Next Article in Journal
Scale Effects in Landslide Susceptibility Assessment: Integrating Slope Unit Division and SHAP-Based Interpretability in a Typical River Basin
Previous Article in Journal
A Coupled Model of System Dynamics and Environmental Models for the Development Process Deduction of the Yangtze River Basin: Model Construction Method
Previous Article in Special Issue
Integrated Groundwater Quality Assessment for Irrigation in the Ras El-Aioun District: Combining IWQI, GIS, and Machine Learning Approaches
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhancing Photocatalytic Activity with the Substantial Optical Absorption of Bi2S3-SiO2-TiO2/TiO2 Nanotube Arrays for Azo Dye Wastewater Treatment

1
Department of Chemistry, College of Science, King Khalid University, Abha 61421, Saudi Arabia
2
Department of Physics, Faculty of Sciences at Bisha, University of Bisha, P.O. Box 199, Bisha 61922, Saudi Arabia
3
Mining Research Center, Northern Border University, Arar 73213, Saudi Arabia
4
Department of Civil Engineering, College of Engineering, University of Diyala, Baqubah 10047, Iraq
5
Department of Chemical Engineering, College of Engineering, King Khalid University, Abha 61411, Saudi Arabia
6
Department of Civil Engineering Science, School of Civil Engineering, and the Built Environment, Faculty of Engineering and the Built Environment, University of Johannesburg, Kingsway Campus, P.O. Box 524, Aukland Park, Johannesburg 2006, South Africa
7
Directorate of Engineering the Future, School of Science, Engineering and Environment, The University of Salford, Newton Building, Greater Manchester M5 4WT, UK
8
Department of Town Planning, Engineering Networks and Systems, South Ural State University (National Research University), 76, Lenin Prospekt, Chelyabinsk 454080, Russia
9
Nexus by Sweden, Skepparbacken 5, 722 11 Västerås, Sweden
*
Author to whom correspondence should be addressed.
Water 2025, 17(13), 1875; https://doi.org/10.3390/w17131875
Submission received: 23 April 2025 / Revised: 10 June 2025 / Accepted: 18 June 2025 / Published: 24 June 2025
(This article belongs to the Special Issue Global Water Resources Management)

Abstract

One-dimensional TiO2 nanotube arrays (TNAs) were vertically aligned and obtained via the electrochemical anodization method. In this study, Bi2S3-TiO2-SiO2/TNA heterojunction photocatalysts were successfully prepared with different amounts of Bismuth(III) sulfide (Bi2S3) loading on the TNAs by the successive ionic layer adsorption and reaction (SILAR) method and characterized by X-ray diffraction (XRD) patterns, field-emission scanning electron microscope–energy-dispersive spectroscopy (FESEM-EDS), Fourier transform infrared (FTIR) spectra, ultraviolet-visible diffuse reflectance spectra (UV–Vis/DRS), and electrochemical impedance spectroscopy (EIS) techniques. The photocatalytic performances of the samples were investigated by degrading Basic Yellow 28 (BY 28) under visible-light irradiation. Optimization of the condition using the response surface methodology (RSM) and central composite rotatable design (CCRD) technique resulted in the degradation of BY 28 dye, showing that the catalyst with 9.6 mg/cm2 (designated as Bi2S3(9.6)-TiO2-SiO2/TNA) showed the maximum yield in the degradation process. The crystallite size of about 17.03 nm was estimated using the Williamson–Hall method. The band gap energies of TiO2-SiO2/TNA and Bi2S3(9.6)-TiO2-SiO2/TNA were determined at 3.27 and 1.87 eV for the direct electronic transitions, respectively. The EIS of the ternary system exhibited the smallest arc diameter, indicating an accelerated charge transfer rate that favors photocatalytic activity.

1. Introduction

Nowadays, due to the rapid development of various industries and population growth, environmental pollution has become one of the most substantial challenges of human society [1]. Utilizing solar energy, a clean, powerful, inexhaustible, affordable, and renewable energy source, to eliminate pollutants is regarded as the best solution to critical problems [2]. Wastewater colored with synthetic organic dye that contains nitrogen as the azo group (N=N) from various industries, especially the paper and textile industries, is essentially soluble in water and is considered the main source of pollution, causing significant environmental issues. Consequently, in today’s world, it appears to be essential to eliminate residual dyes in industrial effluents. Within the different groups of industrial wastewater dyes, basic dyes are visible at a low concentration and exhibit pronounced, vigorous color intensity [3,4]. BY28, a cationic dye that is hazardous to the environment, is frequently used for polyacrylonitrile and to color cotton [3].
Photocatalysts have been developed and used in environmental areas for industrial wastewater elimination [5]. The semiconductor titanium dioxide (TiO2) is widely used, being generally applied in heterogeneous photocatalysis, to remove pollution wastewater. However, due to its photocatalytic properties and inherent defects, the practical applications of TiO2 are significantly limited [6]. The rapid recombination of photogenerated electron and hole pairs leads to a lower rate of desired chemical transformations due to the absorbed light energy and low quantum efficiency [7,8,9]. It is worth noting that photocatalysts play an essential role in photocatalytic processes that regulate the efficiency of sunlight utilization and photocatalytic ability. Creating hybrid heterojunction designs and manufacturing photocatalysts based on semiconductors to increase photocatalytic activity have attracted much attention [10,11,12]. In a conventional type II heterojunction system, the electrons generated by light in the conduction band (CB) of a semiconductor with high reduction potential (i.e., more negative) can be transferred to the CB of a semiconductor with low reduction potential (i.e., less negative) due to more negative CB edge potentials. However, the resulting photoexcited electrons and holes are less active, in accordance with the lower oxidation capacity [13]. In a direct Z-scheme photocatalytic system, the photogenerated electrons on the CB of a semiconductor which has low reduction potential (i.e., less negative) will recombine with the photogenerated holes on the valance band (VB) of a semiconductor which has low oxidation potential (i.e., less positive), giving photogenerated electrons and holes high reduction and oxidation abilities for participating in photocatalytic redox reactions [9,11]. Notably, Z-scheme photocatalysts, compared to conventional heterojunction-type photocatalysts, have been designed to improve the separation efficiency of photogenerated charge carriers. They could play a critical role in improving charge transfer, oxidation potential, and photocatalytic efficiency [11]. This structure enables electrons and holes to collect on different semiconductors, extending the lifetime of the charge carriers and utilizing a broader range of light wavelengths, which ultimately increases light harvesting efficiency. For this purpose, Bi2S3 (band gap energy (Eg) of 1.27 eV)-based Z-scheme heterojunction semiconductors with more negative conduction edge potentials have gained attention for the production of TiO2-based Z-scheme photocatalysts [13], due to their high conduction potential, excellent photoactivity in visible light, and photosensitization, which thoroughly meets the requirements for the construction of photocatalysts based on Z-schematic heterojunctions [14,15,16,17].
For the photodegradation of organic dyes, researchers have resorted to constructing heterojunction systems as an emerging technology. According to the energy band matching principle, these systems are suitable for the fabrication of heterojunction photocatalysis in Bi2S3, which is excited in the visible region with a narrow bandgap energy, and semiconductors, like TiO2, which are activated in the ultraviolet region, which promotes the spatial separation of the excited electron/hole pairs [1,17]. Bai [18] focused on the solvothermal synthesis of a nanocomposite of Bi2S3 and TiO2 that had improved visible-light photocatalytic activity. Arjmand et al. [19] investigated the photocatalytic activity of TiO2/CNT/BiOBr/Bi2S3 in the degradation of Acid black 1721 under visible-light irradiation. As a result, complete decolorization was achieved at an initial pH of 3, a dye concentration of 30 mg/L, a reaction time of 40 min, and a catalyst dosage of 1 g/L, while COD removal was 99% after 120 min. Lin et al. [20] reported an improved photocatalytic performance of a Bi2S3-TiO2 nanotube array heterostructure under visible-light irradiation by sensitizing a wide-band-gap semiconductor with a narrow-band-gap semiconductor. Sun et al. [21] prepared Bi2S3/BiVO4/TiO2 heterostructure photocatalysts using continuous hydrothermal methods, and the photocatalytic activity of the sample was investigated by degrading methyl orange under visible-light irradiation. Compared to TiO2 nanorod arrays, Bi2S3/BiVO4/TiO2 photocatalysts extend the photocatalytic degradation efficiency of Bi2S3/BiVO4/TiO2 during the catalytic decomposition of methyl orange to 76.3%, which is approximately 4 times higher than that of bare TiO2 nanorod arrays. Drmosh et al. [22] synthesized a Bi2S3/MoS2/TiO2 heterostructure using a facile microwave-assisted hydrothermal method. The photocatalytic performance of the as-prepared nanocomposite was evaluated by monitoring the photocatalytic degradation of methylene blue under sunlight. The photocatalytic performance of Bi2S3/MoS2/TiO2 is much better than that of TiO2, MoS2, or Bi2S3. In this work, we investigated the performance of Bi2S3-SiO2-TiO2/TiO2 nanotube array heterojunction photocatalysts in degrading BY 28 under visible-light irradiation. The enhanced performance was mainly attributed to the synergistic effect of Bi2S3 and SiO2, which improved the dispersion of TNAs [2]. The introduction of SiO2 into the TiO2 system significantly enhances the stability of photocatalysts and the dispersivity. The supporting catalysts, using SiO2-TiO2 compared to single-component oxide supports, possess the benefits of significantly larger surface areas, enhanced acidic sites that improve redox abilities, and better degradation of pollutants [8]. Meanwhile, TNAs, which were vertically formed on the titanium (Ti) substrate by electrochemical anodization, were used as scaffolds to construct Bi2S3 and TiO2 nanoparticles to increase the photocatalytic performance of the photocatalyst. TNAs with smooth surfaces possess a one-dimensional (1D) geometry that facilitates charge carrier transfer. Also, paths along the TNA walls reduced charge carrier diffusion, thereby decreasing the occurrence of charge losses [6,7]. In addition, TNAs have shown better photocatalytic efficiency than other morphologies, possessing excellent properties, such as a high orientation; a large internal surface area, preventing the recombination of photogenerated electrons and holes under irradiation; and a short diffusion length to the conductive substrate for charge carrier transfer [8,9]. On the other hand, Bi2S3, which is advantageous for increasing efficient visible-light absorption, photogenerated carriers, and photocatalytic performance, also enhances electron–hole separation efficiency and a reduction in the recombination speed of electron–hole pairs, thereby improving the efficient capture of photogenerated charge carriers [23].
In this research, we report a novel photocatalyst, Bi2S3-TiO2-SiO2/TNA, developed using a successive ion layer adsorption and reaction (SILAR) technique to enhance the photocatalytic performance of the photocatalyst. Scheme 1 shows the experimental setup and strategy adopted for the preparation of Bi2S3-TiO2-SiO2/TNA. The samples were characterized, and their photocatalytic efficiency was tested for the degradation of BY 28 dye solution as a model of textile waste under visible light. Subsequently, the RSM-CCRD approach was employed to evaluate the interactions between the influencing experimental variables and to reduce the number of experiments needed to achieve the optimal photodegradation response. The effects of Bi2S3 loading on the TNAs, initial concentration of dye, and pH were investigated.

