Next Article in Journal
Mathematical Approach to Optimizing the Panchromatic Absorption of Natural Dye Combinations for Dye-Sensitized Solar Cells
Previous Article in Journal
pH-Induced Orthogonal Photoresponse of trans-Chalcone Isomers and Related Compounds in Equilibria
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Crosslinking and Porosity Surface Effects of Photoetching Process on Immobilized Polymer-Based Titanium Dioxide for the Decolorization of Anionic Dye

by
Siti Raihan Hamzah
,
Muhammad Afiq Rosli
,
Nadiah Sabihah Natar
,
Nureel Imanina Abdul Ghani
,
Nur Aien Muhamad
,
Mohammad Saifulddin Azami
,
Mohd Azlan Mohd Ishak
,
Razif Nordin
and
Wan Izhan Nawawi
*
Faculty of Applied Sciences, Universiti Teknologi MARA (UiTM), Cawangan Perlis, Arau 02600, Perlis, Malaysia
*
Author to whom correspondence should be addressed.
Colorants 2023, 2(1), 73-89; https://doi.org/10.3390/colorants2010006
Submission received: 9 December 2022 / Revised: 14 January 2023 / Accepted: 16 January 2023 / Published: 17 February 2023

Abstract

:
The textile industry is suffering a great challenge regarding wastewater management, primarily due to the implementation of improper systems, specifically for dye wastewater treatment. Photocatalysis is one of approaches that have been used to treat wastewater. Titanium dioxide (TiO2) was immobilized by using the dip-coating technique in this research. Epoxidized natural rubber (ENR) and polyvinyl chloride (PVC) were used as a polymer to bind the TiO2 on the glass substrate. This immobilized TiO2/ENR/PVC underwent a photoetching process at various times to study the crosslink and porosity formations. Reactive red 4 dye was used as a model pollutant for photocatalytic performance. All immobilized TiO2/ENR/PVC samples under 12, 24 and 30 h of photoetching process (TEP12, TEP24 and TEP30 samples, respectively) showed higher photocatalytic activity compared to those without photoetching process (TEP0 sample) due to the intermediate charge in crosslinking reaction after the photoetching process. The TEP24 sample showed the highest photocatalytic degradation; light harvesting; photocatalytic degradation.

1. Introduction

Photocatalysis is one of the promising techniques among the various AOPs [1]. The most famous semiconductor photocatalyst is titanium dioxide (TiO2) due to its low cost, high stability and non-toxicity compared to other semiconductors [2]. Recently, a normal practice of TiO2 for treating wastewater is in suspension mode as it has [3,4,5]. However, this practice gives rise to a difficulty in the post-treatment filtration to separate the treated wastewater from TiO2 [6,7,8,9,10]. The post-treatment process can be overcome by developing an immobilized TiO2 that can be recycled and continuously used. There are several methods to immobilize the TiO2 onto a solid substrate, i.e., dip coating from suspension [9,10,11,12,13], electrophoretic deposition [14,15,16], sol-gel related methods [17,18,19] and spray coating and sputtering [20,21,22]. Although the application of immobilized TiO2 has gained much attention in order to solve the problem of the post-treatment process, the low surface area of TiO2 is also considered to cause the decrease in the efficiency of photocatalytic activity [23]. Based on Chen et al. [24], the highest photocatalytic activity is due to the high availability of the active surface area for organic compound absorption. Consequently, the high surface area also can be created by introducing immobilized photocatalysts that can produce a high porosity of the catalyst.
Previously, Tennakone et al. [25] discovered the first technique of immobilization of TiO2 using a polymer as a binder. Their finding has led other researchers to study various types of polymers in immobilizing the TiO2. The addition of a polymer as a binder in immobilized TiO2 is usually applied to improve its adhesion, temperature resisting ability, and durability towards pollutants [26,27,28], and some polymers are also able to enhance TiO2 photoactivity [29,30,31,32,33,34,35,36,37,38]. Generally, solvent-based polymers are used in preparing immobilized TiO2. There are a few polymers that have been used, such as polyethylene sheets [30], thin polyethylene films, polystyrene (PS) beads [31], expanded polystyrene (EPS) beads, ethyl cellulose microspheres [32], fluoropolymer resins, polyethylene terephthalate (PET) bottles [33], polypropylene (PP) granules, cellulose fibers [34], polypropylene fiber (PPF), polyvinyl chloride (PVC) [35], polycarbonate (PC), poly(methyl methacrylate) (PMMA), polyvinyl acetate (PVAc) [36], poly(styrene)-co-poly(4-vinylpyridine) (PSP4VP), rubber latex (an elastic hydrocarbon polymer) [37], parylene and tedlar [38]. These binders have improved the performance of immobilized TiO2 in their own unique characteristics.
Many researchers [7,8,9,39] have studied immobilizing the TiO2 by using epoxidized natural rubber (ENR-50) and polyvinyl chloride (PVC) as a polymer binder due to their properties of adhesiveness and durability. They state that the ENR compound was photoetched out from the immobilized TiO2 and crosslinking reaction occurred during the photoetching process. However, no study reported the chemical reaction pathway process and the end product formed in the crosslinking reaction during the photoetching process. Hence, this paper focuses on the surface behavior and the photocatalytic activity of immobilized TiO2/ENR/PVC after the photoetching process, suggests a possible reaction mechanism, and presents a crosslinked ENR/PVC end product achieved from the photoetching process.

