate analysis to study the effect of bulk and nano copper oxide on peanut plant leaves. J. Sci. Adv. Mater. Devices 2016, 1, 343–350. [Google Scholar] [CrossRef] [Green Version]
  55. Cong, Y.; Li, Z.; Zhang, Y.; Wang, Q.; Xu, Q. Synthesis of α-Fe2O3/TiO2 nanotube arrays for photoelectro-fenton degradation of phenol. Chem. Eng. J. 2012, 191, 356–363. [Google Scholar] [CrossRef]
  56. Shehab, M.A.; Sharma, N.; Karacs, G.; Kristaly, F.; Mrabate, B.E.; Chauhan, T.; Koós, T.; Leskó, A.K.; Palotás, Á.B.; Hernádi, K.; et al. Effect of the synthesis parameters on the formation of TiO2 nanostructures: Controllable synthesis and adsorption properties of nanowires and nanotubes. Circ. Econ. Environ. Prot. 2021, 5, 16–38. [Google Scholar]
Figure 1. HRTEM images of TiO2 NW@Fe2O3 (a) and TiO2 NW@CuO (b) nanocomposite samples.
Figure 1. HRTEM images of TiO2 NW@Fe2O3 (a) and TiO2 NW@CuO (b) nanocomposite samples.
Molecules 27 02951 g001
Figure 2. EDS spectra of TiO2 NW@Fe2O3 (a) and TiO2 NW@CuO (b) nanocomposites.
Figure 2. EDS spectra of TiO2 NW@Fe2O3 (a) and TiO2 NW@CuO (b) nanocomposites.
Molecules 27 02951 g002
Figure 3. SEM images of neat cellulose (a), TiO2 NW (b); photograph and SEM images of TiO2 NW@Fe2O3 (c,d) and TiO2 NW@CuO (e,f) nanocomposite samples.
Figure 3. SEM images of neat cellulose (a), TiO2 NW (b); photograph and SEM images of TiO2 NW@Fe2O3 (c,d) and TiO2 NW@CuO (e,f) nanocomposite samples.
Molecules 27 02951 g003
Figure 4. XRD of TiO2 NW@Fe2O3/cellulose (a) and TiO2 NW@CuO/cellulose (b) hybrid membranes.
Figure 4. XRD of TiO2 NW@Fe2O3/cellulose (a) and TiO2 NW@CuO/cellulose (b) hybrid membranes.
Molecules 27 02951 g004
Figure 5. Photocatalytic performance of TiO2 NW@Fe2O3/cellulose (blue curve) and TiO2 NW@CuO/cellulose (red curve) hybrid membranes in MB degradation.
Figure 5. Photocatalytic performance of TiO2 NW@Fe2O3/cellulose (blue curve) and TiO2 NW@CuO/cellulose (red curve) hybrid membranes in MB degradation.
Molecules 27 02951 g005
Figure 6. Adsorption capacity of pure cellulose (green marks) as well as TiO2 NW@Fe2O3/cellulose (blue marks) and TiO2 NW@CuO/cellulose (red marks) hybrid membranes in MB adsorption tests.
Figure 6. Adsorption capacity of pure cellulose (green marks) as well as TiO2 NW@Fe2O3/cellulose (blue marks) and TiO2 NW@CuO/cellulose (red marks) hybrid membranes in MB adsorption tests.
Molecules 27 02951 g006
Table 1. Elemental (EDS) analysis of the raw materials (cellulose and TiO2 NW) and hybrid membranes.
Table 1. Elemental (EDS) analysis of the raw materials (cellulose and TiO2 NW) and hybrid membranes.
Sample NameCOTiNaFKFeCu
Cellulose4555------
TiO2 NW-6436-----
TiO2 NW@Fe2O3/cellulose2050233-31-
TiO2 NW@CuO/cellulose255811-32-1
Table 2. Specific surface areas of the pure materials and the hybrid membranes.
Table 2. Specific surface areas of the pure materials and the hybrid membranes.
Sample NameSurface Area (m2/g)Pore Diameter (nm)
TiO2 NW168-
cellulose membrane618
Fe2O362-
CuO141-
TiO2 NW@Fe2O3146-
TiO2 NW@CuO139-
TiO2 NW@Fe2O3/cellulose12216
TiO2 NW@CuO/cellulose11716
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Shehab, M.A.; Sharma, N.; Valsesia, A.; Karacs, G.; Kristály, F.; Koós, T.; Leskó, A.K.; Nánai, L.; Hernadi, K.; Németh, Z. Preparation and Photocatalytic Performance of TiO2 Nanowire-Based Self-Supported Hybrid Membranes. Molecules 2022, 27, 2951. https://doi.org/10.3390/molecules27092951

AMA Style

Shehab MA, Sharma N, Valsesia A, Karacs G, Kristály F, Koós T, Leskó AK, Nánai L, Hernadi K, Németh Z. Preparation and Photocatalytic Performance of TiO2 Nanowire-Based Self-Supported Hybrid Membranes. Molecules. 2022; 27(9):2951. https://doi.org/10.3390/molecules27092951

Chicago/Turabian Style

Shehab, Mohammed Ahmed, Nikita Sharma, Andrea Valsesia, Gábor Karacs, Ferenc Kristály, Tamás Koós, Anett Katalin Leskó, Lilla Nánai, Klara Hernadi, and Zoltán Németh. 2022. "Preparation and Photocatalytic Performance of TiO2 Nanowire-Based Self-Supported Hybrid Membranes" Molecules 27, no. 9: 2951. https://doi.org/10.3390/molecules27092951

Article Metrics

Back to TopTop