2. Experimental Section

2.1. Materials

The materials used in this study were of reagent grade, such as tetrabutylorthotitanate (Ti (OC4H9)4, 98%), acetyl acetone (C5H8O2, 99%), hydrochloric acid (HCl, 37%), ethanol (C2H5OH, 99%), tetraethylorthosilicate (TEOS, 99%), ammonium fluoride (NH4F, 99%), ethylene glycol (C2H6O2, 99%), and bismuth nitrate (Bi(NO3)3.5H2O, 99%). Titanium (Ti) plates (99.5% pure) were applied as a base metal with the dimensions of 1 × 4 cm and a thickness of 1 mm. All chemical reagents used in this work were purchased from Sigma–Aldrich (St. Louis, MO, USA) and Merck (Darmstadt, Germany).

2.2. Preparation of TNAs

A traditional electrochemical anodization process at a constant voltage was used to prepare the TNAs [20,24]. Firstly, Ti foils with 100- and 600-grit sandpaper were thoroughly polished, washed successively using acetone, isopropyl alcohol, and ethanol in ultrasound for 10 min, and dried in ambient air. Anodization experiments were carried out in an electrolyte bath consisting of 0.5 wt.% NH4F and 2.5 vol.% DI water in ethylene glycol [24]. The anodization was performed at a constant voltage of 60 V for 1 h, and the gap between the anode and cathode was 10 mm, while a Ti plate was chosen as the anode and platinum was used as the cathode. After the electrochemical process, the as-prepared samples were rinsed several times in DI water and then crystallized the amorphous nanotubes in the anatase phase and annealed in a furnace at a temperature of 500 °C, maintained for 2 h.

2.3. Fabrication of Bi2S3-TiO2-SiO2/TNA

TiO2-SiO2 was synthesized with the sol–gel technique, which is similar to previous reports. The details are as follows [25,26]. Firstly, a mixture of 10 mL of ethanol, 2.5 mL of tetrabutylorthotitanate, 2.5 mL of tetraethylorthosilicate, and 2.5 mL of acetylacetone was mixed at room temperature (23 °C) and stirred for 45 min; then, 2 mL of DI water was added to the reaction mixture, followed by stirring for 10 min. Subsequently, under continuous stirring for 2 h, the pH of the suspension was adjusted to 1.8. Finally, the mixture was aged at room temperature for 24 h.
TiO2-SiO2/TNA prepared by the immersion dip-coating method [27]. For this purpose, the TNAs were transferred into the TiO2-SiO2 sol solution, removed after 15 min, and dried in ambient air. The procedure of dipping, drying, and heating was repeated 3 times, for 30 s each time, and the covered TNAs were dried at 60 °C, for 2 h. Subsequently, the film was annealed by a furnace at a temperature of 500 °C for 2 h.
Bi2S3-TiO2-SiO2/TNA photocatalysts with different amounts of Bi2S3 were assembled uniformly onto the TiO2-SiO2/TNA using the SILAR technique in Bi2S3 solution with various concentrations (0.13, 0.20, 0.30, 0.40, 0.46 M). Briefly, each deposition cycle consisted of four distinct steps: (i) immersing the TiO2-SiO2/TNA in a solution containing the Bi3+ precursor for 5 min; (ii) rinsing the film in ethylene glycol for 1 min to eliminate the loosely bounded ions and drying it in ambient air; (iii) immersing the film in the S2− precursor for 5 min for the adsorption of sulfur ions that will react with pre-adsorbed bismuth ions; and (iv) rinsing the substrate again in ethylene glycol for 1 min to eliminate the loosely bound ions and then drying it at ambient temperature. To investigate the growth of Bi2S3 thin films, each procedure was repeated three times. Eventually, the films were dried at 80 °C for 3 h [28]. Scheme 1 illustrates the experimental setup and depiction outlining the SILAR technique for the preparation of Bi2S3-TiO2-SiO2/TNA.

2.4. Instruments

The X-ray diffraction (XRD) patterns were obtained using an XRD analytical diffractometer (X’PertPro, using a CuKα radiation source with λ = 1.54 Å; Ni-filtered, the Netherlands). The surface morphologies of the prepared samples were imaged using a field-emission scanning electron microscope (FESEM, INSPECT F50 FEI). To reveal the compositions of Bi2S3, TiO2, and SiO2 nanoparticles doped into the TNA surface, energy-dispersive spectroscopy (EDS) was used. The ultraviolet-visible diffuse reflectance spectra (UV–Vis/DRS) were recorded using a UV–vis spectrophotometer (Shimadzu UV-2450). A Fourier transform infrared (FTIR) spectrum was performed using a Nicolet Model Nexus 470 FTIR spectrometer in the range of 400–4000 cm−1. Electrochemical impedance spectroscopy (EIS) experiments (AUTOLAB PGSTAT302N/FRA2) to investigate the photoelectrochemical behaviors of photocatalysts were conducted in three-electrode compartments in a frequency range from 100 kHz to 10 mHz at OCP with an amplitude signal of 10 mV. The electrolyte was 0.2 M Na2SO4 solution, with a saturated calomel electrode, platinum foil, and the as-prepared samples serving as the reference, counter, and working electrodes, respectively.

2.5. Photocatalytic Performance Tests

The decomposition of the azo dye BY 28 by the prepared photocatalysts was evaluated in the presence of an Osram lamp (HQI-BT 400 W/D/DE40FLH1, 400 W). The wavelength range for the lamp was 380 to 750 nm. The distance between the solution and the light source was set to 20 cm. For the photodegradation step, the Bi2S3-TiO2-SiO2/TNA photocatalyst (with a surface area of 1 × 4 cm2) was vertically immersed in 25 mL of BY 28 (22.5 mg/L) aqueous solution and separated by a distance of 20 cm from the light source. Photocatalytic tests were performed with a magnetic stirrer for 30 min in the dark before the test to evaluate the adsorption/desorption equilibrium. The concentration of the degraded solution was determined by taking samples every 30 min and then examined using a UV–Vis spectrophotometer over 180 min. The maximum absorbance (λmax) of BY 28 was at the wavelength of 453 nm [29]. The change in the intensity of the absorbance peak of BY 28 during the photocatalytic process was recorded. The photocatalytic efficiency of the azo dye solution as a percentage was determined according to the following equation [16,17,30]:
B Y   28   p h o t o d e g r a d a t i o n   e f f i c i e n c y   % =   C 0 C C 0   × 100
where C0 is the concentration of BY 28 at 0 min after 30 min of adsorption–desorption equilibrium and C is the BY 28 concentration at time t. The first-order rate constant (kr) was determined from the slope of the fitted line using the Langmuir–Hinshelwood kinetic equation, as below [14,19]:
l n C C 0 =   K r t

2.6. Experimental Design and Statistical Analysis

The RSM model is a reliable statistical tool with important application in the optimization and analysis of the significance of various operational parameters, as it reduces the number of required experiments and highlights the interactions between process variables [31]. Minitab 16 software and a CCRD approach were used for the analysis of variance (ANOVA) of the regression parameters of the predicted response surface models based on CCRD-RSM to determine the interaction effects of three independent parameters, such as the amount of Bi2S3 loaded on the TNA, dye concentration, and pH, which have evident effects on the photocatalytic activity. These parameters and their levels are presented in Table 1.
The following quadratic polynomial model employed a correlation between the operating parameters and the photodegradation efficiency [32].
Y ( d e c o l o r i z a t i o n   e f f i c i e n c y % ) = β 0 + i = 1 n β i X i + i = 1 n β i i X i 2 + i = 1 n j = 1 n β i j X i X j
where Y is the response model, β0 is the constant of the model, Xi and Xj are the coded values of the independent variable, βi, βii, and βij correspond to the linear, quadratic, and cubic coefficients, respectively, and n is the number of independent parameters [33].