2. Materials and Methods

The titanium dioxide used was commercial Degussa P-25 TiO2 (80 anatase, 20% rutile) nano powder procured from Acros Organics, Geel, Belgium. The polymeric binders applied for the immobilized TiO2 were ENR-50 and PVC supplied by Kumpulan Guthrie Sdn. Bhd., Kuala Lumpur, Malaysia and Petrochemicals (M) Sdn. Bhd., Johor, Malaysia respectively. Toluene, C7H8 (99.5%) and dichloromethane (DCM), CH2Cl2 (99.5%) were purchased from R&M Chemicals, Selangor, Malaysia. The model pollutant was reactive red 4 (RR4) dye or known as Cibacron Brilliant Red with 50% dye content (Sigma-Aldrich (M) Sdn. Bhd., Petaling Jaya, Malaysia), as it is widely used in the industry. Throughout this work, ultra-pure water (18.2 MΩ cm−1) was used to prepare all solutions and dilution purposes.
The reflux process of ENR-50 was carried out by using a stirring heating mantle from Fisher Scientific, Gelugor, Penang. The formulation of immobilized TiO2 was homogenized via Crest Ultrasonic cleaner model 4NT-1014-6 (50–60 kHz) from Crest Systems (M) Sdn Bhd, Bayan Lepas, Penang. A 55 W and 65 W compact household fluorescent lamp (Firefly Electric & Lighting Corp., Manila, Philipines) with an emitted UV leakage irradiance of 6.0 W m−2 and aquarium pump model NS 7200 (Ace Story Aquatic, Penang, Malaysia) were used as the light and aeration sources, respectively. Two custom-made glass cells with of length 58 mm, width 10 mm and height 80 mm were used for all photocatalytic processes. The support material used for immobilizing the TiO2 was a glass plate (dimensions of 47 mm × 70 mm) with ground surfaces on one side. The concentration of RR4 was determined by using an ultraviolet-visible (UV-vis) spectrophotometer, model DR2000 from HACH Malaysia Sdn. Bhd., Kuala Lumpur, Malaysia. The surface morphology analysis was characterized by a scanning electron microscope (SEM) analyzer model LEO SUPRA 50 VP field emission SEM (Bruker Malaysia Sdn. Bhd., Penang, Malaysia) and 3D optical profilometer (PEMTRON HAWK 3D WT-250, Pemtron, Seoul, South Korea). A Fourier transform infrared spectroscopy (FTIR) analyzer (Perkin-Elmer, model system 2000 FTIR) from Perkin Elmer Sdn. Bhd., Selangor, Malaysia was used to study the changes of functional groups in the treated samples. The binding energy of the chemical compound was determined by performing X-ray photoelectron spectroscopy (XPS) with an Argus detector using an Omicron’s DAR 400 dual Mg/Al X-ray source (Thermo Fisher Scientific Malaysia, Selangor, Malaysia). X-ray diffraction (XRD) analysis of the prepared samples was performed by using a Bruker D8 Advance diffractometer (Bruker Malaysia Sdn. Bhd., Penang, Malaysia).
The ENR solution was prepared by refluxing 24.8 ± 0.05 g of ENR-50 in 250 mL of toluene at 88–90 °C until all of the ENR-50 was completely dissolved in the toluene and formed an adhesive-like and sticky solution. The preparation of the PVC solution was carried out by dissolving 0.8 g of PVC powder in 35 mL of dichloromethane through sonication for 1 h. Both of these solutions were used to prepare the immobilized TiO2 formulation in which 6.0 g of TiO2 was added slowly into the ENR-50/PVC blend with a mixed ratio of (5:1). Finally, the TiO2 formulation was sonicated by using an ultrasonic machine to completely homogenize the mixture.
A simple dip coating method was applied for immobilizing the TiO2 formulation onto clean glass plates (47 mm × 70 mm, 2 mm thickness). Before the dip coating process, each of the glass plates was dried in the oven at 100 °C for 30 min and weighed by using an analytical balance (Denver Instrument Company, model AA-160) from Alpha Chemicals Sdn. Bhd., Kuala Lumpur, Malaysia. The homogenized TiO2 formulation was poured into a coating glass cell. The glass plate was dipped (cover up to 5.6 cm depth) for 5 s in the formulation and subsequently pulled up manually with a consistent pulling rate. Then, the coated glass plate was dried completely by using an air blower to vaporize the solvent. The weight of the coated glass plate was obtained after scraping off the smooth side of the glass plate. The weight difference between the blank glass plate and coated plate would determine the weight of the immobilized TiO2 and organic binder deposited onto the glass plate. The procedure of coating, drying and weighing was carried out until 2 g of the TiO2 loading was obtained, which typically included three rounds of dip coating. The TiO2 particle loading in the coating was estimated to be around 1.10 to 1.20 g based on the ratio TiO2:ENR:PVC (6:4:1). All of the coated plates were completely dried and stored in plastic dishes before use.
A customized glass cell filled with distilled water (Universiti Teknologi MARA, Perlis, Malaysia) was used to immerse the coated glass plate with aeration air. A 65 W compact fluorescent lamp (Firefly Electric & Lighting Corp., Manila, Philipines) was switched on to initiate the crosslinking reaction. The coated glass plate was photoetched for 1 h. Subsequently, the photoetched coated glass plate was used for the photocatalytic study. The difference of photocatalytic degradation performance between the non-photoetched and photoetched coated glass plates was finally observed.
A 55 W compact fluorescent lamp (Firefly Electric & Lighting Corp., Manila, Philipines) was placed horizontally in front of the custom-made glass cell reactor. A volume of 16 mL of 30 ppm RR4 dye solution was poured into the glass cell. Then, the glass plate was immersed into the glass cell with aeration and a light source. At 15 min time intervals, the degradation of the dye during the photocatalysis process was observed and the absorbance was measured by using a DR2000 spectrophotometer (HACH Malaysia Sdn. Bhd., Kuala Lumpur, Malaysia). The results were converted into ln C0/C, where C0 is the absorbance of the initial concentration and C is the absorbance at any time (t), and thus the values were plotted against irradiation or contact time. Based on the Langmuir–Hinshelwood rate model, the slope of the linear line was taken as the pseudo first-order rate constant.

