Next Article in Journal
Improvement of GaN-Based Device Performance by Plasma-Enhanced Chemical Vapor Deposition (PECVD) Directly Preparing h-BN with Excellent Thermal Management Characteristics
Previous Article in Journal
Advancing Cancer Treatment and Diagnosis: A Review on Photodynamic Therapy Using OLED Technology
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis and Characterization of Se4+@TiO2/PET Composite Photocatalysts with Enhanced Photocatalytic Activity by Simulated Solar Irradiation and Antibacterial Properties

1
School of Textile and Clothing, Nantong University, Nantong 226019, China
2
College of Textile and Clothing, Yancheng Institute of Technology, Yancheng 224051, China
3
School of Microelectronics and Integrated Circuits, Nantong University, Nantong 226019, China
*
Authors to whom correspondence should be addressed.
Molecules 2025, 30(6), 1306; https://doi.org/10.3390/molecules30061306
Submission received: 18 January 2025 / Revised: 7 March 2025 / Accepted: 12 March 2025 / Published: 14 March 2025

Abstract

:
To fabricate recyclable catalytic materials with high catalytic activity, Se4+@TiO2 photocatalytic materials were synthesized by the sol–gel method. By introducing free radicals on the surface of polyester (PET) fabrics through plasma technology, Se4+@TiO2/PET composite photocatalytic materials with high photocatalytic activity were prepared. The surface morphology, crystal structure, chemical composition, and photocatalytic performance were characterized by scanning electron microscopy (SEM), X-ray diffraction (XRD), X-ray photoelectron spectroscopy (XPS), ultraviolet–visible absorption spectroscopy (UV–Vis), and photoluminescence spectroscopy (PL), respectively. The photocatalytic degradation performance was determined by assessing the degradation of azo dye methyl orange under simulated solar irradiation. The results demonstrated that Se4+@TiO2/PET exhibited a superior degradation rate of methyl orange, reaching up to 81% under simulated sunlight. The PL spectra indicated that the electron–hole pair separation rate of Se4+@TiO2/PET was higher than that of TiO2/PET. Furthermore, UV–Vis spectroscopy demonstrated that the relative forbidden band gap of Se4+@TiO2/PET was determined to be 2.9 eV. The band gap of Se4+@TiO2/PET was narrower, and the absorption threshold shifted toward the visible region, indicating a possible increase in its catalytic activity in simulated solar irradiation. In addition, the antibacterial properties of Se4+@TiO2/PET were subsequently investigated, achieving 99.99% and 98.47% inhibition against S. aureus and E. coli, respectively.

1. Introduction

Semiconductor photocatalytic materials are crucial in environmental remediation and sustainable energy applications due to their ability to harness solar energy to drive chemical reactions [1,2]. In recent decades, they have been employed in a variety of applications, including the treatment of organic pollutants [3,4], the production of hydrogen by photolysis of water [5,6], the reduction of CO2 [7,8], and the degradation of microbials [9,10]. These materials typically include metal oxides such as titanium dioxide (TiO2), zinc oxide (ZnO), and other metal sulfides and nitrides.
Titanium dioxide (TiO2) is one of the most commonly used and highly efficient photocatalysts, which is characterized by low cost, non-toxicity, good stability, and high photocatalytic activity [11,12,13]. When TiO2 is exposed to light with energy equal to or greater than its forbidden band energy (Eg), electrons in the valence band (VB) are excited to the conduction band (CB). At the same time, holes are left in the valence band, forming electron (e) and hole (h+) pairs. When exposed to oxygen and water vapor, the TiO2 will produce hydroxyl radicals (·OH), hydrogen peroxide (H2O2), and reactive oxygen species (ROS), such as the superoxide radical (O2·), which can rapidly decompose organic pollutants or bacteria [14]. However, TiO2 has a high forbidden band gap (Eg 3.2 eV), which renders it primarily sensitive to UV light and less applicable to simulated sunlight [15,16]. Its electronic and energy band structures can be modified to enhance its simulated sunlight responsiveness [17].
In recent years, numerous methods have been employed to modify TiO2 to enhance its photocatalytic performance. These include ion doping [18,19], noble metal deposition (Pt, Pd, Au, Ag, etc.) [20,21], and semiconductor compounding to form p-n heterojunctions [22,23]. Among these methods, ion doping modification is regarded as a competent technique that can be employed to achieve an efficient response to solar irradiation by modifying the electronic structure of TiO2 nanoparticles. Doping is a highly effective technique for suppressing the recombination of light-generated carriers and enhancing photocatalytic activity. By introducing dopant ions, TiO2 can extend its active wavelength range from UV to visible light, thereby improving solar photocatalytic efficiency [24,25]. Dopants create new energy levels, reduce the band gap energy, broaden solar absorption, and enhance TiO2 nanoparticles’ photocatalytic activity under simulated sunlight [26]. Additionally, dopant ions can act as shallow traps for electrons or holes, significantly reducing electron–hole recombination rates and prolonging carrier lifetimes, further boosting photocatalytic performance [27,28]. Transition metal ions, including Fe3+, Cr3+, Co2+, and Ni2+, have been employed to dope modified TiO2 nanoparticles, to enhance their photocatalytic performance [29,30,31,32]. Nevertheless, some studies have indicated that during the photocatalytic degradation of organic pollutants in water, the presence of metal cations in nano-TiO2 can result in the formation of composite pollutants with organic pollutants, which may subsequently contribute to environmental degradation [33].
Selenium is an essential component of numerous antioxidant enzymes in living organisms and exhibits several advantageous properties, including high bioactivity, biocompatibility, and environmental friendliness [34,35]. It has been demonstrated that selenium has good broad-spectrum photosensitizing properties and forms p-n heterojunctions with nano-TiO2 to extend the photo response range of nano-TiO2 [36,37]. While Se exists in various chemical valencies, including Se4+, Se6+, and Se2−, some calculations have shown that Se4+ doping of TiO2 can introduce additional electronic states in the band gap [38]. Gurkan et al. [38] prepared Se4+doped TiO2 photocatalysts with varying concentrations using an impregnation method to investigate the photocatalytic degradation kinetics. The results demonstrated that 0.50% Se4+doped TiO2 exhibited the highest degradation rate of 4-nitrophenol. Wei et al. [39] modified nano TiO2 with Se4+ and increased the doping concentration of selenium in nano TiO2 to 17.1%. The band gap width of the nano-TiO2 was significantly reduced, resulting in enhanced photocatalytic activity under solar irradiation.
Traditional photocatalysts are typically in powder form, posing challenges for separation and recovery from wastewater, leading to resource wastage and environmental pollution. Polyester (PET) fibers, valued for their high strength, corrosion resistance, heat tolerance, and affordability, serve as an ideal matrix for loading nano-TiO2 photocatalysts. This integration facilitates the recyclability of photocatalytic materials and reduces overall costs [40,41]. In this study, the synthesis of the Se4+@TiO2 photocatalytic material was achieved through the sol–gel method. PET fabrics were pretreated with plasma technology, which created active sites for the Se4+@TiO2 to enhance the bonding fastness to PET fabrics. The subsequent preparation of the Se4+@TiO2/PET material was conducted by an impregnation method. The surface morphology, chemical composition, crystal structure, and surface properties of the PET fabrics were analyzed. Meanwhile, the alterations in the surface morphology, chemical composition, and crystal structure of polyester fabrics following finishing were analyzed, thereby providing a theoretical foundation for the utilization of Se4+modified nano-TiO2 in the field of photocatalytic antimicrobial textiles.

2. Results and Discussion

2.1. SEM Analyses

Figure 1 illustrates the surface morphology of the PET fabric before and after treatment. Figure 1a depicts an SEM image of the PET fabric with 5000 times magnification. This image reveals that the surface of the untreated PET fabric is smooth and free of impurities. Figure 1b depicts an SEM image of TiO2/PET with 5000 times magnification. It can be observed that the nano-TiO2 is more uniformly distributed on the surface of the PET, with an average size of approximately 100–120 nm. Figure 1c depicts an SEM image of Se4+@TiO2/PET with 5000 times magnification, which has been synthesized by the sol–gel method. The average size of the Se4+@TiO2 is 60–80 nm. There is local aggregation on the surface of the PET fabric due to the easier aggregation of nanoparticles.

2.2. EDS Analyses

Figure 2 illustrates the surface elemental distribution of the PET fabric before and after the treatment process. The elemental composition of the untreated PET fabric consists primarily of carbon and oxygen, with relative contents of 70.83% and 29.17%, respectively, as shown in Figure 2a. In Figure 2b, the elemental distribution of the TiO2/PET composite reveals an increased proportion of titanium (Ti), which results from the incorporation of TiO2 onto the PET surface, with a relative Ti content of 8.29%. Figure 2c demonstrates a uniform distribution of Ti and Se elements across the PET fabric surface. The results indicate that the mass percentages (wt%) of Ti and Se are 7.55% and 8.47%, respectively. The presence of these elements confirms the successful loading of TiO2 and Se.