3. Results and Discussion

3.1. Characterization of Prepared Films

3.1.1. Structural Properties

The TiO2-SiO2/TNA and Bi2S3(9.6)-TiO2-SiO2/TNA films were characterized by X-ray diffraction patterns to determine the phase structure and composition of the samples. As demonstrated in Figure 1A, two of the photocatalyst films had significant 2θ peaks at 25.70°, 37.72°, 38.61°, 48.54°, 53.17°, and 62.80°, which correspond to (101), (103), (112), (105), (200), and (105), which matches well with the anatase phase of TiO2 (JCPDS 21–1272) [9,12]. It can be seen that the XRD pattern of TiO2-SiO2/TNA and Bi2S3(9.6)-TiO2-SiO2/TNA films had the characteristic (101) crystal plane of the anatase type of TiO2 [1]. After the deposition of Bi2S3 by the SILAR method, a new diffraction peak located at 35.54° appeared in Bi2S3(9.6)-TiO2-SiO2/TNA, which could correspond to the (240) lattice plane of Bi2S3 (JCPDS 84-0279) [34]. No signal was observed corresponding to SiO2 on Bi2S3(9.6)-TiO2-SiO2/TNA, consistent with previously reported literature [26,27,28]. This may be attributable to the amorphous nature of the synthesized SiO2 and the incorporation of Si within the crystalline lattice of TiO2 as an interstitial atom [35]. According to the Scherrer formula and Williamson–Hall (W-H) method, the crystallite size (D) of the sample could be analyzed from the X-ray line broadening of the anatase (1 0 1) reflection of the samples using Equation (4) [34,35]:
D = ( k   λ ) β c o s θ
where k, β, λ, and θ denote a constant (k = 0.9), the full width at the half maximum (FWHM) in radians, the wavelength of the X-ray (λ = 0.15406 nm), and the Bragg angle of reflection in radians, respectively. The Williamson–Hall method is a simplified integral width method that separates the contributions of diffraction line broadening, according to the following formula, Equation (5) [15,36];
β h k l c o s θ h k l =   k λ D + η   s i n θ h k l
βhkl is the instrument-corrected FWHM, k is a constant (k = 0.9), D is crystallite size, λ is the wavelength of the X-ray source, and η is the strain of the lattice [5,9,23]. By plotting 4sinθ values along the X-axis against βcosθ on the Y-axis and X-axis, respectively, the plot demonstrates the crystallite size of the sample according to the equation (Figure 1B). By applying the Scherer formula and the Williamson–Hall model to the diffraction patterns, the average particle sizes of 34.41 nm and 17.03 nm, respectively, were calculated for the Bi2S3(9.62)-TiO2-SiO2/TNA sample.

3.1.2. FT-IR Spectra

The chemical structures of Bi2S3(9.6)-TiO2-SiO2/TNA were characterized by the FT-IR spectrum (Figure 2). The FT-IR spectrum of the coating in Figure 2 displays the Si-O bonds at 500–600 cm−1 and 700–800 cm−1 [34]. In each of the two spectra, the formation of Si-O-Si and Si-O-Ti bands was confirmed, with peaks observed at 1080 and 925 cm−1. Saravanan et al. [37] studied the FTIR spectra of a silicon dioxide/titanium dioxide thin film prepared using the sol–gel synthesis process and found that the asymmetric Si–O–Si vibration peak appears at 1073 cm−1. It can be observed that the spectrum of the sorbent was indexed by distinguished peaks at around 3400 cm−1 and 1631 cm−1, which are ascribed to the bending vibration of O–H [38] and the O-H bending mode of the surface hydroxyl groups or free water adsorbed on the surface. The FT-IR spectrum of Bi2S3(9.6)-TiO2-SiO2/TNA (Figure 2B) demonstrated a decrease in some bands, which suggests that Bi2S3 is strongly linked with TiO2-SiO2/TNA. It also indicates that the stronger interactions between Bi2S3 and TiO2-SiO2/TNA could be related to the shifts in band positions and changes in the intensity of Bi2S3(9.6)-TiO2-SiO2/TNA.

3.1.3. SEM-EDX Studies

A cross-sectional FE-SEM of the surface morphology of the Bi2S3(9.6)-TiO2-SiO2/TNA sample is displayed in Figure 3A–C. As evident from the figure, well-defined TiO2 with a tubular structure appeared to be anodized at 60 V. As can be seen, the tubes presented a top-end-open, vertically oriented, and uniform structure (Figure 3D). The SEM images show that the nanotube arrays were covered by Bi2S3, TiO2, and SiO2 layers, so the surface morphology of the supported TNA became rough. EDX analysis data are listed in Figure 3D and confirm that the main elements composing the Bi2S3(9.6)-TiO2-SiO2/TNA sample were Ti, O, Si, Bi, and S.

3.1.4. Band Gap Energy and UV–Vis/DRS Analysis

To evaluate the optical properties of the TiO2-SiO2/TNA and Bi2S3(9.6)-TiO2-SiO2/TNA samples, UV–vis/DRS was applied, as exhibited in Figure 4A. The spectra of TiO2-SiO2/TNA indicated that the sample only absorbed UV light at wavelengths less than 400 nm due to the larger photon energy, which showed a steep absorption edge and had no absorbance in the visible region [37,38]. Similar phenomena have been reported by other researchers. Wang et al. [5] investigated the absorbance spectra of TiO2-ZrO2-SiO2 and TiO2-SiO2 films. All the films only had strong absorption at wavelengths less than 300 nm. The absorbance between 300 nm and 350 nm sharply decreased. Wang et al. [39] carried out UV–Vis/DRS measurements to investigate the optical properties of SiO2- TiO2/TNTs. SiO2- TiO2/TNT film predominantly demonstrates a photoresponse at wavelengths less than 380 nm, which is associated with the absorption threshold of anatase TiO2. Wanag et al. [40] investigated the UV-vis absorption spectra of a SiO2/TiO2 composite system for which strong absorbance only in the ultraviolet range of 250–375 nm was observed, and the process of photoexcitation was observed from the lower valence band to the upper CBis. Additionally, TiO2-SiO2/TNA modified with Bi2S3 exhibits a noticeable increase in the visible-light region, resulting from the combination of Bi2S3 and TiO2-SiO2/TNA, which indicates a significant red-shift effect [41]. The reflectance spectra of samples from the UV–Vis spectrophotometer and the plot of the modified Kubelka–Munk function were used. The Kubelka–Munk function is calculated by ( α h ν ) n = A ( h ν E g ) n / 2 , where α denotes the absorption coefficient, h is the Planck constant, ν corresponds to light frequency, A is a constant, Eg is the band gap, and n depends on the type of transition, where for direct allowed transition, n is equal to 2, and for an indirectly allowed transition, n is equal to ½ [42]. The Eg of semiconductor materials can be calculated by the extrapolation of the linear section of the ( α h ν )2 plot versus photon energy over the wavelength range of 200–700 nm. The results are shown in Figure 4B, and the energy band gap of TiO2-SiO2/TNA and Bi2S3(9.6)-TiO2-SiO2/TNA was determined, which can be reasonably assumed to be 3.27 eV and 1.87 eV, respectively [43].
These results indicate that the energy gap of Bi2S3(9.6)-TiO2-SiO2/TNA was narrower than that of TiO2-SiO2/TNA and, with improving efficiency of solar spectrum absorption, decreased the energy needed to excite the light-induced electrons. This catalyst exhibits noteworthy advantages because it utilizes more than two components [3] through the presence of Bi2S3(9.6)-TiO2-SiO2/TNA. The support of SiO2 reduces the particle size of TiO2 and improves its dispersive capacity, making more surface area of TiO2 available to absorb more light. At the same time, SiO2 has a strong reflective effect on light, and it is believed that the light reflected by SiO2 could be absorbed by the photocatalyst on the surface of SiO2 [4]. The relatively stronger light absorption promotes the creation of charge carriers in the photocatalytic reaction that can be efficiently enhanced to generate hydroxyl radicals (OH) and superoxide radicals (O2•−) on the surface [6]. It is noteworthy that the incorporation of Bi2S3 into TiO2-SiO2/TNA shows a favorable response to visible light because it has a small band gap and a superior absorption coefficient and promotes photocatalytic performance [7].