3. Results

The surface morphology as seen via field scanning electron microscopy (FESEM) of TEP0, TEP12, TEP24 and TEP30 for immobilized TiO2/ENR/PVC without photoetching and with 12, 24 and 30 h photoetching, respectively, is shown in Figure 1. The morphology of immobilized TiO2/ENR/PVC without photoetching in Figure 1a,b shows a compact surface, which indicates that the TiO2 particles are covered with a layer of ENR/PVC. Figure 1c,d illustrates the immobilized TiO2/ENR/PVC after 12 h of photoetching, which reveals that the fully covered layer is being degraded and turning into a hole with a sponge-like layer structure after a 12 h photoetching process. The ENR/PVC layer diminished from the TiO2 after the 24 h photoetching process (Figure 1e,f) and remained constant until 30 h of the photoetching process (Figure 1g,h). The amounts of ENR/PVC layer and TiO2 that leached out in the water during the photoetching were ca. 0.08 and 0.096 g, respectively. Based on previous researchers [7,8,9,40,41], they state that a porous immobilized TiO2 is produced as the ENR-50 leaches out during the 8 to 12 h photoetching process. This finding corresponds with the result obtained and shown in Figure 1a,h, as the layer of ENR/PVC diminished slowly upon time.
The porosity transformations in TEP0 and TEP24 are supported by surface topography characterization via 3D optical profilometer (PEMTRON HAWK 3D WT-250, Pemtron, Seoul, South Korea) using an optical probe at 5× magnification, as shown in Figure 2a,d. The height of the surface can be determined by the color indicator shown in the 2D surface profile (Figure 2a,b). Based on the color indicator, the blue color indicates the deepest height while the red color depicts the tallest height. The 3D surface profile illustrates that the surface of TEP24 was more porous compared to TEP0.
The parameters in the 2D surface roughness category are average roughness (Ra) and familiar root mean square roughness (Rq). Both of these parameters are given by the following equations [42]:
R a = ( 1 / N )   J = 1 N Z J
where N is the total number of measured points and ZJ is the vertical height deviation from the average height of the surface.
R q =   ( Z J ) 2 /   N
where ZJ is the vertical deviation with respect to a mean line for the jth point of the total N measured points.
The Ra values for TEP0 and TEP24 are 8101.95 and 15,674.81 nm, while the Rq values are 10,449.51 and 19,062.62 nm, respectively. The surface roughness of TEP24 is higher than TEP0 due to the high porosity as ENR-50 leaches out from the immobilized TiO2 as signified by the FESEM results in Figure 1.
The XRD analysis for the TEP0 and TEP24 samples was performed on a Bruker D8 advance diffractometer (DKHS Sdn. Bhd., Petaling Jaya, Malaysia) operating in the reflection mode with Cu-Kα radiation (35 kV, 30 mA) and a diffracted beam monochromator (DKHS Sdn. Bhd., Petaling Jaya, Malaysia), using a step scan mode with the step of 0.075° (2θ) and 4 s per step. The XRD patterns of TEP0 and TEP24 are shown in Figure 3. All peaks were detected as anatase phase compared with JCPDS Card No: 01-071-1169. The positions of 2θ at 25.31°, 26.95°, 37.79°, 38.57°, 48.04°, 53.89°, 55.07°, 68.76°, and 70.30° correspond to the Miller indices of (101), (103), (004), (112), (200), (105), (211), (116), (220), and (215) plane, respectively. Assuming that the crystalline powder did not align itself to a preferential orientation during fixation, the strong diffraction peaks at 25.31° and 48.04° confirm the samples in the anatase structure. The higher peak intensities are due to the etching of the polymer, exposing more TiO2 crystals to the X-ray probe. The morphology of the ceramic powder is unlikely to change under the reported experimental condition, other than some electrochemical etching.
Table 1 shows that the crystallinity index (CI) values of TEP0 and TEP24 are 51.21% and 57.30%, respectively. The CI value for TEP24 is 6.07%, which is higher than the TEP0 sample due to the ENR/PVC decrement, as also observed in the FESEM images. Based on the XRD pattern, the crystallinity index (%) can be calculated via XRD data and origin software using the following equation [43,44,45]:
Crystallinity index (CI) = Area CP/Area CAP
where area CP is the area of all the crystalline peaks while area CAP is the area of all the crystalline and amorphous peaks.
Figure 4 depicts an optimized slab model of the TiO2 structure that has been calculated before the calculations of the electronic structures. The slab model was executed to acquire the lattice parameters with the lowest energy. The computational results for the structural parameters such as lattice parameter, volume, and bond length are shown in Table 2. Based on the data obtained, the structural TiO2 can be reported as anatase form due to its tetragonal shape. The volume of the TiO2 between TEP0 and TEP24 has no significant difference, which indicates that there is no structural alteration during the photoetching process.
The electronic structures of TEP0 and TEP24 were investigated by setting the same k-points mesh to sample the first Brillouin zone for TEP0 and TEP24. Figure 5a,b represents the calculated band gaps of TEP0 and TEP24, which are 3.206 and 3.210 eV, respectively. The conduction band minimum (CBM) is located at M, while the valence band maximum (VBM) is located near G. The band gap was not affected by the photoetching treatment, as the band gap values of both samples are obscure to the theoretical band gap of TiO2, which is 3.20 eV. Nonetheless, the calculated band gaps are overvalued compared with the experimental value obtained by UV-Vis DRS analysis as shown in Figure 6b. This may be due to the DFT limitation, as the discontinuity in the exchange correlation is not taken into account within the framework of DFT.
Based on Figure 6a, TEP0 exhibits higher absorbance of the maximum peak (λmax) compared to TEP24. The difference in UV absorption intensity observed is also a consequence of the exposed TiO2 crystals after the binder was photo-etched with the very remote possibility that there might have been some preferential orientation of the crystals during curing. The band gap energy (Eg) of the samples was calculated by the Tauc equation αhν = A(hν − Eg)n, where A is constant, and α, h, ν, and Eg are the absorption coefficient, Planck’s constant, photon energy, and band gap, respectively [46]. n is defined by the optical transition of a semiconductor (n = 1 for TiO2). As shown in Figure 6b, the Eg for TEP0 and TEP24 is approximately 3.18 and 3.00 eV, respectively. The Eg for TEP24 is lower than the TEP0 supporting the reduction of electron hole recombination incidence. Thus, the smaller electron hole recombination after the photoetching process results in a higher photocatalytic degradation.
The total densities of states (TDOS) and partial density of states (PDOS) of TEP0 and TEP24 treated by Gaussian broadening are verified in Figure 7a,b. The band gap can be defined by the separation between the CBM and VBM, which are extending in energy ranges from 0 to 6 eV and from −6 to 0 eV, respectively. The DOS for both samples shows that the VBM is predominantly made up of O 2p states while the CBM is mostly made up of Ti 3d states alongside a minimal amount of O 2p states. These results are consistent with other DFT calculations on TiO2 polymorphs [47].
The FTIR spectra of TEP0, TEP24, and Degussa P-25 TiO2 are shown in Figure 8. The peaks at 1251, 1332, and 1432 cm−1 in Figure 8 indicate the presence of PVC, while the peak at 1432 cm−1 indicates the presence of the ENR-50 compound [7,8,9,41]. The destruction of the CH2Cl deformation in PVC compounds and the reduction of ENR-50 compounds can be seen by the decrease of peak intensity at 1251 and 1432 cm−1, respectively. This is due to the opening ring of ENR-50 and crosslinking reaction between ENR-50 and PVC during the photoetching process as pointed out by [7,8,9]. Moreover, there are two new peaks of the C–O bond in the TEP24 spectrum at 1167 and 1062 cm−1, which confirms the formation of aliphatic ether due to the crosslinking reaction. Hence, even though the porosity formed due to the leaching out of ENR, there is some of the remaining ENR turning into a crosslinking reaction as shown in the possible reaction mechanism as depicted in Figure 10. Instead of a crosslinking reaction occurring, there is some potential that the PVC may be converted into a polyenes precursor, which can serve as a photo-sensitizer, thus enhancing the photocatalytic activity of TiO2 [8].
X-ray photoelectron spectroscopy (XPS) was performed to confirm the binding energy of the elements detected in TEP24. The crosslinking reaction was successfully produced by the photoetching process. Figure 9a,b demonstrates the wide, C1s, O1s, and Cl2p spectra for TEP24. There are four elements in the samples: Ti, C, O, and Cl. Figure 9a shows the wide spectra elements of the sample while Figure 9b represents the C1s spectra in which the C–C and C–O are located at 284.5 and 286.8 eV, respectively. Figure 9c illustrates the O1s characteristic peaks of Ti–O, C–O–C (aliphatic), and epoxide at 529.3, 532.5, and 533.1 eV respectively. The appearance of the peak of C–O–C (aliphatic) verified the presence of a crosslink reaction. This is due to the ether crosslink product consisting of C–O–C (aliphatic). Lastly, Figure 9d displays the Cl2p characteristic peaks that indicate the existence of chloride from the PVC. Thus, it has been confirmed that the crosslink effect occurred after the photoetching process due to the presence of C–O–C (aliphatic).
Based on the structure of ENR-50 in Figure 10, ENR-50 consists of an oxirane ring that is highly strained and readily open under mild conditions [48]. After the photoetching process, the oxirane ring in the ENR-50 structure was altered by a ring-opening reaction that formed radicals, as illustrated in the mechanism proposed in Figure 10. There are two possibilities of the ENR-50 crosslinking reaction as displayed in Figure 10(1,2). The opened ring of ENR-50 tends to react with PVC or another ENR-50, thus producing an ether crosslink product. The products derived from the crosslinking reaction may increase the durability and the adhesiveness of the immobilized TiO2. Moreover, the radicals formed at the intermediate reaction (opening ring of ENR-50) can react with the RR4 dye (model pollutant) during the photocatalytic degradation as light penetrates towards the immobilized TiO2/ENR/PVC, thus enhancing the photocatalytic degradation rate.
The photoluminescence (PL) spectra of TEP0 and TEP24 were extracted to confirm the photoexcitation in the samples as shown in Figure 11. Basically, the PL intensity of TEP24 is quite higher than that of TEP0 at the sub-band range, which is around 500–700 nm wavelength. This situation has been expressed in the occurrence of electron hole recombination from the band to sub-band TiO2 due to the superficial oxygen vacancies and flaws of the semiconductors. The photocatalytic activity of the TEP24 demonstrates a sophisticated response despite the high electron hole recombination since the photoetching process has created open pores through the internal sample, thus increasing the activation of the TiO2 site.