2.3. XRD Analyses

The crystalline structures of the three materials, PET, TiO2/PET, and Se4+@TiO2/PET, are depicted in Figure 3. In Figure 3a, the diffraction peaks of the three curves, indicated by the circles at 17.2°, 22.6°, and 25.1°, correspond to the diffraction peaks of the PET [42]. The diffraction peaks of TiO2/PET, indicated by the five positions of 25.2°, 37.5°, 47.8°, 53.5°, and 62.3°, correspond to the crystallographic planes of (101), (004), (200), (105), and (204) in PDF#71–1168, which are the diffraction peaks of anatase TiO2 [43,44]. As illustrated in Figure 3b, the XRD diffractogram of Se4+@TiO2/PET also exhibits the characteristic diffraction peaks of anatase-phase TiO2, thereby indicating that Se4+ does not affect the physical phase of nano TiO2. Moreover, the diffraction peaks of Se4+@TiO2/PET at the (101) crystal plane exhibit a slight shift to a lower angle, which indicates that Se4+ enters the substitution sites of the TiO2 crystal structure [45,46]. The Deby–Scherrer formula Equation (1) was employed for the calculation of the crystal size [47]:
D = k λ β cos θ
where D is the crystal size, nm; k is the Scherrer constant and taken as 0.89; λ is the X-ray wavelength and taken as 0.15406 for Cu kα; β is the full width at the half maximum intensity of the peak (FWHM); θ is the angle of diffraction, 2θ/2. From this equation, the mean diameter of the TiO2 nanoparticles was determined to be 100 nm, while that of the Se4+@TiO2 particles was found to be 60 nm. Concurrently, calculations demonstrated that the size of the Se4+@TiO2 crystals exhibited a reduction in comparison to that of TiO2, accompanied by an increase in disorder or defects within the crystals. This phenomenon can be attributed to the lattice deformation caused by the doping of Se4+ ions, which results in the formation of new defects in the TiO2 grains, thereby reducing the orderliness of the material [38].

2.4. XPS Elemental Analysis

Figure 4a illustrates the full peaks of TiO2/PET, which contain C 1s, O 1s, and Ti 2p signals, while those of Se4+@TiO2/PET contain C 1s, O 1s, Ti 2p, and Se 3d signals. The peaks of the C and O elements are primarily derived from PET fabrics [48]. Figure 4b illustrates the binding energy of C 1s. The peaks observed in the two curves in the figure correspond to the O-C=O, C-O-C, and C-C bonds in the PET, respectively. Figure 4c depicts the photoelectron spectra of O 1s. Peaks at 531.48 eV in TiO2/PET may be attributed to surface hydroxyl (-OH) groups and another peak at 528.93 eV is caused by lattice oxygen. Figure 4d displays the photoelectron spectra of elemental titanium. The peaks observed at 457.88 eV and 463.58 eV for TiO2/PET correspond to the Ti 2p3/2 and Ti 2p1/2 signals, respectively. The difference in binding energies between the two peaks is 5.7 eV, which indicates the presence of Ti4+ in the material [49,50,51].
The Ti 2p3/2 and Ti 2p1/2 spectra in Se4+@TiO2/PET exhibit two major characteristic peaks at 458.30 eV and 464.00 eV, respectively. The difference in binding energy between the two peaks is 5.7 eV, which also suggests that the elemental valence state of Ti in Se4+@TiO2/PET is also +4 valence. Further analysis demonstrates that Se4+ doping into TiO2 results in a shift of the Ti 2p3/2 peak from 457.88 eV to 458.30 eV, while the Ti 2p1/2 peak shifts from 463.58 eV to 464.00 eV. Both peaks exhibit a shift towards higher binding energy. Any shift greater than 0.2 eV indicates a genuine change in the XPS spectrogram, suggesting the emergence of novel species in the XPS spectrogram of the TiO2 catalyst surface [52]. These shifts are due to the fact that Se4+ occupies the position of Ti4+ after Se4+ doping of TiO2. The electron transfer from Ti4+ to Se4+ decreases the electron cloud density around the Ti atom’s nucleus, increases its positive charge density, and raises the binding energy, indicating successful doping of Se4+ into the TiO2 lattice [38]. Figure 4e presents the Se 3d spectra of Se4+@TiO2/PET, and the two prominent Se 3d peaks are evident in the image. In Figure 4e, the red line represents Se 3d 3/2 while the blue line represents Se 3d 5/2, where the Se 3d 5/2 peak corresponds to Se4+.

2.5. UV–Vis and PL Analyses

In order to ascertain the optical properties of the photocatalytic composites, UV–visible and PL tests were conducted [53]. The energy band structure of the semiconductor materials can be obtained analytically through UV–Vis absorption spectroscopy. Figure 5a illustrates the UV–Vis absorption spectra of TiO2/PET and Se4+@TiO2/PET, spanning a wavelength range of 200–800 nm. Compared to TiO2/PET, Se4+@TiO2/PET exhibited enhanced absorption properties within the visible light range. The absorption spectra of Se4+@TiO2/PET exhibited a redshift within the simulated sunlight range, suggesting an increase in its photocatalytic activity within the simulated sunlight range. This was attributed to the change in the conduction and valence bands in Se4+@TiO2/PET [54]. The photocatalytic effect can be formulated as follows (Equation (2)) [55]:
( α h ν ) 2 = A ( h ν E g )
where A is a constant, α is the light absorption coefficient, h is Planck’s constant, ν is the optical frequency, and Eg is the forbidden band gap of the material.
Figure 5b plots (αhν)2 versus and extrapolates the linear portion of the plot to the energy axis to determine the energy band gap. The forbidden band gap of TiO2/PET is 3.1 eV, while the relative forbidden band gap of Se4+@TiO2/PET is 2.9 eV. The narrowing of the band gap of Se4+@TiO2/PET reduces the rate of photogenerated electron and hole complexation. This, in turn, leads to a shift in the absorption threshold (400 nm for TiO2/PET and 476 nm for Se4+@TiO2/PET) towards the visible light region, thereby broadening the range of light utilization. Consequently, the simulated solar irradiation photocatalytic activity of the composite photocatalytic materials is enhanced by the incorporation of Se4+ into TiO2.
The photogenerated electron–hole complexation and trapping processes of each catalyst were analyzed by photoluminescence spectroscopy, which enables the characterization of the information about the separation and complexation of photogenerated carriers in semiconductors [56,57]. A reduction in spectral intensity indicates a diminished photogenerated electron–hole recombination reaction, which is associated with enhanced photocatalytic performance. Figure 6 illustrates the photoluminescence spectra of TiO2/PET and Se4+@TiO2/PET. Figure 6 reveals a notable decline in the photoluminescence intensity of Se4+@TiO2/PET compared to TiO2/PET. The electron–hole complexation rate is proportional to the photoluminescence spectral intensity, and the faster the complexation rate, the stronger the spectral intensity. Conversely, a lower photoluminescence (PL) intensity indicates that the photogenerated electrons are trapped and safely transferred to the surface of the photocatalytic material, where they react with adsorbed oxygen or water molecules [58,59]. The lower PL intensity of Se4+@TiO2/PET suggests a lower rate of complexation, which in turn implies better photocatalytic activity.