3.1.5. Electrochemical Analysis

To gain deep insight into the properties and photocatalytic mechanism of the Bi2S3-TiO2-SiO2/TNA heterojunction and investigate the charge transfer resistance on the interface during the photocatalytic process, EIS was recorded. Figure 5A,B illustrate the Nyquist and Bode log |Z| curves of EIS analysis. As is well known, the shorter radius of the arc of EIS spectra is indicative of a lower interfacial resistance and a more rapid charge transfer impedance. As the corresponding EIS results depicted in Figure 5A show, the Bi2S3(9.6)-TiO2-SiO2/TNA photocatalyst presented a shorter radius of the arc, reflecting the smallest charge transfer resistance on the interface and the higher charge transfer yield. The results agree with previous literature [8,9]:

3.2. Photocatalytic Performance of Photocatalysts

To evaluate the photocatalytic activity of the different samples by the photodegradation of azo dye, the reaction was carried out in a dye suspension with a concentration of 22.5 mg/L and a natural pH. Prior to the photocatalytic experiment, to determine the surface adsorption extent, experiments were carried out in dark conditions in the presence of the catalyst. The adsorption extent of BY 28 on the photocatalyst was determined to be about 15% after 180 min of contact time. Figure 6A shows a performance comparison for the development of the photocatalytic degradation of TiO2/TNA, TiO2-SiO2/TNA, Bi2S3-TiO2/TNA, and Bi2S3(9.6)-TiO2-SiO2/TNA. As can be observed from the degradation results, the Bi2S3(9.6)-TiO2-SiO2/TNA heterojunction exhibited higher photocatalytic activity over time than TiO2/TNA, TiO2-SiO2/TNA, and Bi2S3-TiO2/TNA within 180 min of irradiation. In contrast, the photocatalytic activity of TiO2/TNA was not noticeable. Moreover, these results show that TiO2-SiO2/TNA and Bi2S3-TiO2/TNA had almost lower degradability [2,4,43,44].
Using the linear correlation between ln(C/C0) and the irradiation time, the first-order kinetic rate constant could be determined, whereby k could be calculated for all samples displayed in Figure 6B. The rate constant k was 0.025 min−1 for the photocatalytic degradation of BY 28 by the obtained Bi2S3(9.6)-TiO2-SiO2/TNA, which is more than that of TiO2/TNA, TiO2-SiO2/TNA and Bi2S3-TiO2/TNA, especially as k was 0.0027, 0.0045, 0.0083, and min−1, which is 8 and 6 times that of TiO2/TNA and TiO2-SiO2/TNA and 3 times that of Bi2S3-TiO2/TNA, respectively. The results show that the deposition of Bi2S3 was beneficial for improving the photocatalytic activity of TiO2-SiO2/TNA photocatalysts. It can be seen that the reaction rate constant of TiO2-SiO2/TNA was 2 times higher than that of TiO2/TNA, indicating the effect of the SiO2, TiO2, and TNA ternary components for enhancing the photocatalytic ability of TiO2-SiO2/TNA. Additionally, the introduction of SiO2 can improve the specific surface area and light absorption, further promoting photocatalytic performance [8,23]. Table 2 shows a comparison of the results for dye removal efficiency with various photocatalysts. Notably, when using powder to photodegrade pollutants in wastewater, the performance of the photocatalytic systems increased, due to the strong oxidizing effect and large surface contact with the pollutants [19,36]. However, separation problems were encountered following repeated utilization. Comparing TiO2 nanotube array structures with photocatalyst nanopowders demonstrates that TNAs have higher surface-to-volume ratios and larger specific surface areas, which not only increases electron transfer as well as the interpenetration of holes, but also enhances the separation and transferring efficient of the photogenerated electron–hole (e–h+) pairs [3,10], offering great potential for enhancing the performance of photocatalytic systems [41].

3.3. Degradation Parameter Optimization Using RSM

In the present study, parameter optimization through RSM for the degradation of azo dye was determined. A five-level, three-factor design was implemented here. A total of 17 runs were necessary to improve the three individual parameters through the RSM. Consequently, the relationship between the degradation of BY 28 and the independent variable parameters was established and expressed by appropriate quadratic equations in the form of coded factors (Equation (6)):
Y B Y 28 = 88.62 + 6.90   X 1 + 4.55   X 2 + 17.24   X 3 4.95   X 1 X 1 20.51   X 2 X 2 8.46   X 3 X 3 + 2.12   X 1 X 2 6.78   X 1 X 3 + 4.87     X 2 X 3
The analysis of variance (ANOVA) showed that the mathematical model can be applied to samples with a probability level of 95% and quantifies the dominance of the control factor. The test results of Bi2S3-TiO2-SiO2/TNA for BY 28 photodegradation are listed in Table 3, which shows the status of considerable parameters and details of the ANOVA. To ensure the applicability of the model, the lack of fit was checked, which, with a p-value of 0.082 and an F-value of 0.76, indicated that the model was not significant regarding pure error, suggesting that the model effectively describes the measured response data within the experimental region. F-values and p-values (p < 0.01) are criteria for determining the significance of the less/more input variable effect on the measured response of the RSM model. F-values higher than the critical value (p < 0.05) indicate the significance of the quadratic model or role of the parameter [44]. p-values < 0.05 indicate a confidence level of 95%. In the present study, a p-value < 0.0001 demonstrated that the proposed model terms were significant, which shows that the second-order polynomial model fit the experimental results well [43]. From Table 3, it can be seen that the selected model was highly significant, with an F-value of 120.0 and a p-value of less than 0.0001, which is obvious from Fisher’s “F” test [41,42]. In addition, the regression coefficient (R2) for the photodegradation of BY 28 was 0.9936% and thus close to one, which indicates the suitability of the applied model and indicates that it reflected a good correlation between the observed and the predicted values. In addition, the values of R2 and the adjusted R2 were very close to each other (R2: 0.9936, Adj. R2: 0.9853), indicating a negligible amount of variation in the predicted and actual experimental data. The results presented in Table 4 indicate that a maximum degradation of 95.45% was attained, which closely aligns with the predicted response from the software (93.09%). The correspondence between the predicted response and experimental results demonstrates the applied model’s validity and its proficiency in forecasting the maximum degradation of BY 28. Figure 7A illustrates the relationship between the predicted responses versus experimental responses to the degradation of the dye, confirming the results were in close agreement. The normal plots of the studentized residuals are shown in Figure 7B. As shown in the plot, except for a few data points, most of the generated data points in the plot were scattered over a straight region, and no individual residuals exceeded the residual variance, confirming that the obtained data have a linear relationship and the errors along a zero mean have a normal distribution. This also confirmed the excellent suitability of the model for degradation purposes [51].

3.4. Key Parameter Optimization Using RSM

The three-dimensional (3D) surface plot generated from the RSM analysis displays a comparison among the contribution of key parameters in the design space. Figure 8A–C represent the effect of three independent parameters: the amount of Bi2S3 loaded on the TNA, dye concentration, and pH. In Figure 8A, interactions between the amount of Bi2S3 loaded on the TNA and the dye concentration on the response are shown. Five Bi2S3-TiO2-SiO2/TNA catalysts were prepared, and their Bi3+ content was determined by atomic absorption spectroscopy by digesting the catalysts in an HF + HNO3 + H2O2 mixture. The summarized results are shown in Table 5. It was observed that, initially, as the Bi2S3 content increased, with a Bi3+ concentration of 0.30 M, the degradation of the dye increased. These results show that increasing the Bi2S3 content enhanced the degradation efficacy of the dye by increasing the number of active sites that can be activated to generate reactive radicals, including superoxide radicals and hydroxyl, thereby facilitating an enhancement in the degradation rate of BY 28 [52]. However, when the Bi2S3 content increased to 15.82, the degradation efficacy of the dye decreased. Produced hydroxyl radicals can be recombined into less reactive H2O2 at higher doses [10,44,53]. These results reveal that an optimum amount of Bi2S3 was required to obtain high BY 28 degradation. Based on the results obtained from the RSM-CCRD technique in the photodegradation of BY 28 dye, the catalyst consisting of 9.6 mg/cm2 (designated as Bi2S3(9.6)-TiO2-SiO2/TNA) showed an increase in degradation efficiency compared to other synthesized catalysts. Hence, the Bi2S3(9.6)-TiO2-SiO2/TNA catalyst was used in the following experiments and designated as the “optimized catalyst”.
According to Figure 8B, the degradation process for BY 28 elimination was enhanced at a concentration of 22.5 mg/L. According to the collision theory, the probability of dye molecules colliding with the surface of the catalyst increases when the dye concentration rises to a certain level, leading to a high degradation efficiency [17,21]. Notably, at low dye concentrations, this probability is low, and as the concentration increases, it increases to the optimum concentration. The degradation efficiency was further enhanced by increasing the dye concentration to levels higher than the optimal range; high levels of BY 28 molecules can have a screening effect by absorbing some of the illuminated photons [54]. Thus, the degradation yield decreases due to the low production of free radicals and the saturation of the catalyst surface. Therefore, low and high BY 28 concentrations can result in a lower efficiency of photochemical degradation [52].
Figure 8C displays the 3D surface plot of the response to the significant quadratic effects of Bi2S3 and pH, indicating that the response increased the degradation efficiency with an increase in pH from 6 to 8 for the degradation of BY 28 dye. The surface of a catalyst exhibits a negative charge when the pH is basic, and a positive charge when the pH is acidic [41,42,45]. BY 28 dye is classified as a cationic azo dye, which has surfaces with a positive charge [52]. Because the surface charge of the Bi2S3(9.6)-TiO2-SiO2/TNA photocatalyst became negatively charged at a solution pH greater than 6, the dye’s degradation was observed to be higher in basic solutions compared to acidic solutions. An increase in the pH value in the solution, producing large numbers of H+ and thereby occupying the active sites on the surface of the photocatalyst and creating electrostatic interactions between the dye and H+ ions, led to a reduction in the interaction rate between the photocatalyst and the dye, which resulted in a lower degradation yield. At high pH values, the catalyst surface gradually became negatively charged [55]. The high pH value was favorable for the adsorption of cationic dyes due to electrostatic interaction, increasing the contact rate between the catalyst and the organic dye molecule [5], and assisted higher photogenerated charge carrier excitation, which resulted in a greater photocatalytic yield. The alkaline solution not only contained a higher concentration of OH ions but influenced the amount of hydroxyl radicals (OH) formed. The OH ions were adsorbed onto the catalyst’s surface and subsequently oxidized to OH radicals by the holes during the photoexcitation process [41]. The hydroxyl radicals, which were an active species in the photocatalytic yield of BY 28, are recognized as a highly reactive species that is a non-selective and extremely strong oxidant, resulting in the breakdown of various organic compounds [4]. At pH values higher than 8 (especially at high basic pH values), the repulsive force between the anionic BY 28 molecules, which were most abundant at these pH values, and the negatively charged catalyst surface repelled BY 28 molecules, resulting in a sharp decrease in degradation efficiency [8,9].
The identification of the operational parameters for the optimization of the process response has significant importance. RSM was used for process optimization using the different parameters listed in Table 4, with the objective of maximizing the degradation efficiencies of the dye.
The assessment of Chemical Oxygen Demand (COD) revealed a decline from 111 to 21 mg/L, thereby suggesting that the Bi2S3(9.6)-TiO2-SiO2/TNA photocatalyst has significant potential for the implementation of a photocatalytic system in the enhancement of water resource purification.