Photodegradation of Immobilized TiO2/ENR/PVC

The photodegradation of the immobilized TiO2/ENR/PVC was carried out by using RR4 dye as a model pollutant. Figure 12 shows that the rate constant (k) value and percent dye removal increased with cycles of sample TEP24. The lowest rate constant (k) value was 0.0158 min−1. After 24 cycles of photoetching, the rate constant (k) value reached the highest value of 0.1066 min−1. The ENR from the immobilized TiO2 was leached out by the photoetching process, thus increasing the surface area of the immobilized TiO2 and improving the photocatalytic activity upon cycles.
The photoetched immobilized TiO2/ENR/PVC was about seven times faster as compared to the non-photoetched one, as shown in Figure 13. This is due to the opening ring and crosslinking effect and high contact area of the immobilized TiO2. The ENR-50 compound on the immobilized TiO2 surface was leached out during the photoetching process, thus increasing the contact area between TiO2 and RR4 dye molecules [49]. However, the P25-TiO2 suspension system was the uppermost, which is 8% better compared to the immobilized TiO2, as it has the highest contact area.
Figure 14 shows the model of immobilized TiO2 before and after the photoetching process. The porous immobilized TiO2 was created due to the leaching of ENR from the immobilized TiO2 during the photoetching process. Meanwhile, some of the ENR that remained in the immobilized TiO2 underwent the ring-opening and crosslinking reactions as proposed in Figure 10. Despite the possibilities of increasing the durability and adhesiveness, the product of the crosslinking also may act as an electron injector to the TiO2 as proven by the PL analysis in Figure 11, since the PL intensity for TEP24 is higher than for TEP0. Furthermore, some of the TiO2 particles on the upper layer will be leached out together with the ENR during the photoetching process, thus activating the internal TiO2 particles, which enlarge the contact area between the TiO2 and dye solution. Therefore, the photoetching process can improve the photodegradation rate.