2.6. Photocatalytic Performance Analysis

The photocatalytic activities of TiO2/PET and Se4+@TiO2/PET on a methyl orange (MO) solution were compared under simulated sunlight irradiation, as illustrated in Figure 7. The samples were magnetically stirred in a dark environment for 30 min to achieve adsorption equilibrium between the photocatalysts and the MO solution. Following a 30 min stirring period in the absence of light, the MO with the composite photocatalyst was subjected to irradiation with an xenon lamp for a duration of 120 min, and the supernatant was taken every 20 min to test the absorbance. The control sample was a 20 mg/L MO, and the degradation efficiency was calculated. Figure 7 illustrates that in the absence of any photocatalytic material, the MO undergoes slight self-degradation, with a degradation efficiency of 10%. The addition of PET results in an increase in the removal efficiency of MO by PET, reaching 17%. Following the loading of nano-TiO2, the removal efficiency of MO by TiO2/PET was found to be 19%, while that of MO by Se4+@TiO2/PET was 81%. In comparison to TiO2/PET, the photocatalytic performance of Se4+@TiO2/PET for MO was found to be significantly enhanced.
The photocatalytic degradation of MO was analyzed according to the Langmuir–Hinshelwood kinetic model [60,61] with the kinetic Equation (3):
r = d C d t = k r k s C 0 1 + k s C 0
where C0 is the initial concentration of the dye, kr is the reaction rate constant, ks is the apparent adsorption constant, and t is the reaction time. When the concentration is low, the ksC0 value is deemed insignificant, allowing the reaction rate to be expressed through a pseudo first-order model (Equation (4)):
d C d t = k r k s C 0 = K a p p C 0
The integration of Equation (3) yields the following result (Equation (5)):
l n ( C C 0 ) = K a p p t
where Kapp is defined as the slope of the linear regression equation. The pseudo-primary kinetic equation is used to linearly fit the experimental data, resulting in the apparent rate constants Kapp and the correlation coefficients R2 for different samples, as shown in Table 1. In conjunction with Figure 7 and Figure 8, it can be observed that following 120 min of simulated sunlight irradiation, Se4+@TiO2/PET exhibited the most pronounced degradation efficiency of MO, reaching 81%. This was accompanied by an apparent rate constant of 0.0101 min−1, which was 4.4 times higher than that of TiO2/PET. The results indicated that the photocatalytic degradation followed a pseudo-primary kinetic equation (Figure 8).
To evaluate the reusability of the Se4+@TiO2/PET composites, the photocatalytic degradation tests were performed on the samples for three times after recovery, collection, washing, and drying as illustrated in Figure 9. The results showed that the degradation efficiency changed from 81% to 76% by Equation (13), indicating the good recyclability and stability of the composites.

2.7. EIS Analyses

As demonstrated in Figure 10, the electrochemical impedance spectra of TiO2/PET and Se4+@TiO2/PET reflect the interfacial resistance during the electron-leaping process in these samples. In comparison with TiO2/PET, the composite photocatalytic material Se4+@TiO2/PET exhibited a significantly reduced circular radius, indicating higher electronic conductivity. This is attributed to the lower resistance of the photocharge transfer at the interface between the Se4+@TiO2 photocatalyst and the electrolyte [62,63].

2.8. Radical Scavenger Analysis

Table 2 demonstrates the degradation efficiency of Se4+@TiO2/PET on a methyl orange solution with the addition of various free radical scavengers. In the absence of free radical scavengers, the degradation efficiency of methyl orange was found to be 81%. The addition of 10 mmol/L isopropanol (IPA) resulted in the capture of the hydroxyl radical (·OH), which led to a degradation rate of 77.28% for methylene orange. The addition of 1 mmol/L p-benzoquinone (BQ) resulted in the capture of the superoxide radical (O2·), with a subsequent reduction in the degradation rate of methyl orange to 47.11%. The degradation rate of methyl orange was 72.64% after the addition of 10 mmol/L EDTA-2Na, which served to capture the hole (h+). It can thus be inferred from the capture experiments that the primary substance responsible for dye degradation and bacterial inactivation is the superoxide radical (O2·).

2.9. Antimicrobial Performance Analysis

The antibacterial property test of TiO2/PET and Se4+@TiO2/PET against Escherichia coli (E. coli) and Staphylococcus aureus (S. aureus) are presented in Figure 11 and Table 3, respectively. As illustrated in Table 3, the inhibition rates of TiO2/PET against E. coli and S. aureus were relatively low, at 58.04% and 88.31%, respectively. Additionally, TiO2 exhibited a toxic effect on bacteria under dark conditions, which was postulated to be potentially due to the equilibrium phase prior to the initiation of photocatalysis. These findings are consistent with previous studies that have demonstrated the inhibition rates of Se4+@TiO2/PET against S. aureus and E. coli were 99.99% and 98.47%, respectively, with significant antibacterial effects [64,65]. This indicates that the Se4+@TiO2/PET material exerts its antimicrobial activity by stimulating the highly active ROS generated by the Se4+@TiO2 composite catalyst, resulting in irreversible damage to microbial cell walls and membranes, leading to leakage of microbial cell contents, impaired cellular functions, and ultimately inactivation and death of the microorganisms [66,67,68].

2.10. The Photocatalytic Antibacterial Reaction Mechanism Analysis

Figure 12 presents a diagram of the antimicrobial mechanism of UV–visible photocatalytic degradation of Se4+@TiO2/PET. This diagram can explain the principle of degradation of MO and inhibition of bacteria by the material. When Se4+@TiO2 is exposed to light with an adsorption energy greater than or equal to its band gap (Eg), electrons ( e C B ) filling the valence band (VB) are excited and jump into the empty conduction band (CB), forming holes ( h V B + ) in the VB (Equation (6)). These excited electrons and holes are usually called carriers. Most of the electron–hole pairs produced by absorbing light recombine to produce emission energy (Equation (7)). In contrast, when the light-generated carriers do not recombine and the electron–hole pairs separate, the electrons and holes are transferred to the catalyst surface, where they are captured by the surface-active sites and drive the redox reaction. The entire photocatalytic reaction can be decomposed into two half-reactions: electron-induced reduction and hole-induced oxidation. Photogenerated electrons can react with dissolved oxygen molecules (O2) in aqueous solution to form superoxide radical anions (O2·) (Equation (8)), while holes can interact with H2O molecules on the surface of catalyst particles to form hydroxyl radicals (·OH) (Equation (9)). All of these generators are collectively referred to as highly reactive oxidants (ROS), which can be further involved in the oxidative degradation of organic pollutants by mineralizing them into CO2 and H2O (Equations (10) and (11)) [69,70]. In addition, it is also important to consider that the presence of ROS may lead to the damage of the molecular as well as organic polymer matrix.
The energy level of Se4+ (2.27 eV) is a determining factor in the high photocatalytic activity of the composite. When the energy level of the Se4+doped ion lies below the conduction band edge, it traps the excited electrons; when the energy level is above the valence band edge, the electrons can burst the photogenerated holes, the center of which can act as an electron or hole trap, thus temporarily separating the light-generated carriers [34]. Furthermore, the photocatalytic reaction is proportional to the number of photons absorbed by the photocatalyst. The doping of titanium dioxide with Se4+ has been observed to enhance the light-absorbing properties of the material. This phenomenon is associated with an increase in the generation of electron and hole pairs on the surface of Se4+@TiO2/PET, which in turn facilitates the participation of these charge carriers in redox reactions. This process has been demonstrated to enhance the degradation efficiency of methyl orange [71].
S e 4 + @ T i O 2 + h v e C B + h V B +
e C B + h V B + e n e r g y
O 2 + e C B O 2 ·
H 2 O + h V B + · O H + H +
O 2 · + M O C O 2 + H 2 O
· O H + M O C O 2 + H 2 O
Several reports have confirmed that bacteria can be decomposed and mineralized by TiO2 during the photocatalytic process [72,73]. TiO2 has the capacity to degrade the impurities into less toxic terminal compounds or even mineralize all of them into CO2 and H2O. While selenium itself has high antimicrobial properties [74], the incorporation of Se4+ into TiO2 will enhance the antibacterial performance of the composite photocatalytic materials. The bacterial inhibition mechanisms of photocatalytic antimicrobial composites can be broadly classified into three categories [75]: (a) biological processes, which lead to the internalization of the nanoparticles in bacteria through ion channels or proteins in the cell wall; (b) physical processes, which include the adsorption of photocatalytic antimicrobial composites on the cell surface; and (c) chemical phenomena, which are either generated by ROS or due to the removal from the toxic effects of ionic substances leached from the nanoparticles. Upon direct contact between Se4+@TiO2/PET and the bacterial surface, the peptidoglycan of the bacterial cell wall was destroyed by the ROS generated in the photocatalytic reaction. Additionally, the cell membrane components were peroxidized during the photocatalytic process, which led to an increase in the permeability of the cell membrane, destroying its integrity. Ultimately, this resulted in the inactivation of the bacteria and further decomposition of the bacterial residues (Equation (12)) [76].
· O H + B a c t e r i u m C O 2 + H 2 O

3. Materials and Methods

3.1. Materials

The polyester knitted fabric was supplied by Jiangsu Kuangda Technology Group Co., Ltd. (Changzhou, China) with a grammage weight of 230 g/m2. The reagents used in this study were titanium dioxide (TiO2, Aladdin Biochemical Technology Co., Ltd., Shanghai, China), glacial acetic acid (Hubao Chemical Reagent Co., Ltd., Yangzhou, China), tetrabutyl titanate (Yuanye Biotechnology Co., Ltd., Shanghai, China), selenium dioxide (SeO2, McLean Biochemical Ltd., Shanghai, China), anhydrous ethanol (Runjie Chemical Reagent Co., Ltd., Shanghai, China), acetone (Lingfeng Chemical Reagent Co., Ltd., Shanghai, China), and methyl orange (Yuanye Biotechnology Co., Ltd., Shanghai, China), all of which were of analytical grade. Deionized water was used for all experiments.

3.2. Preparation of Se4+@TiO2 Composite Photocatalysts

First, 20 mL of anhydrous ethanol was mixed with 5 mL of glacial acetic acid under magnetic stirring at 600 r/min, and 10 mL of tetrabutyl titanate was added drop by drop and recorded as solution A. The solution was mixed with 5 mL of deionized water and 0.33 g of SeO2 under magnetic stirring until SeO2 was completely dissolved. Then 20 mL of anhydrous ethanol was mixed with 0.5 mL of deionized water and 0.33 g of SeO2 with magnetic stirring until the SeO2 was completely dissolved, recorded as solution B. Solution B was mixed with 0.5 mL of deionized water and 0.33 g of SeO2 with magnetic stirring until the SeO2 was completely dissolved. Solution B was added into solution A under magnetic stirring at 1500 r/min, the resulting sol was aged for 12 h to obtain a wet gel, and the wet gel was dried at 80 °C for 24 h to obtain a dry gel. Subsequently, the dry gel was pulverized and calcined in a resistance furnace at 500 °C for a period of 2 h, resulting in the production of a photocatalyst doped with a 10% molar ratio of Se4+ to TiO2.

3.3. Preparation of Se4+@TiO2-Loaded PET Fabrics

The PET fabrics were dispersed in a 1:1:1 mixture of ethanol, acetone, and deionized water at room temperature for 30 min using a 70 W model KH-250DE ultrasonic disperser produced by Kunshan Hexiang Ultrasonic Instrument Co. Ltd. (Kunshan, China), and then placed into the dispersed solution and shaken in a water bath at 60 °C for 2 h. The PET fabrics were taken out and rinsed with deionized water more than 5 times, and dried at 60 °C. The pretreated PET fabrics were cut into 10 cm × 10 cm and placed into a model PDC-VCG-2 plasma chamber manufactured by Harrick Scientific Products, Inc. (New York, NY, USA) for 2 min; the plasma power supply voltage was 220 V, frequency was 50 Hz, and the processing power was 18 W. Then 0.006 mol of Se4+@TiO2 or TiO2 photocatalyst was dispersed in 50 mL of deionized water by ultrasonication for 30 min, respectively, followed by the addition of 1 g of plasma-treated PET fabrics to ensure the identical loading of photocatalysts deposited on the fibers. After that, the suspension was oscillated in a water bath for 2 h. The PET fabrics were then removed and rinsed more than 5 times with deionized water and dried at 60 °C.

3.4. Characterization

Various analytical techniques were carried out on the experimental samples in order to investigate their surface morphology, crystal structure, chemical composition, optical characteristics, band gap width, photocatalytic activity, and antimicrobial properties. Specifically, the surface morphology of the materials was analyzed using a ZEISS Gemini SEM 300, Oberkochen, Germany, field emission scanning electron microscope (SEM) with an accelerating voltage of 5 kV. The elemental composition of the material surface was analyzed using a Genesis XM series X-ray energy dispersive spectrometer (EDS) on a Carl Zeiss EVO15 instrument from Oberkochen, Germany. The crystalline phase structure of the materials was analyzed using an X-ray diffractometer (XRD) model H-12, manufactured by Rigaku Corporation, Tokyo, Japan, operated at a voltage of 40 kV and a current of 100 mA with a 2θ scanning range of 10° to 80°. An X-ray photoelectron spectrometer (XPS) model K-Alpha from Thermo Fisher Scientific Inc., Waltham, MA, USA was used to analyze the elemental species, chemical composition, and information about the electronic structure of the material surfaces, and the elements tested were C, Ti, O, and Se. A UV–visible diffuse reflectance spectrophotometer (UV–Vis) model UH4150 from Hitachi Ltd., Tokyo, Japan was used to analyze the UV–Vis absorption spectra of the samples with the wavelength of 200–800 nm at 1 nm intervals at a scanning speed of 1200 nm/min. The fluorescence spectra of the samples were measured using a full-featured steady state/transient fluorescence spectrometer (PL) model FLS980 from Edinburgh Instruments Ltd., Livingston, UK, under 450 W xenon lamp irradiation. The electrochemical impedance spectroscopy (EIS) test was conducted using a CHI760 electrochemical workstation manufactured by Shanghai Chenhua Instrument Co. Ltd., Shanghai, China. The test was carried out in a standard three-electrode system in the frequency range of 0.1–105 Hz with an amplitude of 10 mV, and the electrolyte was a 1 mol/L KOH solution.

3.5. Radical Trapping Analysis

The presence of the hydroxyl radical (·OH), superoxide anion (O2·), and hole (h+) reactive species during the photocatalytic reaction was confirmed through the addition of 10 mM of IPA (a scavenger of ·OH), 1 mM of BQ (a scavenger of O2·), and 10 mM of EDTA-2Na (a scavenger of h+). Subsequently, the degradation efficiency of Se4+@TiO2/PET on methyl orange was employed to ascertain the species responsible for dye degradation and bacterial inactivation.

3.6. Photocatalytic Degradation Performance Reaction Test

The methyl orange (MO) at a concentration of 20 mg/L was prepared as a model for organic pollutants. The composite photocatalytic materials were cut into pieces, submerged in 100 mL of the prepared MO, and stirred magnetically for 30 min in the dark to reach the adsorption equilibrium between the materials and MO. A PLS-SXE300+ xenon lamp with a power of 300 W and a wavelength of 320–780 nm manufactured by Beijing Porphyry Technology Co. Ltd. (Beijing, China) was used to simulate the sunlight source. The absorbance of the supernatant was measured at 20 min intervals and recorded. According to the Beer–Lambert law, there is a linear relationship between dye concentration and absorbed light, and the degradation efficiency of the photocatalytic material was calculated according to Equation (13) to analyze its simulated solar irradiation photocatalytic performance [77]:
η = C 0 C t C 0 × 100 % = A 0 A t A 0 × 100 %
where η is the degradation efficiency of the material, %; C0 is the concentration of MO at the starting instant, mg/L; Ct is the concentration of MO at the instant t, mg/L; A0 is the absorbance of MO at the starting instant; At is the absorbance of MO at the instant t.
To evaluate the reusability of the Se4+@TiO2/PET composite, the Se4+@TiO2/PET was recycled, collected, washed, and dried. Subsequently, it was carried out three times under visible light irradiation.

3.7. Evaluation of Antimicrobial Performance

The antimicrobial properties of the materials were tested according to GB/T 20944.3–2008 [78] using the oscillation method. The test strains were gram-positive bacteria S.aureus and gram-negative bacteria E.coli [79]. The procedure was as follows: 0.75 g of a 0.5 cm × 0.5 cm sample was placed in a conical flask containing 70 mL of 0.03 mol/L phosphate buffered saline and 5 mL of diluted bacterial solution and incubated on a shaker at 24 °C for 24 h to form a culture solution. Then 1 mL of the diluted culture solution was evenly distributed in Petri dishes containing agar and then incubated in an inverted incubator at 37 °C for 24 h. At the end of the incubation period, the colonies produced were counted and the inhibition rate was calculated according to Equation (14) [80]:
Y = N b N t N b × 100 %
where Y is the inhibition rate of the sample, %; Nb is the number of colonies of the standard blank sample; Nt is the number of colonies of other antimicrobial samples.

4. Conclusions

The Se4+@TiO2 photocatalytic material was prepared by the sol–gel method, and the plasma technique was employed to introduce free radicals on the surface of polyester fabric. This was performed to provide active sites for Se4+@TiO2 and to enhance its binding fastness to polyester fabric, thus preparing the Se4+@TiO2/PET composite photocatalytic material with an enhanced photocatalytic efficiency. The XRD results indicated that Se4+ had entered the substitution site of the TiO2 crystal structure in the composite photocatalytic material. The XPS results revealed that the valence state of Se was +4, and the shift of Ti 2p binding energy also indicated that Se4+ was doped into the TiO2 lattice. The PL intensity of the Se4+@TiO2/PET composite photocatalytic material was lower than that of TiO2/PET, suggesting a higher photocatalytic activity. In the UV–Vis spectrum, the forbidden band gap of TiO2/PET was 3.1 eV, while that of the Se4+@TiO2/PET composite was 2.9 eV. The bandgap of Se4+@TiO2/PET narrowed, the absorption threshold shifted to the visible light region, and the range of light utilization broadened. These observations indicate that Se4+ doping of TiO2 enhanced the simulated sunlight photocatalytic activity of composite photocatalytic materials. The photocatalytic and antimicrobial results demonstrated that Se4+@TiO2/PET exhibited a superior degradation efficiency of 81% for methyl orange solution in comparison to TiO2/PET. The antibacterial properties of Se4+@TiO2/PET were subsequently investigated, achieving 99.99% and 98.47% inhibition against S. aureus and E. coli, respectively. Nevertheless, the ROS generated by the photocatalyst have the potential to cause damage to the organic polymer carrier and degrade the matrix.

Author Contributions

Conceptualization, methodology, and validation, Y.R., M.L. and Y.L.; formal analysis, Y.R., R.L. and Z.Z.; writing—original draft preparation, Y.R., R.L. and Z.Z.; writing—review and editing, Y.R., R.L., L.T. and C.W.; supervision, project administration, and funding acquisition, Y.R., M.L. and Y.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by Natural Science Foundation of Jiangsu Province (No. BK20220613), Postgraduate Research & Practice Innovation Program of Jiangsu Province (No. KYCX24_3528), the Nantong Science and Technology Project (No. JC12022080), Student Innovation Training Program of Nantong University (No. 2024_173), and the Large Instruments Open Foundation of Nantong University (No. KFJN2459).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Niu, L.; Hu, Y.; Hu, H.; Zhang, X.; Wu, Y.; Giwa, A.S.; Huang, S. Kitchen-waste-derived biochar modified nanocomposites with improved photocatalytic performances for degrading organic contaminants. Environ. Res. 2022, 214, 114068. [Google Scholar] [CrossRef] [PubMed]
  2. Jarvin, M.; Rosaline, D.R.; Gopalakrishnan, T.; Kamalam, M.B.R.; Foletto, E.L.; Dotto, G.L.; Inbanathan, S.S.R. Remarkable photocatalytic performances towards pollutant degradation under sunlight and enhanced electrochemical properties of TiO2/polymer nanohybrids. Environ. Sci. Pollut. Res. 2023, 30, 62832–62846. [Google Scholar] [CrossRef] [PubMed]
  3. Liu, Y.; Xu, J.; Ding, Z.; Mao, M.; Li, L. Marigold shaped mesoporous composites Bi2S3/Ni(OH)2 with n-n heterojunction for high efficiency photocatalytic hydrogen production from water decomposition. Chem. Phys. Lett. 2021, 766, 138337. [Google Scholar] [CrossRef]
  4. Valverde-González, A.; López Calixto, C.G.; Barawi, M.; Gomez-Mendoza, M.; de la Peña O’Shea, V.A.; Liras, M.; Gómez-Lor, B.; Iglesias, M. Understanding charge transfer mechanism on effective truxene-based porous polymers–TiO2 hybrid photocatalysts for hydrogen evolution. ACS Appl. Energ. Mater. 2020, 3, 4411–4420. [Google Scholar] [CrossRef]
  5. Ma, Y.; Yi, X.; Wang, S.; Li, T.; Tan, B.; Chen, C.; Majima, T.; Waclawik, E.R.; Zhu, H.; Wang, J. Selective photocatalytic CO2 reduction in aerobic environment by microporous Pd-porphyrin-based polymers coated hollow TiO2. Nat. Commun. 2022, 13, 1400. [Google Scholar] [CrossRef]
  6. Li, N.; Chen, X.; Wang, J.; Liang, X.; Ma, L.; Jing, X.; Chen, D.-L.; Li, Z. ZnSe nanorods–CsSnCl3 perovskite heterojunction composite for photocatalytic CO2 reduction. ACS Nano 2022, 16, 3332–3340. [Google Scholar] [CrossRef] [PubMed]
  7. Yang, R.; Song, G.; Wang, L.; Yang, Z.; Zhang, J.; Zhang, X.; Wang, S.; Ding, L.; Ren, N.; Wang, A.; et al. Full solar-spectrum-driven antibacterial therapy over hierarchical Sn3O4/PDINH with enhanced photocatalytic activity. Small 2021, 17, 2102744. [Google Scholar] [CrossRef]
  8. Shi, L.; Ma, Z.; Qu, W.; Zhou, W.; Deng, Z.; Zhang, H. Hierarchical Z-scheme Bi2S3/CdS heterojunction: Controllable morphology and excellent photocatalytic antibacterial. Appl. Surf. Sci. 2021, 568, 150923. [Google Scholar] [CrossRef]
  9. Alamelu, K.; Raja, V.; Shiamala, L.; Jaffar Ali, B.M. Biphasic TiO2 nanoparticles decorated graphene nanosheets for visible light driven photocatalytic degradation of organic dyes. Appl. Surf. Sci. 2018, 430, 145–154. [Google Scholar] [CrossRef]
  10. Pant, B.; Park, M.; Park, S.-J. Recent Advances in TiO2 films prepared by sol-gel methods for photocatalytic degradation of organic pollutants and antibacterial activities. Coatings 2019, 9, 613. [Google Scholar] [CrossRef]
  11. Pant, B.; Pant, H.R.; Barakat, N.A.M.; Park, M.; Jeon, K.; Choi, Y.; Kim, H.-Y. Carbon nanofibers decorated with binary semiconductor (TiO2/ZnO) nanocomposites for the effective removal of organic pollutants and the enhancement of antibacterial activities. Ceram. Int. 2013, 39, 7029–7035. [Google Scholar] [CrossRef]
  12. Verma, V.; Al-Dossari, M.; Singh, J.; Rawat, M.; Kordy, M.G.M.; Shaban, M. A review on green synthesis of TiO2 NPs: Photocatalysis and antimicrobial applications. Polymers 2022, 14, 1444. [Google Scholar] [CrossRef] [PubMed]
  13. Pant, B.; Barakat, N.A.M.; Pant, H.R.; Park, M.; Saud, P.S.; Kim, J.-W.; Kim, H.-Y. Synthesis and photocatalytic activities of CdS/TiO2 nanoparticles supported on carbon nanofibers for high efficient adsorption and simultaneous decomposition of organic dyes. J. Colloid Interface Sci. 2014, 434, 159–166. [Google Scholar] [CrossRef]
  14. Ren, Y.; Zhao, Z.; Fan, T.; Luan, R.; Yao, L.; Shen, H.; Hu, X.; Cui, L.; Li, M.-X. Chitosan and TiO2 functionalized polypropylene nonwoven fabrics with visible light induced photocatalytic antibacterial performances. Int. J. Biol. Macromol. 2023, 250, 126305. [Google Scholar] [CrossRef] [PubMed]
  15. Pant, B.; Park, M.; Park, S.-J. Hydrothermal synthesis of Ag2CO3-TiO2 loaded reduced graphene oxide nanocomposites with highly efficient photocatalytic activity. Chem. Eng. Commun. 2019, 207, 688–695. [Google Scholar] [CrossRef]
  16. Dinh, C.T.; Yen, H.; Kleitz, F.; Do, T.O. Three-Dimensional ordered assembly of thin-shell Au/TiO2 hollow nanospheres for enhanced visible-light-driven photocatalysis. Angew. Chemie. Int. Ed. 2014, 53, 6618–6623. [Google Scholar] [CrossRef]
  17. Komaraiah, D.; Radha, E.; Sivakumar, J.; Reddy, M.V.R.; Sayanna, R. Influence of Fe3+ ion doping on the luminescence emission behavior and photocatalytic activity of Fe3+, Eu3+-codoped TiO2 thin films. J. Alloys Compd. 2021, 868, 159109. [Google Scholar] [CrossRef]
  18. Khan, H.; Swati, I.K. Fe3+-doped anatase TiO2 with d–d transition, oxygen vacancies and Ti3+ centers: Synthesis, characterization, UV–vis photocatalytic and mechanistic studies. Ind. Eng. Chem. Res. 2016, 55, 6619–6633. [Google Scholar] [CrossRef]
  19. Radha, E.; Komaraiah, D.; Sayanna, R.; Sivakumar, J. Photoluminescence and photocatalytic activity of rare earth ions doped anatase TiO2 thin films. J. Lumines. 2022, 244, 118727. [Google Scholar] [CrossRef]
  20. Tan, X.; Zhang, J.; Tan, D.; Shi, J.; Cheng, X.; Zhang, F.; Liu, L.; Zhang, B.; Su, Z.; Han, B. Ionic liquids produce heteroatom-doped Pt/TiO2 nanocrystals for efficient photocatalytic hydrogen production. Nano Res. 2019, 12, 1967–1972. [Google Scholar] [CrossRef]
  21. Stucchi, M.; Bianchi, C.L.; Argirusis, C.; Pifferi, V.; Neppolian, B.; Cerrato, G.; Boffito, D.C. Ultrasound assisted synthesis of Ag-decorated TiO2 active in visible light. Ultrason. Sonochem. 2018, 40, 282–288. [Google Scholar] [CrossRef] [PubMed]
  22. Liao, X.; Li, T.-T.; Ren, H.-T.; Mao, Z.; Zhang, X.; Lin, J.-H.; Lou, C.-W. Enhanced photocatalytic performance through the ferroelectric synergistic effect of p-n heterojunction BiFeO3/TiO2 under visible-light irradiation. Ceram. Int. 2021, 47, 10786–10795. [Google Scholar] [CrossRef]
  23. Pan, X.; Xu, Y.-J. Graphene-templated bottom-up fabrication of ultralarge binary CdS–TiO2 nanosheets for photocatalytic selective reduction. J. Phys. Chem. C 2015, 119, 7184–7194. [Google Scholar] [CrossRef]
  24. Yadav, H.M.; Kim, J.-S.; Pawar, S.H. Developments in photocatalytic antibacterial activity of nano TiO2: A review. Korean J. Chem. Eng. 2016, 33, 1989–1998. [Google Scholar] [CrossRef]
  25. Huang, F.; Yan, A.; Zhao, H. Influences of doping on photocatalytic properties of TiO2 photocatalyst. In Semiconductor Photocatalysis—Materials, Mechanisms and Applications; InTechOpen: Rijeka, Croatia, 2016. [Google Scholar] [CrossRef]
  26. Shi, J.; Chen, G.; Zeng, G.; Chen, A.; He, K.; Huang, Z.; Hu, L.; Zeng, J.; Wu, J.; Liu, W. Hydrothermal synthesis of graphene wrapped Fe-doped TiO2 nanospheres with high photocatalysis performance. Ceram. Int. 2018, 44, 7473–7480. [Google Scholar] [CrossRef]
  27. Zhu, J.; Chen, F.; Zhang, J.; Chen, H.; Anpo, M. Fe3+-TiO2 photocatalysts prepared by combining sol–gel method with hydrothermal treatment and their characterization. J. Photochem. Photobio. A Chem. 2006, 180, 196–204. [Google Scholar] [CrossRef]
  28. Zhou, Y.; Zhang, Q.; Shi, X.; Song, Q.; Zhou, C.; Jiang, D. Photocatalytic reduction of CO2 into CH4 over Ru-doped TiO2: Synergy of Ru and oxygen vacancies. J. Colloid Interface Sci. 2022, 608, 2809–2819. [Google Scholar] [CrossRef]
  29. Vidhya, R.; Malliga, P.; Karthikeyan, R.; Neyvasagam, K. Influence of nickel dopant concentration on structural, optical, magnetic and electrochemical properties of TiO2 nanoparticle. Int. J. Nanotechnol. 2021, 18, 610–621. [Google Scholar] [CrossRef]
  30. Xiang, H.; Tuo, B.; Tian, J.; Hu, K.; Wang, J.; Cheng, J.; Tang, Y. Preparation and photocatalytic properties of Bi-doped TiO2/montmorillonite composite. Opt. Mater. 2021, 117, 111137. [Google Scholar] [CrossRef]
  31. Zhang, D.; Chen, J.; Xiang, Q.; Li, Y.; Liu, M.; Liao, Y. Transition-metal-ion (Fe, Co, Cr, Mn, Etc.) doping of TiO2 nanotubes: A general approach. Inorg. Chem. 2019, 58, 12511–12515. [Google Scholar] [CrossRef]
  32. Mancuso, A.; Sacco, O.; Vaiano, V.; Sannino, D.; Pragliola, S.; Venditto, V.; Morante, N. Visible light active Fe-Pr co-doped TiO2 for water pollutants degradation. Catal. Today 2021, 380, 93–104. [Google Scholar] [CrossRef]
  33. Yalçın, Y.; Kılıç, M.; Çınar, Z. Fe+3-doped TiO2: A combined experimental and computational approach to the evaluation of visible light activity. Appl. Catal. B Environ. 2010, 99, 469–477. [Google Scholar] [CrossRef]
  34. Li, Y.; Ma, X.; Zhang, J.; Pan, X.; Li, N.; Chen, G.; Zhu, J. Degradable selenium-containing polymers for low cytotoxic antibacterial materials. ACS Macro Lett. 2022, 11, 1349–1354. [Google Scholar] [CrossRef]
  35. Huang, Y.; Chen, Q.; Zeng, H.; Yang, C.; Wang, G.; Zhou, L. A review of selenium (Se) nanoparticles: From synthesis to applications. Part. Part. Syst. Charact. 2023, 40, 2300098. [Google Scholar] [CrossRef]
  36. Cheng, H.; Zhang, M.; Hu, H.; Gong, Z.; Zeng, Y.; Chen, J.; Zhu, Z.; Wan, Y. Selenium-modified TiO2 nanoarrays with antibacterial and anticancer properties for postoperation therapy applications. ACS Appl. Bio Mater. 2018, 1, 1656–1666. [Google Scholar] [CrossRef]
  37. Lin, Z.H.; Roy, P.; Shih, Z.Y.; Ou, C.M.; Chang, H.T. Synthesis of anatase Se/Te-TiO2 nanorods with dominant {100} facets: Photocatalytic and antibacterial activity induced by visible light. ChemPlusChem 2013, 78, 302–309. [Google Scholar] [CrossRef]
  38. Gurkan, Y.Y.; Kasapbasi, E.; Cinar, Z. Enhanced solar photocatalytic activity of TiO2 by selenium(IV) ion-doping: Characterization and DFT modeling of the surface. Chem. Eng. J. 2013, 214, 34–44. [Google Scholar] [CrossRef]
  39. Xie, W.; Li, R.; Xu, Q. Enhanced photocatalytic activity of Se-doped TiO2 under visible light irradiation. Sci. Rep. 2018, 8, 8752. [Google Scholar] [CrossRef]
  40. Magnone, E.; Kim, M.K.; Lee, H.J.; Park, J.H. Facile synthesis of TiO2-supported Al2O3 ceramic hollow fiber substrates with extremely high photocatalytic activity and reusability. Ceram. Int. 2021, 47, 7764–7775. [Google Scholar] [CrossRef]
  41. Rashid, M.M.; Simončič, B.; Tomšič, B. Recent advances in TiO2-functionalized textile surfaces. Surf. Interfaces 2021, 22, 100890. [Google Scholar] [CrossRef]
  42. Wang, Q.; Li, J.; Bai, Y.; Lu, X.; Ding, Y.; Yin, S.; Huang, H.; Ma, H.; Wang, F.; Su, B. Photodegradation of textile dye Rhodamine B over a novel biopolymer–metal complex wool-Pd/CdS photocatalysts under visible light irradiation. J. Photochem. Photobiol. B Biol. 2013, 126, 47–54. [Google Scholar] [CrossRef] [PubMed]
  43. D’Amato, C.; Giovannetti, R.; Zannotti, M.; Rommozzi, E.; Minicucci, M.; Gunnella, R.; Di Cicco, A. Band gap implications on nano-TiO2 surface modification with ascorbic acid for visible light-active polypropylene coated photocatalyst. Nanomaterials 2018, 8, 599. [Google Scholar] [CrossRef] [PubMed]
  44. Min, K.S.; Manivannan, R.; Son, Y.-A. Porphyrin Dye/TiO2 imbedded PET to improve visible-light photocatalytic activity and organosilicon attachment to enrich hydrophobicity to attain an efficient self-cleaning material. Dyes Pigm. 2019, 162, 8–17. [Google Scholar] [CrossRef]
  45. Alotaibi, A.M.; Williamson, B.A.D.; Sathasivam, S.; Kafizas, A.; Alqahtani, M.; Sotelo-Vazquez, C.; Buckeridge, J.; Wu, J.; Nair, S.P.; Scanlon, D.O.; et al. Enhanced photocatalytic and antibacterial ability of Cu-doped anatase TiO2 thin films: Theory and experiment. ACS Appl. Mater. Interfaces 2020, 12, 15348–15361. [Google Scholar] [CrossRef]
  46. Khan, H.; Berk, D. Selenium modified oxalate chelated titania: Characterization, mechanistic and photocatalytic studies. Appl. Catal. A Gen. 2015, 505, 285–301. [Google Scholar] [CrossRef]
  47. Ahmad, D.; Manzoor, M.Z.; Kousar, R.; Somaily, H.H.; Buzdar, S.A.; Ullah, H.; Nazir, A.; Warsi, M.F.; Batool, Z. Cost effective synthesis of CuxBi2-xSe3 photocatalysts by sol-gel method and their enhanced photodegradation and antibacterial activities. Ceram. Int. 2022, 48, 32787–32797. [Google Scholar] [CrossRef]
  48. Wu, C.; Corrigan, N.; Lim, C.-H.; Jung, K.; Zhu, J.; Miyake, G.; Xu, J.; Boyer, C. Guiding the design of organic photocatalyst for PET-RAFT polymerization: Halogenated xanthene dyes. Macromolecules 2018, 52, 236–248. [Google Scholar] [CrossRef]
  49. Huang, J.; Dou, L.; Li, J.; Zhong, J.; Li, M.; Wang, T. Excellent visible light responsive photocatalytic behavior of N-doped TiO2 toward decontamination of organic pollutants. J. Hazard. Mater. 2021, 403, 123857. [Google Scholar] [CrossRef]
  50. Wu, T.; Zhao, H.; Zhu, X.; Xing, Z.; Liu, Q.; Liu, T.; Gao, S.; Lu, S.; Chen, G.; Asiri, A.M.; et al. Identifying the origin of Ti3+ activity toward enhanced electrocatalytic N2 reduction over TiO2 nanoparticles modulated by mixed-valent copper. Adv. Mater. 2020, 32, e2000299. [Google Scholar] [CrossRef]
  51. Wu, S.X.; Ma, Z.; Qin, Y.N.; He, F.; Jia, L.S.; Zhang, Y.J. XPS study of copper doping TiO2 photocatalyst. Acta Phys.-Chim. Sin. 2003, 19, 967–969. [Google Scholar] [CrossRef]
  52. Mangayayam, M.; Kiwi, J.; Giannakis, S.; Pulgarin, C.; Zivkovic, I.; Magrez, A.; Rtimi, S. FeOx magnetization enhancing E. coli inactivation by orders of magnitude on Ag-TiO2 nanotubes under sunlight. Appl. Catal. B Environ. 2017, 202, 438–445. [Google Scholar] [CrossRef]
  53. Shan, C.; Su, Z.; Liu, Z.; Xu, R.; Wen, J.; Hu, G.; Tang, T.; Fang, Z.; Jiang, L.; Li, M. One-step synthesis of Ag2O/Fe3O4 magnetic photocatalyst for efficient organic pollutant removal via wide-spectral-response photocatalysis–fenton coupling. Molecules 2023, 28, 4155. [Google Scholar] [CrossRef]
  54. Rao, Z.; Shi, G.; Wang, Z.; Mahmood, A.; Xie, X.; Sun, J. Photocatalytic degradation of gaseous VOCs over Tm3+-TiO2: Revealing the activity enhancement mechanism and different reaction paths. Chem. Eng. J. 2020, 395, 125078. [Google Scholar] [CrossRef]
  55. Murugesan, R.; Sivakumar, S.; Karthik, K.; Anandan, P.; Haris, M. Structural, optical and magnetic behaviors of Fe/Mn-doped and co-doped CdS thin films prepared by spray pyrolysis method. Appl. Phy. A 2019, 125, 281. [Google Scholar] [CrossRef]
  56. Zhang, H.Y.; Yu, H.; Dai, J.D.; Huang, X.Y.; Liu, C.M. Microstructure, photoluminescence and photocatalytic activity of ZnO-MoS2-TiO2 composite. Chin. J. Phys. 2018, 56, 3053–3061. [Google Scholar] [CrossRef]
  57. Kansal, S.K.; Lamba, R.; Mehta, S.K.; Umar, A. Photocatalytic degradation of Alizarin Red S using simply synthesized ZnO nanoparticles. Mater. Lett. 2013, 106, 385–389. [Google Scholar] [CrossRef]
  58. Kibombo, H.S.; Weber, A.S.; Wu, C.-M.; Raghupathi, K.R.; Koodali, R.T. Effectively dispersed europium oxide dopants in TiO2 aerogel supports for enhanced photocatalytic pollutant degradation. J. Photochem. Photobio. A Chem. 2013, 269, 49–58. [Google Scholar] [CrossRef]
  59. Qureshi, A.A.; Javed, H.M.A.; Javed, S.; Bashir, A.; Usman, M.; Akram, A.; Ahmad, M.I.; Ali, U.; Shahid, M.; Rizwan, M.; et al. Incorporation of Zr-doped TiO2 nanoparticles in electron transport layer for efficient planar perovskite solar cells. Surf. Interfaces 2021, 25, 101299. [Google Scholar] [CrossRef]
  60. Pattanayak, D.S.; Mishra, J.; Nanda, J.; Sahoo, P.K.; Kumar, R.; Sahoo, N.K. Photocatalytic degradation of cyanide using polyurethane foam immobilized Fe-TCPP-S-TiO2-rGO nano-composite. J. Environ. Manag. 2021, 297, 113312. [Google Scholar] [CrossRef]
  61. Enesca, A.; Isac, L. Tandem structures semiconductors based on TiO2_SnO2 and ZnO_SnO2 for photocatalytic organic pollutant removal. Nanomaterials 2021, 11, 200. [Google Scholar] [CrossRef]
  62. Chang, F.; Li, J.; Kou, Y.; Bao, W.; Shi, Z.; Zhu, G.; Kong, Y. The intense charge migration and efficient photocatalytic NO removal of the S-scheme heterojunction composites Bi7O9I3-BiOBr. Sep. Purif. Technol. 2025, 353, 128402. [Google Scholar] [CrossRef]
  63. Bai, Y.; Hu, X.; Tian, T.; Cai, B.; Li, Y. Carbon pre-protected iron strategy to construct Fe, C-codoped g-C3N4 for effective photodegradation of organic pollutants via hole oxidation mechanism. J. Cleaner Prod. 2024, 437, 140739. [Google Scholar] [CrossRef]
  64. Henriksen-Lacey, M.; Carregal-Romero, S.; Liz-Marzán, L.M. Current Challenges toward In vitro cellular validation of inorganic nanoparticles. Bioconjugate Chem. 2016, 28, 212–221. [Google Scholar] [CrossRef]
  65. Horie, M.; Sugino, S.; Kato, H.; Tabei, Y.; Nakamura, A.; Yoshida, Y. Does photocatalytic activity of TiO2 nanoparticles correspond to photo-cytotoxicity? Cellular uptake of TiO2 nanoparticles is important in their photo-cytotoxicity. Toxicol. Mech. Methods 2016, 26, 284–294. [Google Scholar] [CrossRef]
  66. Zhu, Z.; Cai, H.; Sun, D.W.; Wang, H.W. Photocatalytic effects on the quality of pork packed in the package combined with TiO2 coated nonwoven fabrics. J. Food Process Eng. 2019, 42, e12993. [Google Scholar] [CrossRef]
  67. Wang, W.; Li, G.; Xia, D.; An, T.; Zhao, H.; Wong, P.K. Photocatalytic nanomaterials for solar-driven bacterial inactivation: Recent progress and challenges. Environ. Sci.-Nano 2017, 4, 782–799. [Google Scholar] [CrossRef]
  68. Bozzi, A.; Yuranova, T.; Guasaquillo, I.; Laub, D.; Kiwi, J. Self-cleaning of modified cotton textiles by TiO2 at low temperatures under daylight irradiation. J. Photochem. Photobiol. A Chem. 2005, 174, 156–164. [Google Scholar] [CrossRef]
  69. Wang, G.; Cheng, H. Application of photocatalysis and sonocatalysis for treatment of organic dye wastewater and the synergistic effect of ultrasound and light. Molecules 2023, 28, 3706. [Google Scholar] [CrossRef] [PubMed]
  70. Guo, Q.; Zhou, C.; Ma, Z.; Yang, X. Fundamentals of TiO2 photocatalysis: Concepts, mechanisms, and challenges. Adv. Mater. 2019, 31, 1901997. [Google Scholar] [CrossRef]
  71. Wang, C.-Y.; Böttcher, C.; Bahnemann, D.W.; Dohrmann, J.K. A comparative study of nanometer sized Fe(iii)-doped TiO2photocatalysts: Synthesis, characterization and activity. J. Mater. Chem. 2003, 13, 2322–2329. [Google Scholar] [CrossRef]
  72. Amiri, M.R.; Alavi, M.; Taran, M.; Kahrizi, D. Antibacterial, antifungal, antiviral, and photocatalytic activities of TiO2 nanoparticles, nanocomposites, and bio-nanocomposites: Recent advances and challenges. J. Public Health Res. 2022, 11, 22799036221104151. [Google Scholar] [CrossRef]
  73. Uyguner Demirel, C.S.; Birben, N.C.; Bekbolet, M. A comprehensive review on the use of second generation TiO2 photocatalysts: Microorganism inactivation. Chemosphere 2018, 211, 420–448. [Google Scholar] [CrossRef]
  74. Zhang, T.; Qi, M.; Wu, Q.; Xiang, P.; Tang, D.; Li, Q. Recent research progress on the synthesis and biological effects of selenium nanoparticles. Front. Nutr. 2023, 10, 1183487. [Google Scholar] [CrossRef]
  75. Raksha, K.R.; Ananda, S.; Madegowda, N.M. Study of kinetics of photocatalysis, bacterial inactivation and • OH scavenging activity of electrochemically synthesized Se4+ doped ZnS nanoparticles. J. Mol. Catal. A Chem. 2015, 396, 319–327. [Google Scholar] [CrossRef]
  76. RokickA Konieczna, P.; Morawski, A.W. Photocatalytic bacterial destruction and mineralization by TiO2-based photocatalysts: A mini review. Molecules 2024, 29, 2221. [Google Scholar] [CrossRef] [PubMed]
  77. Dariani, R.S.; Esmaeili, A.; Mortezaali, A.; Dehghanpour, S. Photocatalytic reaction and degradation of methylene blue on TiO2 nano-sized particles. Optik 2016, 127, 7143–7154. [Google Scholar] [CrossRef]
  78. GB/T 20944.3-2008; Textiles—Evaluation for Antibacterial Activity—Part 3: Shake Flask Method. China National Standards: Beijing, China, 2008.
  79. Zhang, G.; Wang, D.; Yan, J.; Xiao, Y.; Gu, W.; Zang, C. Study on the photocatalytic and antibacterial properties of TiO2 nanoparticles-coated cotton fabrics. Materials 2019, 12, 2010. [Google Scholar] [CrossRef]
  80. Wang, Q.; Liu, S.; Lu, W.; Zhang, P. Fabrication of curcumin@Ag loaded core/shell nanofiber membrane and its synergistic antibacterial properties. Front. Chem. 2022, 10, 870666. [Google Scholar] [CrossRef]
Figure 1. Scanning electron microscopy images of composite photocatalytic materials: (a) untreated PET; (b) TiO2/PET; (c) Se4+@TiO2/PET.
Figure 1. Scanning electron microscopy images of composite photocatalytic materials: (a) untreated PET; (b) TiO2/PET; (c) Se4+@TiO2/PET.
Molecules 30 01306 g001
Figure 2. Dispersion of photocatalytic composite material elements: (a) untreated PET; (b) TiO2/PET; (c) Se4+@TiO2/PET.
Figure 2. Dispersion of photocatalytic composite material elements: (a) untreated PET; (b) TiO2/PET; (c) Se4+@TiO2/PET.
Molecules 30 01306 g002
Figure 3. X-ray diffraction pattern images of composite photocatalytic materials: (a) XRD full spectrum; (b) Diffraction peak of (101) crystal plane.
Figure 3. X-ray diffraction pattern images of composite photocatalytic materials: (a) XRD full spectrum; (b) Diffraction peak of (101) crystal plane.
Molecules 30 01306 g003
Figure 4. XPS image of TiO2/PET and Se4+@TiO2/PET: (a) wide spectrum; (b) C1s spectrum; (c) O1s spectrum; (d) Ti2p spectrum; (e) Se3d spectrum.
Figure 4. XPS image of TiO2/PET and Se4+@TiO2/PET: (a) wide spectrum; (b) C1s spectrum; (c) O1s spectrum; (d) Ti2p spectrum; (e) Se3d spectrum.
Molecules 30 01306 g004
Figure 5. (a) UV–vis absorption spectra of TiO2/PET and Se4+@TiO2/PET; (b) The relationship curve between (αhv)2 and hv of TiO2/PET and Se4+@TiO2/PET.
Figure 5. (a) UV–vis absorption spectra of TiO2/PET and Se4+@TiO2/PET; (b) The relationship curve between (αhv)2 and hv of TiO2/PET and Se4+@TiO2/PET.
Molecules 30 01306 g005
Figure 6. Photoluminescence spectra of TiO2/PET and Se4+@TiO2/PET.
Figure 6. Photoluminescence spectra of TiO2/PET and Se4+@TiO2/PET.
Molecules 30 01306 g006
Figure 7. Degradation curves of methyl orange by TiO2/PET and Se4+@TiO2/PET under simulated sunlight irradiation.
Figure 7. Degradation curves of methyl orange by TiO2/PET and Se4+@TiO2/PET under simulated sunlight irradiation.
Molecules 30 01306 g007
Figure 8. Degradation kinetics curve of methyl orange.
Figure 8. Degradation kinetics curve of methyl orange.
Molecules 30 01306 g008
Figure 9. The reusability of Se4+@TiO2/PET by degradation of methyl orange.
Figure 9. The reusability of Se4+@TiO2/PET by degradation of methyl orange.
Molecules 30 01306 g009
Figure 10. Electrochemical impedance spectra of TiO2/PET and Se4+@TiO2/PET.
Figure 10. Electrochemical impedance spectra of TiO2/PET and Se4+@TiO2/PET.
Molecules 30 01306 g010
Figure 11. Antibacterial activity of TiO2/PET and Se4+@TiO2/PET against E. coli and S. aureus.
Figure 11. Antibacterial activity of TiO2/PET and Se4+@TiO2/PET against E. coli and S. aureus.
Molecules 30 01306 g011
Figure 12. Se4+@TiO2/PET UV–Vis photocatalytic degradation and antimicrobial mechanism map.
Figure 12. Se4+@TiO2/PET UV–Vis photocatalytic degradation and antimicrobial mechanism map.
Molecules 30 01306 g012
Table 1. The constant of pseudo first-order kinetic equation for degradation of methyl orange under different photocatalysts.
Table 1. The constant of pseudo first-order kinetic equation for degradation of methyl orange under different photocatalysts.
SamplesKappR2
MO0.00070.98443
PET0.00140.99331
TiO2/PET0.002270.99609
Se4+@TiO2/PET0.01010.97474
Table 2. Degradation efficiency of Se4+@TiO2/PET for methyl orange under different radical scavengers.
Table 2. Degradation efficiency of Se4+@TiO2/PET for methyl orange under different radical scavengers.
Radical Scavenger BlankIPABQEDTA-2Na
Degradation efficiency %8177.2847.1172.64
Table 3. Antibacterial rates of TiO2/PET and Se4+@TiO2/PET against E. coli and S. aureus.
Table 3. Antibacterial rates of TiO2/PET and Se4+@TiO2/PET against E. coli and S. aureus.
SamplesConcentration of Live
E. coli Bacteria/(CFU·mL−1)
Inhibition
Rate (%)
Concentration of Live
S. aureus Bacteria/
(CFU·mL−1)
Inhibition
Rate (%)
Blank9.2 × 105n/a8.9 × 105n/a
TiO2/PET3.86 × 10558.041.04 × 10588.31
Se4+@TiO2/PET1.4 × 10498.47099.99
n/a: Not applicable.
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

Ren, Y.; Luan, R.; Zhao, Z.; Tang, L.; Wang, C.; Li, Y.; Li, M. Synthesis and Characterization of Se4+@TiO2/PET Composite Photocatalysts with Enhanced Photocatalytic Activity by Simulated Solar Irradiation and Antibacterial Properties. Molecules 2025, 30, 1306. https://doi.org/10.3390/molecules30061306

AMA Style

Ren Y, Luan R, Zhao Z, Tang L, Wang C, Li Y, Li M. Synthesis and Characterization of Se4+@TiO2/PET Composite Photocatalysts with Enhanced Photocatalytic Activity by Simulated Solar Irradiation and Antibacterial Properties. Molecules. 2025; 30(6):1306. https://doi.org/10.3390/molecules30061306

Chicago/Turabian Style

Ren, Yu, Rui Luan, Ziyao Zhao, Lina Tang, Chunxia Wang, Yuehui Li, and Meixian Li. 2025. "Synthesis and Characterization of Se4+@TiO2/PET Composite Photocatalysts with Enhanced Photocatalytic Activity by Simulated Solar Irradiation and Antibacterial Properties" Molecules 30, no. 6: 1306. https://doi.org/10.3390/molecules30061306

APA Style

Ren, Y., Luan, R., Zhao, Z., Tang, L., Wang, C., Li, Y., & Li, M. (2025). Synthesis and Characterization of Se4+@TiO2/PET Composite Photocatalysts with Enhanced Photocatalytic Activity by Simulated Solar Irradiation and Antibacterial Properties. Molecules, 30(6), 1306. https://doi.org/10.3390/molecules30061306

Article Metrics

Back to TopTop