3.5. Energy Level Structure and the Photocatalytic Degradation Mechanism

Theoretical calculations based on Mulliken electronegativity theory [11] were used to determine the relative band edge positions of Bi2S3(9.6)-TiO2-SiO2/TNA, as shown:
E V B = χ E c + 1 2 E g
E C B = E V B E g
Specifically, the valance band potential (EVB), conduction band potential (ECB), band gap of the semiconductor (Eg), free electron energy of the standard hydrogen electrode potential (Ec = 4.5 eV), and absolute electronegativity of the corresponding semiconductor (X) were used to estimate the energies of the bands [52]. A possible photocatalytic mechanism during a photocatalytic reaction for Bi2S3(9.6)-TiO2-SiO2/TNA heterojunction was estimated. There were two charge transfer mechanisms: a conventional type II heterojunction and a direct Z-scheme heterojunction (Figure 9A,B). Irradiation by visible light can excite both TiO2 and Bi2S3 to generate charge carriers (e, h+). In a typical type II heterojunction system (Figure 9A), the charge carriers (e, h+) migrate through the typical charge transfer process [56,57].
The CB of TiO2 hosted the photoexcited electrons (eCB), and the CB potential of TiO2 (−0.29 eV) was more positive and not adequate to reduce O2 to O2•− (O2/O2•− = −0.33 eV vs. NHE) as an active radical species to degrade dye [56,57]. In addition, the VB of Bi2S3 hosted the photogenerated holes (hVB+), and the position of the VB was well above the standard potential of H2O/HO (OH/OH = 2.40 eV vs. NHE), and it is an inactive band which leads to no HO production [55,57,58]. In the second mechanism in Figure 9B, based on the potential edges in the Z-scheme mechanism, in the charge transfer pathway, the photogenerated charges migrated the photoexcited electrons at the CB of TiO2 transfer and combined with the photogenerated holes to the VB of Bi2S3, resulting in the photogenerated electrons in the CB of Bi2S3 and holes in the VB of TiO2. Therefore, the abundant accumulated electrons in Bi2S3 with strong reducing power could rapidly release CB to the adsorbed O2 to generate O2•−, while the holes were formed in the TiO2 valance band, which can not only react with pollutants but also generate OH by reacting with H2O, which oxidizes BY 28 to H2O and CO2. As a result, the formation of Z-schemes in the photocatalytic system could explain the improvement of photocatalytic performance for organic pollutants [55,56,57].

3.6. Reusability of Bi2S3(9.6)-TiO2-SiO2/TNA

The reuse experiment was conducted to evaluate the practicability and cost-effectiveness of using photocatalysts for environmental remediation [43,51]. The photocatalyst was reused for multiple runs at the optimal conditions [59]. The results are shown in Figure 10. The reused photocatalyst was washed and dried at a temperature of 100 °C and used in further experiments in fresh conditions. The results indicate that the efficiency of photocatalyst degradation decreased to about 9% of the initial photocatalytic activity after five reuse runs.

4. Conclusions

In summary, TiO2, SiO2, and Bi2S3 nanoparticles were loaded on TNAs (Bi2S3-TiO2-SiO2/TNA) and used for the degradation of BY 28 under visible light. The XRD results indicate that TiO2-SiO2/TNA and Bi2S3(9.6)-TiO2-SiO2/TNA films have the characteristics of the (101) crystal plane of the anatase type of TiO2. UV–Vis/DRS measurements of TiO2-SiO2/TNA modified with Bi2S3 exhibited a noticeable increase in the visible-light region, which provided effective visible-light absorption and electron excitation. The Nyquist curves of the EIS analysis showed the Bi2S3(9.6)-TiO2-SiO2/TNA photocatalyst had a shorter radius of arc, reflecting the smallest charge transfer resistance on the interface and the higher charge transfer yield. In this research paper, CCRD-RSM was employed for the optimization of the key parameters in BY 28 photocatalytic degradation. The RSM model was found to have an R2 and adjusted R2 value of 0.9936 and 0.9853, respectively, and could significantly (p < 0.0001) predict the response variables. Using Bi2S3(9.6)-TiO2-SiO2/TNA, the optimum conditions were set as 0.30 M, 22.5 mg/L, and 8.34 for the amount of Bi2S3 loaded on the TNA, dye concentration, and pH, respectively, with a maximum degradation efficiency (%) of 95.45%. The amount of Bi2S3 loaded on the TNA is a critical factor in the photocatalytic performance of Bi2S3-TiO2-SiO2/TNA due to the enhancement of light absorption and the ability of the Bi2S3 to absorb more visible light and promote the photocatalytic performance to generate a higher number of electrons and holes. On the other hand, SiO2 has an evidently enhanced effect on the photocatalytic performance of Bi2S3(9.6)-TiO2-SiO2/TNA because it considerably enhances the thermal stability and dispersivity and increases the surface area. The Bi2S3(9.6)-TiO2-SiO2/TNA photocatalyst has two mechanisms: the conventional type II heterojunction and direct Z-scheme systems. The direct Z-scheme mechanism is suitable for the production of superoxide due to the accumulating electrons in the CB of Bi2S3. In addition, the direct Z-scheme can facilitate charge transportation and separation during photocatalytic reactions, which suppresses the electron–hole recombination rate, improving the photodegradation ability of BY 28.

Author Contributions

Conceptualization, A.A. (Amal Abdulrahman); data curation, N.G.; formal analysis, A.A. (Amal Abdulrahman) and S.S.S.; funding acquisition, M.S.; investigation, A.A. (Abdelfattah Amari) and M.S.; resources, N.G.; software, N.G. and A.A. (Abdelfattah Amari); supervision, M.S.; validation, A.A. (Amal Abdulrahman), Z.A., S.S.S., and M.S.; visualization, Z.A.; writing—original draft, A.A. (Amal Abdulrahman) and A.A. (Abdelfattah Amari); writing—review and editing, Z.A. and S.S.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Deanship of Research and Graduate Studies at King Khalid University under grant number RGP.2/243/46 and by Northern Border University under the project number NBU-FFR-2025-2105-XX.

Data Availability Statement

Data are available from the corresponding author upon reasonable request.