4. Discussion

This works represented the effect of a photoetching process on the surface of immobilized TiO2/ENR/PVC and its reaction during the photoetching process. Photocatalytic degradation of RR4 dye by using a TiO2/ENR/PVC plate under various times of the photoetching process has successfully enhanced the photocatalytic activity, making them able to degrade the dye solution into a colorless solution. This finding is significant since other immobilized catalyst may use the photoetching process to enhanced their photocatalytic activity. Moreover, this method can be applied to real textile wastewater in the future.

Author Contributions

Conceptualization, W.I.N. and S.R.H.; methodology, W.I.N., S.R.H. and M.A.R.; software, M.S.A., S.R.H. and N.S.N.; validation, M.A.M.I., R.N. and W.I.N.; formal analysis, M.S.A.; investigation, N.A.M. and S.R.H.; resources, M.A.M.I., W.I.N., M.S.A. and R.N.; data curation, W.I.N.; writing—original draft preparation, S.R.H.; writing—review and editing, W.I.N., N.S.N., M.A.R. and N.I.A.G.; visualization, W.I.N. and S.R.H.; supervision, W.I.N.; project administration, W.I.N.; funding acquisition, M.A.M.I. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ministry of Higher Education, Malaysia (MOHE), grant number 600-RMC/GIP 5/3 (006/2022).

Data Availability Statement

Not applicable.

Acknowledgments

The authors would like to thank the Ministry of Higher Education, Malaysia (MOHE) for providing financial support under GIP grants: 600-RMC/GIP 5/3 (006/2022), Universiti Teknologi MARA (UiTM) in conducting this study. We would also like to acknowledge Universiti Malaysia Perlis (UniMAP) and Universiti Teknologi MARA (UiTM) for providing all of the facilities.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Aziz, K.H.H. Application of different advanced oxidation processes for the removal of chloroacetic acids using a planar falling film reactor. Chemosphere 2019, 228, 377–383. [Google Scholar] [CrossRef] [PubMed]
  2. Ijaz, M.; Zafar, M. Titanium dioxide nanostructures as efficient photocatalyst: Progress, challenges and perspective. Int. J. Energy Res. 2021, 45, 3569–3589. [Google Scholar] [CrossRef]
  3. Mortazavian, S.; Saber, A.; James, D.E. Optimization of photocatalytic degradation of Acid Blue 113 and Acid Red 88 textile dyes in a UV-C/TiO2 suspension system: Application of response surface methodology (RSM). Catalysts 2019, 9, 360. [Google Scholar] [CrossRef] [Green Version]
  4. Tetteh, E.K.; Rathilal, S.; Naidoo, D.B. Photocatalytic degradation of oily waste and phenol from a local South Africa oil refinery wastewater using response methodology. Sci. Rep. 2020, 10, 1–12. [Google Scholar] [CrossRef]
  5. Ikhwan, M.S.; Nazeri, N.S.; Natar, N.S.; Abdul Ghani, N.I.; Hamzah, S.R.; Rosli, M.A.; Mohd Ishak, M.A.; Mamauod, S.N.; Mohd Azami, M.S.; Wan Ismail, W.I.N. The surface distribution and recyclability study on photodegradation methylene blue dye of immobilized TiO2 under surfactant effect. Sci. Lett. 2022, 16, 40–50. [Google Scholar]
  6. 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]
  7. Nawi, M.A.; Zain, S.M. Enhancing the surface properties of the immobilized Degussa P-25 TiO2 for the efficient photocatalytic removal of methylene blue from aqueous solution. Appl. Surf. Sci. 2012, 258, 6148–6157. [Google Scholar] [CrossRef]
  8. Nawi, M.A.; Ngoh, Y.S.; Zain, S.M. Photoetching of Immobilized TiO2-ENR50-PVC Composite for Improved Photocatalytic Activity. Int. J. Photoenergy 2012. [Google Scholar] [CrossRef]
  9. Razak, S.; Nawi, M.A.; Haitham, K. Fabrication, characterization and application of a reusable immobilized TiO2–PANI photocatalyst plate for the removal of reactive red 4 dye. Appl. Surf. Sci. 2014, 319, 90–98. [Google Scholar] [CrossRef]
  10. Natar, N.S.; Ikhwan, S.; Nazeri, N.S.; Hamzah, S.R.; Rosli, M.A.; Ghani, N.I.; Ishak MA, M.; Razak, S.; Azami, M.S.; Nawawi, W.I. Preparation of water-base immobilized N doped TiO2 using DSAT technique for photocatalytic degradation of methylene blue dye. Mater. Today Proc. 2022, 66, 4036–4044. [Google Scholar] [CrossRef]
  11. Gonzalez, Z.; Yus, J.; Sanchez-Herencia, A.J.; Dewalque, J.; Manceriu, L.; Henrist, C.; Ferrari, B. A colloidal approach to prepare binder and crack-free TiO2 multilayer coatings from particulate suspensions: Application in DSSCs. J. Eur. Ceram. Soc. 2019, 39, 366–375. [Google Scholar] [CrossRef]
  12. Jadaa, W.; Prakash, A.; Ray, A.K. Photocatalytic Degradation of Diazo Dye over Suspended and Immobilized TiO2 Catalyst in Swirl Flow Reactor: Kinetic Modeling. Processes 2021, 9, 1741. [Google Scholar] [CrossRef]
  13. Silva, S.A.; Mineiro, S.L.; Oliveira, R.M.; Nono, M.D.C.A. Processing and Characterization of ZrO2-TiO2 Porous Films by Dip Coating for Application as Environmental Sensors. In Materials Science Forum; Trans Tech Publications Ltd.: Stafa-Zurich, Switzerland, 2018; Volume 930, pp. 631–636. [Google Scholar]
  14. Peiris, D.S.U.; Ekanayake, P.; Petra, M.I. Stacked rGO–TiO2 photoanode via electrophoretic deposition for highly efficient dye-sensitized solar cells. Org. Electron. 2018, 59, 399–405. [Google Scholar] [CrossRef]
  15. Zouzelka, R.; Remzova, M.; Brabec, L.; Rathousky, J. Photocatalytic performance of porous TiO2 layers prepared by quantitative electrophoretic deposition from organic solvents. Appl. Catal. Environ. 2018, 227, 70–78. [Google Scholar] [CrossRef]
  16. Khanmohammadi, S.; Ojaghi-Ilkhchi, M.; Farrokhi-Rad, M. Development of bioglass coating reinforced with hydroxyapatite whiskers on TiO2 nanotubes via electrophoretic deposition. Ceram. Int. 2021, 47, 1333–1343. [Google Scholar] [CrossRef]
  17. Thompson, W.A.; Perier, C.; Maroto-Valer, M.M. Systematic study of sol-gel parameters on TiO2 coating for CO2 photoreduction. Appl. Catal. Environ. 2018, 238, 136–146. [Google Scholar] [CrossRef]
  18. Ahadi, S.; Moalej, N.S.; Sheibani, S. Characteristics and photocatalytic behavior of Fe and Cu doped TiO2 prepared by combined sol-gel and mechanical alloying. Solid State Sci. 2019, 96, 105975. [Google Scholar] [CrossRef]
  19. Estrada-Flores, S.; Martínez-Luévanos, A.; Perez-Berumen, C.M.; García-Cerda, L.A.; Flores-Guia, T.E. Relationship between morphology, porosity, and the photocatalytic activity of TiO2 obtained by sol–gel method assisted with ionic and nonionic surfactants. Boletín Soc. Española Cerámica Vidr. 2020, 59, 209–218. [Google Scholar] [CrossRef]
  20. Jaramillo-Raquejo, D.; Palacio-Espinosa, C.C.; Ageorges, H. Plasma Spray Parameters of TiO2 Targets Used in Magnetron Sputtering. TecnoLógicas 2020, 23, 135–155. [Google Scholar]
  21. Kozlovskiy, A.; Shlimas, I.; Dukenbayev, K.; Zdorovets, M. Structure and corrosion properties of thin TiO2 films obtained by magnetron sputtering. Vacuum 2019, 164, 224–232. [Google Scholar] [CrossRef]
  22. Samiee, M.; Hanachi, M.; Seyedraoufi, Z.S.; Eshraghi, M.J.; Shajari, Y. Biodegradable magnesium alloy coated with TiO2/MgO two-layer composite via magnetic sputtering for orthopedic applications: A study on the surface characterization, corrosion, and biocompatibility. Ceram. Int. 2021, 47, 6179–6186. [Google Scholar] [CrossRef]
  23. Abdul Ghani, N.; Rosli, M.; Hamzah, S.; Natar, N.; Nazeri, N.; Ab Aziz, S.; Mohamad, S.; Wan Ismail, W. Water-based Preparation of Immobilized Ag-doped TiO2 Photocatalyst for Photocatalytic Degradation of RR4 Dye. Sci. Lett. 2022, 16, 24–39. [Google Scholar]
  24. Chen, D.; Cheng, Y.; Zhou, N.; Chen, P.; Wang, Y.; Li, K.; Huo, S.; Cheng, P.; Peng, P.; Zhang, R.; et al. Photocatalytic degradation of organic pollutants using TiO2-based photocatalysts: A review. J. Clean. Prod. 2020, 268, 121725. [Google Scholar] [CrossRef]
  25. Tennakone, K.; Tilakaratne, C.T.K.; Kottegoda, I.R.M. Photocatalytic degradation of organic contaminants in water with TiO2 supported on polythene films. J. Photochem. Photobiol. Chem. 1995, 87, 177–179. [Google Scholar] [CrossRef]
  26. Nawawi, W.I.; Zaharudin, R.; Ishak, M.A.M.; Ismail, K.; Zuliahani, A. The preparation and characterization of immobilized TiO2/PEG by using DSAT as a support binder. Appl. Sci. 2017, 7, 24. [Google Scholar] [CrossRef] [Green Version]
  27. Tasić, N.; Branković, Z.; Marinković-Stanojević, Z.; Branković, G. Effect of binder molecular weight on morphology of TiO2 films prepared by tape casting and their photovoltaic performance. Sci. Sinter. 2012, 44, 365–372. [Google Scholar] [CrossRef]
  28. Wan Ismail, W.I.N.; Ain, S.K.; Zaharudin, R.; Jawad, A.H.; Ishak, M.A.M.; Ismail, K.; Sahid, S. New TiO2/DSAT immobilization system for photodegradation of anionic and cationic dyes. Int. J. Photoenergy 2015, 3, 1–6. [Google Scholar] [CrossRef]
  29. Zhang, X.; Xiao, J.; Peng, C.; Xiang, Y.; Chen, H. Enhanced photocatalytic hydrogen production over conjugated polymer/black TiO2 hybrid: The impact of constructing active defect states. Appl. Surf. Sci. 2019, 465, 288–296. [Google Scholar] [CrossRef]
  30. Nawawi, W.I.; Zaharudin, R.; Zuliahani, A.; Shukri, D.S.; Azis, T.F.; Razali, Z. Immobilized TiO2-polyethylene glycol: Effects of aeration and pH of methylene blue dye. Appl. Sci. 2017, 7, 508. [Google Scholar] [CrossRef] [Green Version]
  31. Altın, I.; Sökmen, M. Preparation of TiO2-polystyrene photocatalyst from waste material and its usability for removal of various pollutants. Appl. Catal. Environ. 2014, 144, 694–701. [Google Scholar] [CrossRef]
  32. Cai, W.; Yang, H.; Han, D.; Guo, X. A physical route to porous ethyl cellulose microspheres loaded with TiO2 nanoparticles. J. Appl. Polym. Sci. 2014, 131, 19. [Google Scholar] [CrossRef]
  33. Ribeiro, L.N.; Fonseca, A.C.; da Silva, E.F.; Oliveira, E.D.; Ribeiro, A.T.; Maranhão, L.C.; Pacheco, J.G.; Machado, G.; Almeida, L.C. Residue-based TiO2/PET photocatalytic films for the degradation of textile dyes: A step in the development of green monolith reactors. Chem. Eng. Process.-Process Intensif. 2020, 147, 107792. [Google Scholar] [CrossRef]
  34. Pan, Y.; Liu, B.; Hu, H.; Jiang, H.; Wei, F.; Zhou, M.; Fan, X.; Yu, H.; Niu, G.; Huang, J. Preparation and photocatalytic performance of the rod-shaped Ni-NiO/TiO2 hollow composite structure based on metallization cellulose fibers and TBOT. Vacuum 2019, 159, 1–8. [Google Scholar] [CrossRef]
  35. Bui, D.P.; Huu Pham, H.; Minh Cao, T.; Van Pham, V. Preparation of conjugated polyvinyl chloride/TiO2 nanotubes for Rhodamine B photocatalytic degradation under visible light. J. Chem. Technol. Biotechnol. 2020, 95, 2707–2714. [Google Scholar]
  36. Petković, G.; Vukoje, M.; Bota, J.; Pasanec Preprotić, S. Enhancement of polyvinyl acetate (PVAc) adhesion performance by SiO2 and TiO2 nanoparticles. Coatings 2019, 9, 707. [Google Scholar] [CrossRef] [Green Version]
  37. Toh-ae, P.; Junhasavasdikul, B.; Lopattananon, N.; Sahakaro, K. Mechanical properties and stability towards heat and UV irradiation of natural rubber/nanotitanium dioxide composites. Procedia Chem. 2016, 19, 139–147. [Google Scholar]
  38. Singh, S.; Mahalingam, H.; Singh, P.K. Polymer-supported titanium dioxide photocatalysts for environmental remediation: A review. Appl. Catal. Gen. 2013, 462, 178–195. [Google Scholar] [CrossRef]
  39. Sabri, N.A.; Nawi, M.A.; Bakar, N.A. Recyclable immobilized carbon coated nitrogen doped TiO2 for photocatalytic degradation of quinclorac under UV–vis and visible light. J. Environ. Chem. Eng. 2018, 6, 898–905. [Google Scholar] [CrossRef]
  40. Bahrudin, N.N.; Nawi, M.A.; Nawawi, W.I. Photocatalytic enhancement of immobilized TiO2-polyaniline bilayer (TiO2-PBL) system for decolorization of methyl orange dye. Mater. Res. Bull. 2018, 106, 388–395. [Google Scholar] [CrossRef]
  41. Razak, S.; Salleh, N.; Nawawi, W.I.; Natar, N. Effect of Photoetching Process onto Immobilized PANI/TiO2 Films Towards Photocatalytic Degradation of RR2 dye. In IOP Conference Series: Materials Science and Engineering; IOP Publishing: Bristol, UK, 2020; Volume 957, p. 12025. [Google Scholar]
  42. Sadeghi, S.M.; Vaezi, M.; Kazemzadeh, A.; Jamjah, R. 3D networks of TiO2 nanofibers fabricated by sol-gel/electrospinning/calcination combined method: Valuation of morphology and surface roughness parameters. Mater. Sci. Eng. 2021, 271, 115254. [Google Scholar] [CrossRef]
  43. Theivasanthi, T.; Alagar, M. Titanium dioxide (TiO2) nanoparticles XRD analyses: An insight. arXiv 2013, arXiv:1307.1091. [Google Scholar]
  44. Pratheepa, M.I.; Lawrence, M. X-ray diffraction analyses of titanium dioxide nanoparticles. J. Sci. Res. Sci. Tec. 2018, 3, 83. [Google Scholar]
  45. Joni, I.M.; Nulhakim, L.; Panatarani, C. Characteristics of TiO2 particles prepared by simple solution method using TiCl3 precursor. In Journal of Physics: Conference Series; IOP Publishing: Bristol, UK, 2018; Volume 1080, p. 12042. [Google Scholar]
  46. Yao, L.; Li, X.; Liu, H.; Li, Z.; Lu, Q. One-dimensional hierarchical CeVO4/TiO2 heterostructures with enhanced photocatalytic performance. J. Nanoparticle Res. 2019, 21, 1–13. [Google Scholar] [CrossRef]
  47. Quesada-Gonzalez, M.; Williamson, B.A.; Sotelo-Vazquez, C.; Kafizas, A.; Boscher, N.D.; Quesada-Cabrera, R.; Scanlon, D.O.; Carmalt, C.J.; Parkin, I.P. Deeper understanding of interstitial boron-doped anatase thin films as a multifunctional layer through theory and experiment. J. Phys. Chem. 2018, 122, 714–726. [Google Scholar] [CrossRef]
  48. Hamzah, R.; Bakar, M.A.; Dahham, O.S.; Zulkepli, N.N.; Dahham, S.S. A structural study of epoxidized natural rubber (ENR-50) ring opening under mild acidic condition. J. Appl. Polym. Sci. 2016, 133, 43. [Google Scholar] [CrossRef]
  49. Sabri, N.A.; Nawi, M.A.; Nawawi, W.I. Porous immobilized C coated N doped TiO2 containing in-situ generated polyenes for enhanced visible light photocatalytic activity. Opt. Mater. 2015, 48, 258–266. [Google Scholar] [CrossRef]
Figure 1. Field emission scanning electron microscopy (FESEM) of (a,b) TEP0, (c,d) TEP12, (e,f) TEP24 and (g,h) TEP30 under 5000× and 10,000× magnification, respectively.
Figure 1. Field emission scanning electron microscopy (FESEM) of (a,b) TEP0, (c,d) TEP12, (e,f) TEP24 and (g,h) TEP30 under 5000× and 10,000× magnification, respectively.
Colorants 02 00006 g001
Figure 2. (a) 2D surface profile and (b) 3D surface profile of TEP0 and (c) 2D surface profile and (d) 3D surface profile of TEP24.
Figure 2. (a) 2D surface profile and (b) 3D surface profile of TEP0 and (c) 2D surface profile and (d) 3D surface profile of TEP24.
Colorants 02 00006 g002
Figure 3. X-ray diffraction patterns of TEP0 and TEP24.
Figure 3. X-ray diffraction patterns of TEP0 and TEP24.
Colorants 02 00006 g003
Figure 4. Structure of TiO2. The grey and red balls indicate Ti and O atoms, respectively.
Figure 4. Structure of TiO2. The grey and red balls indicate Ti and O atoms, respectively.
Colorants 02 00006 g004
Figure 5. Calculated band structure of TiO2 in (a) TEP0 and (b) TEP24 samples.
Figure 5. Calculated band structure of TiO2 in (a) TEP0 and (b) TEP24 samples.
Colorants 02 00006 g005
Figure 6. (a) UV–Vis DRS and (b) corresponding Kubelka–Munk function plots of different samples.
Figure 6. (a) UV–Vis DRS and (b) corresponding Kubelka–Munk function plots of different samples.
Colorants 02 00006 g006
Figure 7. Density of states for (a) TEP0 and (b) TEP24. Green solid lines: TDOS, red solid lines: O 2p states, and blue solid lines: Ti 3d states. The black dashed line represents the position of the Fermi level.
Figure 7. Density of states for (a) TEP0 and (b) TEP24. Green solid lines: TDOS, red solid lines: O 2p states, and blue solid lines: Ti 3d states. The black dashed line represents the position of the Fermi level.
Colorants 02 00006 g007
Figure 8. Comparison of Fourier transform infrared (FTIR) analysis between TEP0, TEP24, and degussa P25–TiO2 in (a) a wide range wavenumber (4000–600 cm−1) and (b) narrow range wavenumber (1300–1000 cm−1).
Figure 8. Comparison of Fourier transform infrared (FTIR) analysis between TEP0, TEP24, and degussa P25–TiO2 in (a) a wide range wavenumber (4000–600 cm−1) and (b) narrow range wavenumber (1300–1000 cm−1).
Colorants 02 00006 g008
Figure 9. (a) The wide X-ray photoelectron spectroscopy (XPS) spectrum of (b) C1s (c) O1s, and (d) Cl2p for TEP24. CPS, cycles per second.
Figure 9. (a) The wide X-ray photoelectron spectroscopy (XPS) spectrum of (b) C1s (c) O1s, and (d) Cl2p for TEP24. CPS, cycles per second.
Colorants 02 00006 g009
Figure 10. Possible reaction mechanism during the photoetching process.
Figure 10. Possible reaction mechanism during the photoetching process.
Colorants 02 00006 g010
Figure 11. Photoluminescence (PL) spectroscopy of TEP0 and TEP24.
Figure 11. Photoluminescence (PL) spectroscopy of TEP0 and TEP24.
Colorants 02 00006 g011
Figure 12. Pseudo first-order rate constant (min−1) and percent removal of RR4 dye (%) vs. number of photocatalysis cycles using the TEP24 sample.
Figure 12. Pseudo first-order rate constant (min−1) and percent removal of RR4 dye (%) vs. number of photocatalysis cycles using the TEP24 sample.
Colorants 02 00006 g012
Figure 13. Pseudo first-order rate constant, (min−1) for removal of 30 mg L−1 RR4 dye via photocatalysis using P25–TiO2 suspension system, and immobilized TEP0 and TEP24 catalyst plate.
Figure 13. Pseudo first-order rate constant, (min−1) for removal of 30 mg L−1 RR4 dye via photocatalysis using P25–TiO2 suspension system, and immobilized TEP0 and TEP24 catalyst plate.
Colorants 02 00006 g013
Figure 14. Diagram of immobilized TiO2 surface before and after the photoetching process.
Figure 14. Diagram of immobilized TiO2 surface before and after the photoetching process.
Colorants 02 00006 g014
Table 1. Crystallinity index (CI) of TEP0 and TEP24.
Table 1. Crystallinity index (CI) of TEP0 and TEP24.
SampleCrystallinity Index (%)
TEP051.21
TEP2457.30
Table 2. Lattice parameter, volume, and bond length of TEP24 and TEP0.
Table 2. Lattice parameter, volume, and bond length of TEP24 and TEP0.
Lattice Parameter (Å)Volume (Å3)Bond Length (Å)
TEP24a = 3.7752136.4475Ti(1)-O1.9310
b = 3.7752 Ti(2)-O1.9861
c = 9.5742 O(1)-O2.4608
O(2)-O2.7910
TEP0a = 3.7758136.3951Ti(1)-O1.93102
b = 3.7758 Ti(2)-O1.98602
c = 9.5670 O(1)-O2.46201
O(2)-O2.79050
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Hamzah, S.R.; Rosli, M.A.; Natar, N.S.; Ghani, N.I.A.; Muhamad, N.A.; Azami, M.S.; Ishak, M.A.M.; Nordin, R.; Nawawi, W.I. The Crosslinking and Porosity Surface Effects of Photoetching Process on Immobilized Polymer-Based Titanium Dioxide for the Decolorization of Anionic Dye. Colorants 2023, 2, 73-89. https://doi.org/10.3390/colorants2010006

AMA Style

Hamzah SR, Rosli MA, Natar NS, Ghani NIA, Muhamad NA, Azami MS, Ishak MAM, Nordin R, Nawawi WI. The Crosslinking and Porosity Surface Effects of Photoetching Process on Immobilized Polymer-Based Titanium Dioxide for the Decolorization of Anionic Dye. Colorants. 2023; 2(1):73-89. https://doi.org/10.3390/colorants2010006

Chicago/Turabian Style

Hamzah, Siti Raihan, Muhammad Afiq Rosli, Nadiah Sabihah Natar, Nureel Imanina Abdul Ghani, Nur Aien Muhamad, Mohammad Saifulddin Azami, Mohd Azlan Mohd Ishak, Razif Nordin, and Wan Izhan Nawawi. 2023. "The Crosslinking and Porosity Surface Effects of Photoetching Process on Immobilized Polymer-Based Titanium Dioxide for the Decolorization of Anionic Dye" Colorants 2, no. 1: 73-89. https://doi.org/10.3390/colorants2010006

Article Metrics

Back to TopTop