Acknowledgments

The authors extend their appreciation to the Deanship of Research and Graduate Studies at King Khalid University for funding this work as a Large Research Project under grant number RGP.2/243/46. The authors extend their appreciation to the Deanship of Scientific Research at Northern Border University, Arar, SA, for funding this research work through the project number “NBU-FFMRA-2025-2105-07”.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Wang, J.; Wan, Y.; He, Q.; Zhou, X.; Song, Q.; Chen, Z.; Wang, P.; Wang, J.; Wang, T.; Chen, P.; et al. Enhanced photoelectrocatalytic performance of ZnO/Co3O4 nanoflowers modified TiO2 nanotube array under visible light. J. Environ. Chem. Eng. 2023, 11, 111604. [Google Scholar] [CrossRef]
  2. Liu, X.; Liu, W.; Xin, S.; Gao, S.; Huo, S.; Fu, W.; Gao, M.; Xie, H. Photoelectrocatalytic degradation of tetracycline through C3N5/TiO2 nanotube arrays photoanode under visible light: Performance, DFT calculation and mechanism. J. Environ. Chem. Eng. 2023, 11, 110576. [Google Scholar] [CrossRef]
  3. Belal, R.M.; Zayed, M.A.; El-Sherif, R.M.; Ghany, N.A.A. Advanced electrochemical degradation of basic yellow 28 textile dye using IrO2/Ti meshed electrode in different supporting electrolytes. J. Electroanal. Chem. 2021, 882, 114979. [Google Scholar] [CrossRef]
  4. Shi, J.; Liao, R.; Jia, R.; Liu, Y.; Wu, D.; Chang, S.; Zhang, N.; Gao, G.; Wang, X.; Hu, D.; et al. A novel combustion drying synthesis route of 3D WO3–SiO2 composite aerogels for enhanced adsorption and visible light photocatalytic activity. J. Non-Cryst. Solids 2023, 609, 122259. [Google Scholar] [CrossRef]
  5. Wu, Y.; Guan, M.; Chang, X.; Wang, J.; Xu, S. Homogeneous double-layer TiO2-ZrO2-SiO2 photocatalyst with multi-heterojunction structure for enhanced visible light-responsive photocatalytic activity. J. Mol. Liq. 2023, 369, 120959. [Google Scholar] [CrossRef]
  6. Wang, Q.; Zhao, S.; Zhao, Y.; Deng, Y.; Yang, W.; Ye, Y.; Wang, K. Construction of Z-scheme Bi2O3/CeO2 heterojunction for enhanced photocatalytic capacity of TiO2 NTs. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2024, 304, 123405. [Google Scholar] [CrossRef]
  7. Fakhravar, S.; Farhadian, M.; Tangestaninejad, S. Excellent performance of a novel dual Z-scheme Cu2S/Ag2S/BiVO4 heterostructure in metronidazole degradation in batch and continuous systems: Immobilization of catalytic particles on α-Al2O3 fiber. Appl. Surf. Sci. 2020, 505, 144599. [Google Scholar] [CrossRef]
  8. Geioushy, R.; El-Sheikh, S.; Azzam, A.B.; Salah, B.A.; El-Dars, F.M. One-pot fabrication of BiPO4/Bi2S3 hybrid structures for visible-light driven reduction of hazardous Cr (VI). J. Hazard. Mater. 2020, 381, 120955. [Google Scholar] [CrossRef]
  9. Wang, Q.; Ren, C.; Zhao, Y.; Fang, F.; Yin, Y.; Ye, Y.; Yang, K.; Yang, Q.; Wang, K. Photocatalytic pollutant elimination and hydrogen production over TiO2 NTs/Bi2S3-MoS2 with Z-scheme configuration: Kinetics and mechanism. Mater. Res. Bull. 2023, 167, 112430. [Google Scholar] [CrossRef]
  10. Hou, J.; Chen, Y.; Yimiti, M.; Xue, Y.; Zhang, H.; Wang, Q. One-pot solvothermal synthesis of Bi2S3-CdS sensitized TiO2 NT films for improved photocatalytic performance. Mol. Catal. 2024, 569, 114560. [Google Scholar] [CrossRef]
  11. Hmamouchi, S.; El Yacoubi, A.; El Hezzat, M.; Sallek, B.; El Idrissi, B.C. Optimization of photocatalytic parameters for MB degradation by g-C3N4 nanoparticles using Response Surface Methodology (RSM). Diam. Relat. Mater. 2023, 136, 109986. [Google Scholar] [CrossRef]
  12. Eskandari, P.; Amarloo, E.; Zangeneh, H.; Rezakazemi, M.; Aminabhavi, T.M. Photocatalytic degradation of metronidazole and oxytetracycline by novel l-Arginine (C, N codoped)-TiO2/g-C3N4: RSM optimization, photodegradation mechanism, biodegradability evaluation. Chemosphere 2023, 337, 139282. [Google Scholar] [CrossRef] [PubMed]
  13. Raj, S.; Chainy, Y.; Sardar, P.; Samanta, A.N. Construction of a novel Bi2S3/MIL-101 (Fe) Z-scheme heterojunction for the effective visible-light assisted photocatalytic degradation of Tetracycline Hydrochloride: Performance, mechanism, and degradation pathway. Surf. Interfaces 2025, 60, 106011. [Google Scholar] [CrossRef]
  14. Ali, H.; Yasir, M.; Ngwabebhoh, F.A.; Sopik, T.; Zandraa, O.; Sevcik, J.; Masar, M.; Machovsky, M.; Kuritka, I. Boosting photocatalytic degradation of estrone hormone by silica-supported g-C3N4/WO3 using response surface methodology coupled with Box-Behnken design. J. Photochem. Photobiol. A Chem. 2023, 441, 114733. [Google Scholar] [CrossRef]
  15. Karimi-Shamsabadi, M.; Behpour, M. Comparing photocatalytic activity consisting of Sb2S3 and Ag2S on the TiO2–SiO2/TiO2 nanotube arrays-support for improved visible-light-induced photocatalytic degradation of a binary mixture of basic blue 41 and basic red 46 dyes. Int. J. Hydrogen. Energy 2021, 46, 26989–27013. [Google Scholar] [CrossRef]
  16. Ansari, F.; Sheibani, S.; Fernández-García, M. A response surface methodology optimization for efficient photocatalytic degradation over reusable CuxO/TiO2 nanocomposite on copper wire. Mater. Res. Bull. 2023, 166, 112342. [Google Scholar] [CrossRef]
  17. Koutavarapu, R.; Babu, B.; Reddy, C.V.; Reddy, I.N.; Reddy, K.R.; Rao, M.; Aminabhavi, T.M.; Cho, M.; Kim, D.; Shim, J. ZnO nanosheets-decorated Bi2WO6 nanolayers as efficient photocatalysts for the removal of toxic environmental pollutants and photoelectrochemical solar water oxidation. J. Environ. Manag. 2020, 265, 110504. [Google Scholar] [CrossRef]
  18. Bai, H. Synthesis of Bi2S3-TiO2 nanocomposite and its electrochemical and enhanced photocatalytic properties for phenol degradation. Int. J. Electrochem. Sci. 2023, 18, 100071. [Google Scholar]
  19. Asadpoor, M.; Arjmand, M.; Farhadian, M.; Omidkhah, M.; Zinatizadeh, A. Optimization and modeling of the photocatalytic activities of a novel visible driven CNT/TiO2/BiOBr/Bi2S3 nanocomposite. Desalination Water Treat. 2021, 209, 219–229. [Google Scholar] [CrossRef]
  20. Wu, Z.; Yuan, D.; Lin, S.; Guo, W.; Zhan, D.; Sun, L.; Lin, C. Enhanced photoelectrocatalytic activity of Bi2S3–TiO2 nanotube arrays hetero-structure under visible light irradiation. Int. J. Hydrogen. Energy 2020, 45, 32012–32021. [Google Scholar] [CrossRef]
  21. Liu, Z.; Xu, K.; Yu, H.; Zhang, M.; Sun, Z. In-situ preparation of double Z-scheme Bi2S3/BiVO4/TiO2 ternary photocatalysts for enhanced photoelectrochemical and photocatalytic performance. Appl. Surf. Sci. 2021, 545, 148986. [Google Scholar] [CrossRef]
  22. Drmosh, Q.; Hezam, A.; Hendi, A.; Qamar, M.; Yamani, Z.; Byrappa, K. Ternary Bi2S3/MoS2/TiO2 with double Z-scheme configuration as high performance photocatalyst. Appl. Surf. Sci. 2020, 499, 143938. [Google Scholar] [CrossRef]
  23. Ghattavi, S.; Nezamzadeh-Ejhieh, A. A visible light driven AgBr/g-C3N4 photocatalyst composite in methyl orange photodegradation: Focus on photoluminescence, mole ratio, synthesis method of g-C3N4 and scavengers. Compos. Part B Eng. 2020, 183, 107712. [Google Scholar] [CrossRef]
  24. Nasseh, N.; Pour, A.K.; Azqandi, M.; Kassim, W.M.; Barikbin, B.; Al-Musawi, T.J. Integrated Fenton-like and photocatalytic system for metronidazole degradation in water using FeNi3/SiO2/CuS magnetic nanocomposite. Desalination Water Treat. 2023, 303, 160–170. [Google Scholar] [CrossRef]
  25. Luo, X.-L.; Yang, S.-Y.; Wang, Z.-L.; Xu, Y.-H. Synthesis of Z–scheme Bi2S3/RGO/BiVO4 photocatalysts with superior visible light photocatalytic effectiveness for pollutant degradation. Sep. Purif. Technol. 2023, 318, 123966. [Google Scholar] [CrossRef]
  26. Muniyandi, G.R.; Ubagaram, J.; Srinivasan, A.; James, D.R.; Pugazhenthiran, N.; Govindasamy, C.; Joseph, J.A.; Bosco, A.J.; Mahalingam, S.; Kim, J. Advanced Z-scheme Hg-C3N4/Bi2S3 nanocomposites: Boosting photocatalytic degradation of antibiotics under visible light exposure. J. Ind. Eng. Chem. 2024, 140, 647–657. [Google Scholar] [CrossRef]
  27. Chawla, A.; Sudhaik, A.; Sonu; Kumar, R.; Raizada, P.; Ahamad, T.; Khan, A.A.P.; Van Le, Q.; Nguyen, V.-H.; Thakur, S.; et al. Bi2S3-based photocatalysts: Properties, synthesis, modification strategies, and mechanistic insights towards environmental sustainability and green energy technologies. Coord. Chem. Rev. 2025, 529, 216443. [Google Scholar] [CrossRef]
  28. Lan, M.; Dong, X.; Zheng, N.; Zhang, Y. One-step construction of sulfur-vacancy-rich SnS/Bi2S3 Z-scheme heterostructure photocatalyst for significant removal of hazardous Cr (VI). Chem. Eng. Sci. 2025, 311, 121626. [Google Scholar] [CrossRef]
  29. Verma, M.K.; Sharma, M.; Dudi, M.; Singla, L.; Kumar, R. Curcumin stabilized Bi2S3@Ce2S3 nanoflowers: An efficient photocatalyst against harmful bacteria and toxic cationic dyes in aqueous medium. Next Mater. 2025, 8, 100694. [Google Scholar] [CrossRef]
  30. Mirzaeinia, S.; Pazhang, M.; Imani, M.; Chaparzadeh, N.; Amani-Ghadim, A.R. Improving the stability of uricase from Aspergillus flavus by osmolytes: Use of response surface methodology for optimization of the enzyme stability. Process Biochem. 2020, 94, 86–98. [Google Scholar] [CrossRef]
  31. Niknam, H.; Sadeghzadeh-Attar, A. Mg-doped TiO2 nanorods-SrTiO3 heterojunction composites for efficient visible-light photocatalytic degradation of basic yellow 28. Opt. Mater. 2023, 136, 113395. [Google Scholar] [CrossRef]
  32. Chahkandi, M.; Zargazi, M.; Ghiasabadi, K.B.; Chung, J.S. Core-shell structured Ag-HAp@Bi2S3 as effective S-scheme heterojunction photocatalyst: Induced with internal polarization electric field effect and co-catalyst fashions. J. Mol. Liq. 2024, 399, 124423. [Google Scholar] [CrossRef]
  33. Rao, V.N.; Reddy, N.L.; Kumari, M.M.; Ravi, P.; Sathish, M.; Kuruvilla, K.; Preethi, V.; Reddy, K.R.; Shetti, N.P.; Aminabhavi, T.M.; et al. Photocatalytic recovery of H2 from H2S containing wastewater: Surface and interface control of photo-excitons in Cu2S@TiO2 core-shell nanostructures. Appl. Catal. B Environ. 2019, 254, 174–185. [Google Scholar]
  34. Reddy, C.V.; Reddy, I.N.; Harish, V.; Reddy, K.R.; Shetti, N.P.; Shim, J.; Aminabhavi, T.M. Efficient removal of toxic organic dyes and photoelectrochemical properties of iron-doped zirconia nanoparticles. Chemosphere 2020, 239, 124766. [Google Scholar] [CrossRef]
  35. Reddy, N.R.; Bharagav, U.; Kumari, M.M.; Cheralathan, K.; Shankar, M.; Reddy, K.R.; Saleh, T.A.; Aminabhavi, T.M. Highly efficient solar light-driven photocatalytic hydrogen production over Cu/FCNTs-titania quantum dots-based heterostructures. J. Environ. Manag. 2020, 254, 109747. [Google Scholar] [CrossRef]
  36. Jiang, B.; Brandt, S.A. A fractal perspective on scale in geography. ISPRS Int. J. Geo-Inf. 2016, 5, 95. [Google Scholar] [CrossRef]
  37. Saravanan, S.; Dubey, R.S.; Subbarao, P.S. Optical investigation of sol–gel-synthesized and spin-coated SiO2/TiO2 multilayers. Nanomater. Energy 2023, 12, 44–48. [Google Scholar] [CrossRef]
  38. Shamsabadi, M.K.; Behpour, M. Fabricated CuO–ZnO/nanozeolite X heterostructure with enhanced photocatalytic performance: Mechanism investigation and degradation pathway. Mater. Sci. Eng. B 2021, 269, 115170. [Google Scholar] [CrossRef]
  39. Wang, J.; Ran, Q.; Xu, X.; Zhu, B.; Zhang, W. Preparation and optical properties of TiO2-SiO2 thin films by sol-gel dipping method. IOP Conf. Ser. Earth Environ. Sci. 2019, 310, 042029. [Google Scholar] [CrossRef]
  40. Babyszko, A.; Wanag, A.; Kusiak-Nejman, E.; Morawski, A.W. Effect of calcination temperature of SiO2/TiO2 photocatalysts on UV-VIS and VIS removal efficiency of color contaminants. Catalysts 2023, 13, 186. [Google Scholar] [CrossRef]
  41. Stelo, F.; Kublik, N.; Ullah, S.; Wender, H. Recent advances in Bi2MoO6 based Z-scheme heterojunctions for photocatalytic degradation of pollutants. J. Alloys Compd. 2020, 829, 154591. [Google Scholar] [CrossRef]
  42. Sundaram, P.S.; Sangeetha, T.; Rajakarthihan, S.; Vijayalaksmi, R.; Elangovan, A.; Arivazhagan, G. XRD structural studies on cobalt doped zinc oxide nanoparticles synthesized by coprecipitation method: Williamson-Hall and size-strain plot approaches. Phys. B Condens. Matter 2020, 595, 412342. [Google Scholar] [CrossRef]
  43. Uma, K.; Chen, S.-W.; KrishnaKumar, B.; Jeyaprabha, C.; Yang, T.C.-K.; Lin, J.-H. Enhanced photocatalytic activity of CdS nanostar decorated SiO2/TiO2 composite spheres and the simulation effect using FDTD model. Ionics 2021, 27, 397–406. [Google Scholar] [CrossRef]
  44. Huerta-Flores, A.M.; García-Gómez, N.A.; de la Parra, S.M.; Sánchez, E.M. Comparative study of Sb2S3, Bi2S3 and In2S3 thin film deposition on TiO2 by successive ionic layer adsorption and reaction (SILAR) method. Mater. Sci. Semicond. Process. 2015, 37, 235–240. [Google Scholar] [CrossRef]
  45. Belghiti, M.; El Mersly, L.; Tanji, K.; Belkodia, K.; Lamsayety, I.; Ouzaouit, K.; Faqir, H.; Benzakour, I.; Rafqah, S.; Outzourhit, A. Sol-gel combined mechano-thermal synthesis of Y2O3, CeO2, and PdO partially coated ZnO for sulfamethazine and basic yellow 28 photodegradation under UV and visible light. Opt. Mater. 2023, 136, 113458. [Google Scholar] [CrossRef]
  46. Niknam, H.; Sadeghzadeh-Attar, A. Constructing trinary heterostructure of TiO2/CoCr2O4/SrTiO3 to enhance photocatalytic activity toward degradation of yellow 28 dye. Mater. Chem. Phys. 2023, 299, 127489. [Google Scholar] [CrossRef]
  47. Behpour, M.; Mehrzad, M.; Hosseinpour, M.S. TiO2 thin film: Preparation, characterization, and its photocatalytic degradation of basic yellow 28 dye. J. Nanostruct. 2015, 5, 183–187. [Google Scholar]
  48. Foulady-Dehaghi, R.; Behpour, M. Visible and solar photodegradation of textile wastewater by multiple doped TiO2/Zn nanostructured thin films in fixed bed photoreactor mode. Inorg. Chem. Commun. 2020, 117, 107946. [Google Scholar] [CrossRef]
  49. Behpour, M.; Foulady-Dehaghi, R.; Mir, N. Considering photocatalytic activity of N/F/S-doped TiO2 thin films in degradation of textile waste under visible and sunlight irradiation. Sol. Energy 2017, 158, 636–643. [Google Scholar] [CrossRef]
  50. Dehaghi, R.F.; Behpour, M.; Mir, N. Purification of textile wastewater by using coated Sr/S/N doped TiO2 nanolayers on glass orbs. Korean J. Chem. Eng. 2018, 35, 1441–1449. [Google Scholar] [CrossRef]
  51. Wang, D.; Hou, P.; Yang, P.; Cheng, X. BiOBr@ SiO2 flower-like nanospheres chemically-bonded on cement-based materials for photocatalysis. Appl. Surf. Sci. 2018, 430, 539–548. [Google Scholar] [CrossRef]
  52. Wang, Z.; Su, B.; Xu, J.; Hou, Y.; Ding, Z. Direct Z-scheme ZnIn2S4/LaNiO3 nanohybrid with enhanced photocatalytic performance for H2 evolution. Int. J. Hydrogen. Energy 2020, 45, 4113–4121. [Google Scholar] [CrossRef]
  53. Helal, A.; Harraz, F.A.; Ismail, A.A.; Sami, T.M.; Ibrahim, I. Hydrothermal synthesis of novel heterostructured Fe2O3/Bi2S3 nanorods with enhanced photocatalytic activity under visible light. Appl. Catal. B Environ. 2017, 213, 18–27. [Google Scholar] [CrossRef]
  54. Reddy, C.V.; Koutavarapu, R.; Reddy, K.R.; Shetti, N.P.; Aminabhavi, T.M.; Shim, J. Z-scheme binary 1D ZnWO4 nanorods decorated 2D NiFe2O4 nanoplates as photocatalysts for high efficiency photocatalytic degradation of toxic organic pollutants from wastewater. J. Environ. Manag. 2020, 268, 110677. [Google Scholar] [CrossRef]
  55. Cao, J.; Chen, X.; Chen, C.; Chen, A.; Zheng, W. High efficiency Bi2S3/Bi2MoO6/TiO2 photoanode for photoelectrochemical hydrogen generation. J. Alloys Compd. 2023, 965, 171139. [Google Scholar] [CrossRef]
  56. Zabihi, M.; Motavalizadehkakhky, A. PbS/ZIF-67 nanocomposite: Novel material for photocatalytic degradation of basic yellow 28 and direct blue 199 dyes. J. Taiwan Inst. Chem. Eng. 2022, 140, 104572. [Google Scholar] [CrossRef]
  57. Zhou, X.; Mu, Y.; Zhang, S.; Gao, L.; Chen, H.; Mu, J.; Zhang, X.; Zhang, M.; Liu, W. CdS nanoparticles sensitized high energy facets exposed SnO2 elongated octahedral nanoparticles film for photocatalytic application. Mater. Res. Bull. 2019, 111, 118–125. [Google Scholar] [CrossRef]
  58. Li, X.; Wang, K.; Mao, Z.; Kumar, J.M.; Han, C.; Li, L.; Yu, R.; Gu, Y.; Zhang, Y. Facile synthesis of hierarchical structure Bi2S3/TiO2 heterojunction and enhancing light-harvesting performance. Vacuum 2023, 217, 112579. [Google Scholar] [CrossRef]
  59. Zare, M.H.; Mehrabani-Zeinabad, A. Photocatalytic activity of ZrO2/TiO2/Fe3O4 ternary nanocomposite for the degradation of naproxen: Characterization and optimization using response surface methodology. Sci. Rep. 2022, 12, 10388. [Google Scholar] [CrossRef]
Scheme 1. The experimental setup and strategy adopted for the preparation of Bi2S3-TiO2-SiO2/TNA.
Scheme 1. The experimental setup and strategy adopted for the preparation of Bi2S3-TiO2-SiO2/TNA.
Water 17 01875 sch001
Figure 1. (A) XRD patterns of TiO2-SiO2/TNA and Bi2S3(9.6)-TiO2-SiO2/TNA. (B) Typical W-H diagram of Bi2S3(9.6)-TiO2-SiO2/TNA for determining the size of photocatalysts.
Figure 1. (A) XRD patterns of TiO2-SiO2/TNA and Bi2S3(9.6)-TiO2-SiO2/TNA. (B) Typical W-H diagram of Bi2S3(9.6)-TiO2-SiO2/TNA for determining the size of photocatalysts.
Water 17 01875 g001aWater 17 01875 g001b
Figure 2. FT-IR of (A) TiO2-SiO2/TNA and (B) Bi2S3(9.6)-TiO2-SiO2/TNA.
Figure 2. FT-IR of (A) TiO2-SiO2/TNA and (B) Bi2S3(9.6)-TiO2-SiO2/TNA.
Water 17 01875 g002
Figure 3. (AC) SEM images and (D) EDX spectrum of the Bi2S3(9.6)-TiO2-SiO2/TNA.
Figure 3. (AC) SEM images and (D) EDX spectrum of the Bi2S3(9.6)-TiO2-SiO2/TNA.
Water 17 01875 g003aWater 17 01875 g003b
Figure 4. (A) UV-Vis/DRS and (B,C) the plots of typical Tauc TiO2-SiO2/TNA and Bi2S3(9.6)-TiO2-SiO2/TNA.
Figure 4. (A) UV-Vis/DRS and (B,C) the plots of typical Tauc TiO2-SiO2/TNA and Bi2S3(9.6)-TiO2-SiO2/TNA.
Water 17 01875 g004aWater 17 01875 g004b
Figure 5. (A) Nyquist plots and (B) EIS Bode plots in the log (|Z|) for TiO2-SiO2/TNA and Bi2S3(9.6)-TiO2-SiO2/TNA samples.
Figure 5. (A) Nyquist plots and (B) EIS Bode plots in the log (|Z|) for TiO2-SiO2/TNA and Bi2S3(9.6)-TiO2-SiO2/TNA samples.
Water 17 01875 g005
Figure 6. (A) Typical plot of C/Co and (B) ln C/Co versus time for calculating the rate constant of BY 28 photodegradation (C BY 28: 22.5 mg/L at pH 8.24, time: 180 min).
Figure 6. (A) Typical plot of C/Co and (B) ln C/Co versus time for calculating the rate constant of BY 28 photodegradation (C BY 28: 22.5 mg/L at pH 8.24, time: 180 min).
Water 17 01875 g006
Figure 7. Diagnostic plots for the proposed model in BY 28 photodegradation. (A) Predicted value versus experimental data of BY 28 photodegradation. (B) Normal probability plot of residuals.
Figure 7. Diagnostic plots for the proposed model in BY 28 photodegradation. (A) Predicted value versus experimental data of BY 28 photodegradation. (B) Normal probability plot of residuals.
Water 17 01875 g007
Figure 8. (A–C) Three-dimensional surface plot of the simultaneous interaction of parameters for the photodegradation efficiency (%) of BY 28 dye.
Figure 8. (A–C) Three-dimensional surface plot of the simultaneous interaction of parameters for the photodegradation efficiency (%) of BY 28 dye.
Water 17 01875 g008aWater 17 01875 g008b
Figure 9. The mechanism of e/h+ production and charge separation in a Bi2S3(9.6)-TiO2-SiO2/TNA system: (A) type II mechanism and (B) Z-type mechanism.
Figure 9. The mechanism of e/h+ production and charge separation in a Bi2S3(9.6)-TiO2-SiO2/TNA system: (A) type II mechanism and (B) Z-type mechanism.
Water 17 01875 g009
Figure 10. Results obtained in successive reuse runs of Bi2S3(9.6)-TiO2-SiO2/TNA in BY 28 dye photodegradation (C BY 28: 22.5 mg/L at pH 8.24, time: 180 min).
Figure 10. Results obtained in successive reuse runs of Bi2S3(9.6)-TiO2-SiO2/TNA in BY 28 dye photodegradation (C BY 28: 22.5 mg/L at pH 8.24, time: 180 min).
Water 17 01875 g010
Table 1. Experimental variables and their levels in CCRD.
Table 1. Experimental variables and their levels in CCRD.
FactorsLevels
LowCentral (0)High
−10+1
X1: [Bi3+] (M)0.300.400.50
X2: [BY 28] (mg/L)102030
X3: pH468
Table 2. Comparison of BY 28 dye degradation efficiency among catalysts.
Table 2. Comparison of BY 28 dye degradation efficiency among catalysts.
PhotocatalystsSources of IrradiationTime (min)Degradation
(%)
Ref.
1.5 Mg–TiO2/SrTiO3Visible-light irradiation15097%[31]
Y2O3, CeO2, and ZnO/PdOVisible-light irradiation12094%[45]
TiO2/CoCr2O4/SrTiO3Visible-light irradiation12098%[46]
N-S doped TiO2Visible-light irradiation18098%[47]
Zn/Tu-TiO2Visible-light irradiation24095%[48]
N/F/S-doped TiO2Visible-light irradiation36094%[49]
Sr/S/N doped TiO2 nanolayers on glass orbsVisible-light irradiation48091%[50]
Bi2S3-TiO2-SiO2/TNAVisible-light irradiation18095.45%This work
Table 3. ANOVA table from CCRD for BY 28 photodegradation.
Table 3. ANOVA table from CCRD for BY 28 photodegradation.
ItemDegrees of FreedomSum of SquaresF-Valuep-Value
Model910,396.8120.050.000
Linear34996.3173.080.000
X11651.467.690.000
X21283.529.470.001
X314061.4422.070.000
Square34907.7170.010.000
X1*X11176.618.360.004
X2*X214744.7493.080.000
X3*X31808.083.970.000
2-Way Interaction3492.717.070.001
X1*X2136.13.750.094
X1*X31267.227.760.001
X2*X31189.419.690.003
Error767.4
Lack of Fit531.80.360.847
Pure Error235.6
Total1610,464.1
Table 5. Amount of Bi2S3 loaded on the TNA.
Table 5. Amount of Bi2S3 loaded on the TNA.
CatalystsC (M)Mass Loading (mg/cm2)
Bi3+Bi2S3
Bi2S3(6.1)-TiO2-SiO2/TNA0.136.1
Bi2S3(9.6)-TiO2-SiO2/TNA0.209.6
Bi2S3(11.4)-TiO2-SiO2/TNA0.3011.4
Bi2S3(13.7)-TiO2-SiO2/TNA0.4013.7
Bi2S3(15.8)-TiO2-SiO2/TNA0.4615.8
The Bi2S3-TiO2-SiO2/TNA photocatalytic process under optimal conditions in degradation of BY 28.
DyesX1X2X3Deg. (%)
CodedActual
(M)
CodedActual
(mg/L)
CodedActual
BY 280.180.300.2522.51.038.3495.45
Table 4. Experimental design for optimization of significant variables.
Table 4. Experimental design for optimization of significant variables.
RunIndependent Variables% Degradation
X1: [Bi3+] (M)X2: Dye Concentration (mg/L)X3: pHExperimental ValuePredicted Value
111184.2183.58
211−153.6652.73
3−1.680067.3366.41
41−1−147.1348.16
500092.5788.68
6−11179.3578.56
7−11−122.1423.56
81.680090.2588.06
900088.6988.68
101−1161.7662.09
11001.6895.4593.09
12−1−1−127.6527.43
13−1−1161.8565.52
1401.68038.2837.42
150−1.68025.6422.61
1600084.1488.68
1700−1.6836.6335.11
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

Abdulrahman, A.; Algarni, Z.; Ghazouani, N.; Sammen, S.S.; Amari, A.; Scholz, M. Enhancing Photocatalytic Activity with the Substantial Optical Absorption of Bi2S3-SiO2-TiO2/TiO2 Nanotube Arrays for Azo Dye Wastewater Treatment. Water 2025, 17, 1875. https://doi.org/10.3390/w17131875

AMA Style

Abdulrahman A, Algarni Z, Ghazouani N, Sammen SS, Amari A, Scholz M. Enhancing Photocatalytic Activity with the Substantial Optical Absorption of Bi2S3-SiO2-TiO2/TiO2 Nanotube Arrays for Azo Dye Wastewater Treatment. Water. 2025; 17(13):1875. https://doi.org/10.3390/w17131875

Chicago/Turabian Style

Abdulrahman, Amal, Zaina Algarni, Nejib Ghazouani, Saad Sh. Sammen, Abdelfattah Amari, and Miklas Scholz. 2025. "Enhancing Photocatalytic Activity with the Substantial Optical Absorption of Bi2S3-SiO2-TiO2/TiO2 Nanotube Arrays for Azo Dye Wastewater Treatment" Water 17, no. 13: 1875. https://doi.org/10.3390/w17131875

APA Style

Abdulrahman, A., Algarni, Z., Ghazouani, N., Sammen, S. S., Amari, A., & Scholz, M. (2025). Enhancing Photocatalytic Activity with the Substantial Optical Absorption of Bi2S3-SiO2-TiO2/TiO2 Nanotube Arrays for Azo Dye Wastewater Treatment. Water, 17(13), 1875. https://doi.org/10.3390/w17131875

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop