Next Article in Journal
Comparative Analysis of Five Numerical Methods and the Whale Optimization Algorithm for Wind Potential Assessment: A Case Study in Whittlesea, Eastern Cape, South Africa
Previous Article in Journal
Study and Effect of Agitation on Kojic Acid Production by Aspergillus oryzae in Liquid Fermentation
Previous Article in Special Issue
Decolorization of Rhodamine B by Fenton Processes Enhanced by Cysteine and Hydroxylamine Reducers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Development of a New Bi12ZnO20/AgI Heterosystem for the Degradation of Dye-Contaminated Water by Photocatalysis Under Solar Irradiation: Synthesis, Characterization and Kinetics

1
Laboratory of Transfer Phenomena, Faculty of Mechanical and Process Engineering (USTHB), B.P.32, El-Alia, Algiers 16000, Algeria
2
Applied Organic Synthesis Laboratory, Faculty of Exact and Applied Sciences, University Oran 1, Oran 31000, Algeria
3
UMR IPREM (Institut des Sciences Analytiques et de Physico-Chimie Pour l’Environnement et les Matériaux), Université de Pau et des Pays de l’Adour, CNRS—Technopôle Helioparc, 2 Avenue du Président Pierre Angot, 64053 Pau, France
*
Authors to whom correspondence should be addressed.
Processes 2025, 13(5), 1342; https://doi.org/10.3390/pr13051342
Submission received: 26 March 2025 / Revised: 23 April 2025 / Accepted: 25 April 2025 / Published: 27 April 2025

Abstract

:
This study explores the efficiency of heterogeneous photocatalysis in wastewater treatment, which is recognized for inducing significant rates of degradation and mineralization of various contaminants, including dyes. The study focuses on the development of an innovative composite via a combination of the sillenite type semiconductor Bi12ZnO20 and the halide-type semiconductor AgI. Both semiconductors were synthesized via co-precipitation, and their phases were identified using X-ray diffraction and characterized by scanning electron microscopy, Raman spectroscopy, Brunauer–Emmett–Teller analysis for specific surface area, UV–Visible diffuse reflectance spectroscopy, and the point of zero charge. The evaluation of the photocatalytic activity of the Bi12ZnO20/AgI heterosystem was carried out by monitoring the degradation process of Basic Blue 41 (BB41) under solar irradiation conditions. The results of this study revealed that the Bi12ZnO20/AgI heterosystem achieved the efficient degradation of BB41, with a removal rate of 98% after 150 min of treatment. The mineralization study showed that the TOC value decreased from 19.89 mg L−1 to 6.87 mg L−1, indicating that a significant portion of BB41 was mineralized. Via kinetic research, it was established that the degradation process followed a pseudo-first-order mechanism. Furthermore, recycling tests showed that the synthesized heterostructures maintained good structural stability and acceptable reusability over several cycles. These findings highlight the potential of heterogeneous photocatalysis as a promising approach to addressing environmental challenges associated with azo dyes.

1. Introduction

One of the major causes contributing to the scarcity of drinking water and freshwater is the degradation of the aquatic environment due to pollution. The deterioration of water quality is primarily caused by human activities, whether industrial, medical, agricultural, or domestic [1]. To date, industrial pollution ranks among the main factors significantly impacting the degradation of aquatic ecosystems [2]. The textile industry has become one of the leading contributors to the pollution of surface and groundwater reservoirs. The production and use of synthetic dyes for fabric dyeing have thus grown into massive industries. More than 10,000 pigments and dyes are used in the textile sector [3]. Each year, approximately 60,000 tons of dyes, nearly 80% of which are azo dyes, are released into the environment worldwide as waste [4]. Even at low concentrations (≤1 ppm), textile dyes can significantly harm aquatic ecosystems due to their high chemical oxygen demand (COD) and biological oxygen demand (BOD). These metrics indicate the potential for oxygen depletion in water bodies, which is critical for aquatic life. Most textile dyes are not only toxic but also highly resistant to biodegradation [5]. Their complex chemical structures prevent effective breakdown by microorganisms, leading to sustained high BOD levels in water. The elevated BOD levels resulting from the presence of these non-biodegradable dyes can lead to hypoxic conditions, which are detrimental to fish and other aquatic organisms. Azo dyes account for 50% of the dyes used in the textile industry. This type of dye is known for its persistence in the environment and the complexity of its degradation [6]. Currently, 2000 types of azo dyes are used worldwide, with 70,000 tons produced annually [7]. These dyes contain a double bond between two nitrogen atoms (-N=N-) [8]. Basic Blue 41 was selected as the model azo dye for this study, as it is among the most widely used dyes in the textile industry [9]. Furthermore, this dye is classified as a substance with a high level of toxicity [10].
In response to this challenge, research must focus on developing economically viable and competitive processes compared to conventional water treatment technologies. Heterogeneous photocatalysis emerges as a promising alternative for wastewater purification [11,12,13,14]. Within the framework of advanced oxidation processes (AOPs), the application of heterogeneous photocatalysis stands out as an optimal solution for the removal of organic pollutants, primarily due to its ability to efficiently harness natural and renewable solar energy [15]. This approach is based on the principle of photoexcitation of a semiconductor material, achieved through a light source with sufficient energy (hʋ ≥ Eg). Several high-performance semiconductor nanomaterials have been developed and utilized for this purpose. Among these photocatalysts, bismuth oxide-based semiconductors have been attracting increasing attention from researchers due to their numerous applications in photocatalysis, as well as their unique properties, such as chemical and thermal stability, strong visible light absorption, rapid interfacial charge separation, and low optical band gap energy [16,17].
The present study focuses on the sillenite family, characterized by the general formula Bi12MO20 [18], as a promising new series of bismuth oxide-based photocatalysts, where M (Zn, Ti, Co, Si, Ge,…) is a tetravalent ion. The crystalline structure of sillenite can be described as an interconnection between MO4 tetrahedra and BiO8polyhedra through oxygen atoms [19], as shown in Figure 1. The high recombination rate of electron-hole pairs in semiconductors is a significant limitation to their efficiency as photocatalysts, as well as their practical applications in pollution remediation [20]. Consequently, research is focused on overcoming this challenge by exploring innovative strategies. This research aims to develop a new photocatalyst with a high photocatalytic performance for the elimination of an organic pollutant (textile dye) in an aqueous solution, with the perspective of providing an economically and ecologically viable solution. One crucial approach involves combining these photocatalysts with other semiconductor nanomaterials, thereby forming heterosystem photocatalysts. Our choice focused on combining a zinc oxide sillenie (Bi12ZnO20) with silver iodide (AgI), which efficiently absorbs visible light [21]. This selection aims to maximize the advantages of each component to achieve more effective and environmentally friendly pollution remediation. The synthesized Bi12ZnO20/AgI heterosystem was used for the photodegradation of a textile dye (Basic Blue 41, BB41). Various characterization techniques were employed to determine the structural, morphological, and optical behavior of the nanomaterials and the developed heterosystem, notably TGA-DSC, XRD, SEM, Raman, optical band gap, and pHpzc. Parametric analysis was performed to study pollutant elimination in detail, with the aim of optimizing the operating conditions to ensure efficient removal. To our knowledge, this is the first study to report both the synthesis of this specific heterostructure and its successful application for BB41 removal under natural sunlight conditions, offering a sustainable and cost-effective alternative to existing photocatalytic systems.

2. Materials and Methods

2.1. Materials

All necessary precautions were taken before the start of our study to prevent any contamination of the raw materials and reagents used. The chemicals used are as follows: bismuth nitrate pentahydrate Bi(NO3)3·5H2O (98.5%, Chem-Lab Algiers, Algeria), zinc acetate dihydrate (CH3COO)2Zn·2H2O (99.5%, Chem-Lab), silver nitrate AgNO3 (99.8%, Biochem Chemo-pharma Algiers, Algeria), potassium iodide KI (99%, Chem-Lab), and nitric acid HNO3 (65%, Chem-Lab). These compounds were used for the synthesis of semiconductors as well as the Bi12ZnO20/AgI heterosystem. Additionally, HCl (37%, Biochem Chemo-pharma) and NaOH (98%, Biochem Chemo-pharma) were also used to adjust the pH of the different solutions. Double-distilled water was used for solution preparation, and Basic Blue 41 C20H26N4O6S2 was supplied by the textile industry Fital.

2.2. Synthesis of Semiconductors and Preparation of the Heterosystem

The sillenite Bi12ZnO20 was synthesized via the co-precipitation method [22] using stoichiometric amounts of Bi(NO3)3·5H2O and (CH3COO)2Zn·2H2O. A precise quantity of each reagent was measured and then separately dissolved in distilled water with the addition of 5% nitric acid (HNO3), followed by stirring. The two homogeneous solutions were then mixed under continuous stirring, and NaOH was added dropwise until a pH of 12 was reached. The resulting suspension was left to settle for 24 h in a closed environment to allow complete precipitation. The precipitate was regularly washed with distilled water until a pH of 7 was reached; then, it was filtered and dried in an oven at 100 °C for 24 h. The dried solid was ground using an agate mortar. Finally, the obtained powder was subjected to calcination at 850 °C for 4 h (this temperature was selected based on the results of the TG-DSC analysis) to obtain the desired zinc oxide sillenite phase. Figure 2 illustrates the key steps of the sillenite formation process.
Silver iodide was synthesized by separately dissolving 16.98 g of AgNO3 and 16.6 g of KI in 100 mL of distilled water under constant stirring for 20 to 30 min. The KI solution was then gradually added to the AgNO3 solution. The resulting light green precipitate was filtered, dried at 80 °C for 48 h, and then ground in an agate mortar.
The Bi12ZnO20/AgI heterosystem was obtained through a simple physical mixing process. The synthesized zinc oxide sillenite was combined with AgI in a ratio of 75% Bi12ZnO20 and 25% AgI. The two powders were carefully homogenized and then ground in an agate mortar until a fine and homogeneous powder was obtained.

2.3. Characterization Techniques

To deepen our understanding of the physicochemical properties as well as the structural, morphological, and optical behavior of the synthesized nanomaterials, various characterization techniques were employed. The temperature required for the formation of the sillenite zinc oxide phase was determined using thermogravimetric analysis coupled with differential scanning calorimetry (TGA-DSC); this was conducted with a Setaram KEP TECHNOLOGIES analyzer. The identification of the crystalline phases of the nanomaterials was carried out by X-ray diffraction (XRD) using a PANalytical Empyrean diffractometer equipped with monochromatic copper radiation Cu Kα (λ = 1.540598 Å). We scanned within the 2θ range (5–90°). Surface morphology analysis was performed using scanning electron microscopy (SEM) with a JEOL JCM 6000 microscope. The synthesized samples were analyzed by Raman spectroscopy using a RENISHAW Centrus 0LL824 microscope system. The measurements were performed with a 532 nm excitation laser, scanning the spectral range between 5 and 700 cm−1. Surface area measurements were conducted via nitrogen physisorption analysis employing the BET (Brunauer–Emmett–Teller) method. The experiments were performed at a set liquid nitrogen temperature (77.35 K) using N2 gas with a molecular cross-sectional area of 16.2 Å2 and a liquid density of 0.808 g/cm³. To enhance the electrical conductivity of the samples and optimize image clarity, a gold coating was applied via metallization for 2 min. The optical band gap energy (Eg) was determined using UV–Visible diffuse reflectance spectroscopy with a SPECORD 210 PLUS spectrophotometer. This instrument operates within a wavelength range of 250 to 1100 nm at room temperature, with a scanning speed of 10 nm s−1, and the point of zero charge (pHPZC) of the heterosystem was determined following the method reported by Ferro Garcia [23].

2.4. Study of the Photodegradation of Basic Blue 41

To assess the ability of the synthesized heterosystem to adsorb BB41, a kinetic study was conducted to determine the adsorption equilibrium, thereby facilitating subsequent photocatalytic process. Adsorption was performed by mixing 0.2 g of the photocatalyst at a dosage of 1 g L−1 in 200 mL of the polluted solution with an initial concentration of 10 mg L−1. The solution was continuously stirred in the dark at room temperature for 4 h using a magnetic stirrer at an average speed of 350 rpm to establish adsorption–desorption equilibrium while preventing any unwanted photocatalytic activity by the semiconductors.
All photocatalytic experiments were conducted in an open batch system during peak sunlight hours (10:00 AM to 2:00 PM) on clear days in June and July to ensure consistent solar irradiation intensity. The reactors, which had previously undergone the adsorption process, were exposed to solar irradiation; the average light source intensity was approximately 910 W·m−2. The solution was continuously stirred to maintain mixture homogeneity. At regular time intervals, 5 mL aliquots were collected, protected from light using opaque packaging, and then centrifuged at 4500 rpm for 15 min before being analyzed by UV–Visible spectrophotometry at a wavelength of 610 nm (Specord 200 Plus). Photolysis experiments were also performed to assess the influence of direct solar irradiation on pollutant degradation in the absence of the photocatalyst. All photocatalytic experiments were conducted in 250 mL cylindrical glass reactors (68 mm diameter × 90 mm height) under standardized conditions.
The photocatalytic degradation efficiency was evaluated using Equation (1):
R ( % ) = C 0 C t C 0 × 100
where C0 is the initial concentration of BB41 after adsorption (mg L−1) and Ct is the concentration at time t (mg L−1).

3. Results and Discussion

3.1. Physicochemical Characterization of Nanomaterials

3.1.1. Evaluation of Thermal Properties

The thermal decomposition of the Bi12ZnO20 mass occurs primarily in three stages during heating in the temperature range of 27 to 1000 °C (Figure 3). The first weight loss (3.75%) is mainly attributed to evaporation, induced by the breaking of molecular water bonds, as well as the elimination of volatile compounds present in the precursors (CH3COOC)2Zn·2H2O and Bi(NO3)3·5H2O. A second weight loss of 1.5% is observed between 250 and 350 °C, due to the decomposition of organic matter, including the acetate CH3COO, followed by a mass loss of 1.1%, which is attributed to the complete desorption of water. Consequently, it can be concluded that the material’s mass decreases as the temperature increases during thermal analysis [24,25]. The DSC curve reveals a significant endothermic peak at 717 °C, attributed to the decomposition of water molecules, followed by an exothermic peak at 850 °C, associated with the formation of the crystalline structure of Bi12ZnO20. The calcination temperature determined in this study is close to the value reported in the literature [26].

3.1.2. Identification of the Crystalline Phase

The confirmation of the crystalline phases of the synthesized nanomaterials and the formed heterosystem was carried out using X-ray diffraction (XRD), followed by XRD data processing with HighScore. The results of this analysis are presented in Figure 4.
The observed diffraction peaks (Figure 4b) correspond to the standard pattern of a zinc oxide sillenite, indexed in terms of cubic symmetry with space group I23 (PDF 01-078-1325) [27]. The presence of three intense diffraction peaks at 27.619°, 30.316°, and 32.822°, respectively, associated with the crystallographic planes (310), (222), and (321), confirms the well-established crystallization of the sillenite and the formation of a pure phase. The average crystallite size of Bi12ZnO20 was estimated using Scherrer’s formula [28], yielding a size of 53 nm.
D = K   λ β cos θ
The X-ray density was also evaluated (Equation (3)). The results of these parameters are listed in Table 1.
ρ = Z M N A V
where k is a dimensional form factor (=0.9), β is the wavelength of the X-ray (radian), and θ is the Braggs angle. Z is the number of formula units in a single unit cell (Z = 2), M is the molecular weight of Bi12ZnO20 (2893.12 g mol−1), NA is Avogadro’s number (6.02 × 1023 mol−1) and V is the unit cell volume (1062.77 × 106 pm3).
The specific surface area was calculated using the crystallite diameter and the X-ray density of BZO (Equation (4)) [29].
S = 6   × 10 3 D   × ρ
A previous study [25] on the same photocatalyst, synthesized using a different sol–gel method, reported a specific surface area of 11.22 m2 g−1, which is lower than the value obtained in our research (12.54 m2 g−1). This discrepancy highlights the efficiency of our synthesis method, co-precipitation, while the observed difference is primarily attributed to variations in crystallite size.
The diffraction data of the AgI sample (Figure 4a) reveal the presence of peaks at 2θ values of 22.41°, 23.63°, 39.28°, and 46.28°, corresponding to the crystallographic planes (010), (002), (110), and (112), respectively [30]. All these peaks are indexed to a hexagonal unit cell, in accordance with the JCPDS card No. 98-007-9678. Furthermore, the diffractogram of Bi12ZnO20/AgI (Figure 4c) confirms the successful formation of the heterosystem, highlighting the presence of all the crystallization planes of both semiconductors, Bi12ZnO20 and AgI, without any shift or deformation.

3.1.3. Surface Analysis

The surface morphology of the synthesized nanomaterials, Bi12ZnO20 and AgI, and the formed Bi12ZnO20/AgI heterosystem was examined using SEM. The results of the imaging are presented in Figure 5. The morphology of Bi12ZnO20 sillenite is characterized by a regular and homogeneous arrangement of small particles (aggregates), taking the form of spheres and nanotubes, as shown in the SEM images (Figure 5a). The average size of these particles is estimated at 1.6 μm. Pure AgI nanoparticles exhibit both hexagonal and spherical shapes (Figure 5b). This observation was made at magnifications of 20 μm and 10 μm, revealing an average particle size of 5 μm. The presence of AgI particles and their homogeneous distribution on the surface of the sillenite confirms the successful formation of the Bi12ZnO20/AgI heterosystem (Figure 5c).

3.1.4. Raman Spectroscopy

The Raman spectrum analysis of Bi12ZnO20 (Figure 6) reveals 10 distinct spectral bands spanning a broad range of 5 to 700 cm−1. These bands show excellent correspondence with the characteristic vibrational modes of sillénite [18,31], thereby confirming its crystalline phase. A detailed comparison between the Raman peaks observed in our study and those reported in the literature is presented in Table 2.
The intense peaks observed below 200 cm−1 can be attributed to Bi-O bond vibrations, while the broad peaks between 200 and 600 cm−1 correspond to oxygen atom vibrations and breathing modes. The weak Raman mode detected at 630 cm−1 is associated with vibrational stretching within ZnO4 tetrahedra [26]. The Raman analysis of AgI nanomaterials revealed two distinct vibrational modes at 83 and 103 cm−1. These modes correspond to the symmetric stretching vibrations of silver iodide and are characteristic of the β-AgI polymorph, specifically its E2 and A1 transitions [32]. Remarkably, the Raman spectrum of the BZO/AgI heterosystem shows that the characteristic vibrational modes of both sillénite and AgI remain intact, exhibiting no peak shifts or distortions. This observation strongly suggests that both nanomaterials maintain their original crystalline structures despite being combined in the heterosystem.

3.1.5. BET Analysis

The nanomaterials’ specific surface area was determined through BET adsorption isotherm analysis, as graphically represented in Figure 7.
The BET surface area measured for AgI was remarkably low (0.116 m2 g−1). This primarily results from particle agglomeration, that reduces the available gas adsorption surface. Furthermore, semiconductors typically exhibit intrinsically low specific surface areas due to their crystalline structure and atomic arrangement. Interestingly, the sillénite Bi12ZnO20 showed a significant surface area increase from 0.757 m2 g−1 to 1.076 m2 g−1 upon forming the Bi12ZnO20/AgI heterosystem. This enhancement led to improved pollutant degradation efficiency when sillénite was incorporated into the heterostructure. Notably, the SBET value of zinc sillénite falls within the typical range reported for sillénites (~0.8 m2 g−1), confirming the consistency of our measurements with the established literature values [33].

3.1.6. Optical Properties Analysis

The analysis of the direct transition (αhʋ)2 and indirect transition (αhʋ)1/2 curves as a function of photon energy hʋ allowed us to determine the band gap energy Eg of the nanomaterials (Equation (5)) [31].
α h ʋ n = K h ʋ E g
where α is the absorption coefficient, is the energy of the incident photon (hʋ = 1240/λ), K is the energy constant, and n is the electronic transition. The following values were obtained (Figure 8): 2.08 eV for AgI, 2.78 eV for Bi12ZnO20, and 1.58 eV for the Bi12ZnO20/AgI heterosystem. These results demonstrate that all photocatalysts exhibit strong absorption in the visible region, making them effective when exposed to visible light irradiation. Therefore, solar energy was used as an irradiation source for the application of these photocatalysts in heterogeneous photocatalysis. A significant reduction in the optical band gap was observed following the formation of Bi12ZnO20/AgI. This can be explained by the intimate contact between the two semiconductors, which facilitates interfacial charge transfer during photocatalytic excitation. The optical band gap values of AgI and Bi12ZnO20 are consistent with those reported in the literature [32,33].

3.1.7. Determination of the Point of Zero Charge

The isoelectric point (pHPZC) of Bi12ZnO20/AgI was determined by plotting the evolution of the final pH as a function of the initial pH, as illustrated in Figure 9. The final pH of the recovered NaCl solution gradually increases with the rise in initial pH, reaching an isoelectric point at pH~8.45. This indicates that the Bi12ZnO20/AgI heterosystem exhibits a basic character, demonstrating that when pH values are lower than pH pHPZC, the surface of the heterosystem acquires a positive charge, whereas when pH values are higher than pH pHPZC, the surface carries a negative charge. These results provide valuable insights into the attraction and repulsion between the heterosystem’s surface and the molecules of the targeted pollutant.

3.2. Evaluation of the Adsorption Effect on the Degradation of Basic Blue 41

Adsorption shows a removal rate of 30%, which is relatively low in terms of adsorption efficiency (Figure 10). These results are consistent with those of previous studies [11]. It can also be observed that after approximately 120 min, the adsorption equilibrium is reached. For all subsequent photocatalytic experiments, it is essential to first conduct a 120 min adsorption step to enhance the efficiency of the photocatalysis process.

3.3. Study of the Photodegradation of Basic Blue 41

A parametric study was conducted to evaluate the effect of mass ratio, pH, photocatalyst dosage, and initial pollutant concentration on the photocatalytic degradation of BB41. The objective was to optimize the best operating conditions in order to maximize photocatalytic performance.

3.3.1. Influence of the Mass Ratio of the Bi12ZnO20/AgI Heterosystem

The histogram in Figure 11 shows that 69% and 49% of BB41 was degraded using Bi12ZnO20 and AgI, respectively. The addition of AgI to the Bi12ZnO20 surface led to a significant improvement in degradation efficiency, reaching an optimal value of 80% for a mixture of 75% BZO and 25% AgI. The efficiency of the heterosystem increases with the Bi12ZnO20 concentration, clearly highlighting the significant advantages of the Bi12ZnO20/AgI heterosystem for treating dye-contaminated water. This enhancement is attributed to the interfacial charge transfer phenomenon between AgI and Bi12ZnO20. This transfer facilitates the migration of photogenerated charge carriers until the Fermi levels of both semiconductors align, creating an electric field between their surfaces, which prevents recombination. As a result, the degradation efficiency of BB41 under visible light is significantly improved. These findings suggest that AgI could be a promising choice for photosensitizing Bi12ZnO20 during junction formation.

3.3.2. Influence of pH on the Photodegradation of Basic Blue 41

The effect of pH was studied by varying the pH range from an acidic medium (pH 5.3) to a basic medium (pH 10) in a suspension containing 10 mg L−1 of BB41 and 1 g L−1 of Bi12ZnO20/AgI, as shown in Figure 12. The results reveal a low dye removal efficiency at pH~7 and isoelectric point~5.3. These are characterized by a slower degradation rate. This behavior can be attributed to the weak adsorption of dye molecules on the Bi12ZnO20/AgI surface due to the similarity in surface charge. Since Bi12ZnO20/AgI is positively charged at pH ˂ 8.45, it experiences electrostatic repulsion with the cationic dye molecules, thereby reducing the degradation efficiency. The maximum level of BB41 degradation was observed at pH~10, achieving a 99% removal efficiency. However, a slight decrease in decolorization efficiency was noted at pH~8, with a yield of 97% after 150 min of treatment. This is due to the electrostatic attraction between the negatively charged Bi12ZnO20/AgI surface and the positively charged cationic BB41 molecules, which enhances dye adsorption onto the heterosystem surface. To maintain ecological balance and the health of aquatic ecosystems [34], pH~8 was selected as the optimal value. These results are consistent with the literature findings, confirming that BB41 photodegradation under UV and visible light irradiation is favored in a basic medium, particularly at pH values ranging from 8 to 10 [35,36].

3.3.3. Influence of Bi12ZnO20/AgI Dose on Photodegradation of Basic Blue 41

According to the results illustrating the variation in BB41 decolorization rate as a function of irradiation time (Figure 13), it is observed that the decolorization efficiency increases from 66% to 98% with the increase in photocatalyst dosage, reaching a maximum value at 0.75 g L−1. This improvement is further enhanced by the small average crystallite size (~53 nm), which promotes both an increase in reactive sites and a larger photon absorption surface, thereby enhancing the efficiency of the photocatalytic treatment process. However, beyond this optimal dosage, the degradation decreases by 8% due to the saturation of the solution with the Bi12ZnO20/AgI heterosystem. These observations are consistent with the work of Y. Benrighi and colleagues [35], who reported that increasing the CoCr2O4 dosage from 0.125 to 0.5 g L−1 enhanced BB41 degradation from 63% to 99% under visible light irradiation. They also observed that this maximum decolorization rate decreased to 86% when the photocatalyst dosage exceeded 1 g L−1.

3.3.4. Influence of the Initial Concentration of Basic Blue 41

Under visible light irradiation, the concentration of BB41 appears to have a negative effect on the photocatalytic degradation process occurring the surface of the Bi12ZnO20 /AgI heterosystem. A higher initial concentration leads to a decrease in degradation efficiency. The decolorization rate increases from 20% to 98% as the dye concentration decreases from 50 to 10 mg L−1 (Figure 14a). At an initial concentration of 10 mg L−1, the degradation process is remarkably fast, reaching a maximum efficiency of 98% in just 150 min of treatment. These results can be explained as follows: At high concentrations, the increased presence of dye creates an optical shielding effect, making the medium opaquer. This opacity prevents incident photons from reaching the surface of Bi12ZnO20/AgI, thereby reducing the generation of radicals responsible for pollutant decomposition. Conversely, at lower concentrations, incident photons can more easily reach the surface of the heterosystem particles. This enhances the interaction between pollutant molecules and the reactive species generated during photocatalytic activation. Many researchers have also observed a similar trend in the degradation of BB41 [37,38].
Using the Bi12ZnO20/AgI heterosystem, the evolution of the UV-Vis absorption spectrum during the photodegradation of Basic Blue 41 as a function of irradiation time is presented in Figure 14b. A decrease in absorbance in the visible region of the absorption spectrum is observed over time due to the fading of the suspension color. This phenomenon is attributed to the breakdown of the azo group (-N=N-), which is a characteristic feature of the BB41 molecule [39].

3.4. Photocatalytic Degradation Kinetics

The kinetic study of the photodegradation reaction of BB41 as a function of the initial concentration was conducted by varying it from 10 to 50 ppm, while keeping the optimal conditions of pH (~8) and photocatalyst dose (0.75 g L−1) constant. Two kinetic models were used [22]: the pseudo-first-order model (Equation (6)) and the pseudo-second-order model (Equation (7)).
ln C t C 0 = k 1 t
1 C t = 1 C 0 + k 2 t
where k1 is the apparent first-order kinetic constant of degradation (min −1), k2 is the apparent second-order kinetic constant of degradation (L mg−1 min−1), C0 is the initial concentration of BB41 after adsorption (mg L−1), Ct is the concentration at time t (mg L−1), and t is irradiation time (min).
The graphical representation of these two models allows us to determine the kinetic order of the photocatalytic degradation reaction of BB41. The experimental results are presented in Figure 15 and the kinetic parameters of both models are summarized in Table 3.
The kinetic curves, representing ln(C0/C) as a function of irradiation time for different initial concentrations, exhibit good linearity, indicating that the photodegradation kinetics are described well by the pseudo-first-order model, with correlation coefficients very close to 1 (R2~0.94). These results are consistent with previous studies [39]. The initial rate r0 (Equation (8)) and the half-reaction time t1/2 (Equation (9)) can be calculated using the apparent rate constant k1.
r 0 = k a p p   C 0
t 1 / 2 = l n 2 k 1
The results in Table 4 show that the apparent rate constant, kapp, decreases with the increase in initial dye concentration C0 due to the screening effect. This leads to a reduction in the solution’s permeability to visible radiation, preventing incident photons from reaching the surface of Bi12ZnO20/AgI and thereby reducing the generation of radicals responsible for dye decomposition. However, the half-reaction time (t1/2) increases as the initial concentration C0 increases.

3.5. Analysis of Dye Mineralization

To evaluate the mineralization rate of organic matter in an aqueous solution, the temporal variation in total organic carbon was monitored. This study was conducted under the optimal conditions as determined by a parametric study of BB41 photodegradation. The results of this study are illustrated in Figure 16. A significant mineralization rate of 65.5% was observed, while the degradation rate determined by UV-Vis analysis reached 98%. After 360 min of treatment, the TOC value decreased from 19.89 mg L−1 to 6.87 mg L−1, indicating that a large portion of BB41 was mineralized. To achieve complete mineralization, the photocatalytic treatment process requires a sufficiently long duration to ensure the decomposition of BB41 organic molecules and the intermediate products formed.

3.6. Study of the Regeneration of the Bi12ZnO20/AgI Heterosystem

We conducted series of recycling tests using the Bi12ZnO20/AgI heterosystem under optimal reaction conditions for the photocatalytic degradation of BB41. After each experiment, the photocatalyst was separated from the aqueous suspension by filtration, cleaned with ethanol, then washed with distilled water to remove the adsorbed dye molecules [40], and finally dried at 80 °C for 24 h. After five recycling cycles (Figure 17), a decrease in degradation efficiency was observed, leading to a reduction in yield from 96% to 46% after 4 h of solar light irradiation. This indicates that the heterosystem is relatively stable and can retain 50% of its photocatalytic properties.
After the recycling experiments, the recovered Bi12ZnO20/AgI nanoparticles were characterized using X-ray diffraction and scanning electron microscopy coupled with energy-dispersive X-ray spectroscopy. Figure 18 illustrates the main results obtained. The XRD spectra of the Bi12ZnO20/AgI heterosystem before and after recycling (Figure 18a) exhibit peaks at similar 2θ positions, indicating the preservation of the main crystalline structure. SEM analysis reveals that the overall morphology of the nanoparticles remains remarkably unchanged (Figure 18b). However, a more intimate attachment between AgI and Bi12ZnO20 has been observed, which may contribute to enhancing the semiconductor performance of the heterosystem. The retention of all constituent elements of Bi12ZnO20/AgI after the recycling process (Figure 18c) confirms the purity, stability, and durability of the synthesized photocatalyst.

3.7. Comparison of BB41 Photodegradation with Previous Studies

It would be relevant at this stage to compare the photocatalytic performance of our heterostructure with that reported in previous studies on the photodegradation of Basic Blue 41. The main results are summarized in Table 5. These data confirm that the synthesized BZO/AgI heterostructure could represent a competitive solution for the treatment of textile wastewater under solar irradiation.

4. Conclusions

In this work, a Bi12ZO20/AgI heterosystem was prepared by combining sillenite-type zinc oxide and silver iodide. This heterosystem was used for the photodegradation of the textile dye Basic Blue 41. Various characterization techniques demonstrated that the prepared heterosystem exhibits significantly superior semiconductor properties compared to Bi12ZnO20 and AgI individually. the efficient degradation of BB41 with a removal rate of 98% was observed after 150 min of treatment under the following operating conditions: pH~8, dose = 0.75 g L−1, [BB41] = 10 mg L−1. Experimental results revealed that the photodegradation kinetics are perfectly described by the pseudo-first-order model. The recycling test demonstrated that the synthesized heterosystem retained its crystalline structure and morphology after use, while showing a moderate decrease in reuse capacity over successive cycles. In light of this limitation, future work will focus on the development and evaluation of alternative heterosystems with different compositions to enhance long-term photocatalytic activity and reusability. In summary, this Bi12ZnO20/AgI sillenite-based heterosystem has demonstrated remarkable photocatalytic efficiency, achieving the near-complete degradation of Basic Blue 41 dye. This exceptional performance positions the material as a highly promising solution for environmental remediation applications.

Author Contributions

S.M. contributed to experimental methods, interpretation, writing and review. M.B. data treatment, review and figures. E.M. analyzes, discussion and review. S.Z.: Ressources. S.E.I.L. supervising. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding authors.

Acknowledgments

The authors would like to express their sincere gratitude to all individuals and institutions who contributed to the completion of this work.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
BB41Basic Blue 41
BZOBi12ZnO20
TOCtotal organic carbon
XRDX-ray diffraction
SEMscanning electron microscopy
pHPZCpoint of zero charge
Egoptical band gap energy

References

  1. Madji, S.; Mekatel, E.; Belmedani, M.; Cosme, J.; Zouaoui, S.; Vieillard, J. Synthesis of an Original Bi12CoO20/ZnO Nanocomposites for Lead Adsorption. Mater. Sci. Eng. B 2025, 319, 118345. [Google Scholar] [CrossRef]
  2. Arif, A.; Malik, M.F.; Liaqat, S.; Asifa, A.; Mumtaz, K.; Afzal, A.; Ch, D.M.; Nisa, K.; Khurshid, F.; Arif, F.; et al. Water pollution and industries. Pure Appl. Biol. 2020, 9, 2214–2224. [Google Scholar] [CrossRef]
  3. Chang Chien, J.-R.; Joshiba Ganesan, J. Advancing Sustainable Approaches for the Removal and Recycling of Toxic Dyes from the Aquatic Environment. In Dye Chemistry-Exploring Colour from Nature to Lab; IntechOpen: London, UK, 2024. [Google Scholar]
  4. Liu, Q. Pollution and Treatment of Dye Waste-Water. IOP Conf. Ser. Earth Environ. Sci. 2020, 514, 052001. [Google Scholar] [CrossRef]
  5. Herath, I.S.; Udayanga, D.; Jayasanka, D.J.; Hewawasam, C. Textile dye decolorization by white rot fungi—A review. Bioresour. Technol. Rep. 2024, 25, 101687. [Google Scholar] [CrossRef]
  6. Ali, H. Biodegradation of Synthetic Dyes—A Review. Water Air Soil Pollut. 2010, 213, 251–273. [Google Scholar] [CrossRef]
  7. Saba, B.; Kjellerup, B.V.; Christy, A.D. Eco-friendly bio-electro-degradation of textile dyes wastewater. Bioresour. Technol. Rep. 2021, 15, 100734. [Google Scholar] [CrossRef]
  8. Mojiri, A.; Zhou, J.L.; KarimiDermani, B.; Razmi, E.; Kasmuri, N. Anaerobic Membrane Bioreactor (AnMBR) for the Removal of Dyes from Water and Wastewater: Progress, Challenges, and Future Perspectives. Processes 2023, 11, 855. [Google Scholar] [CrossRef]
  9. Liu, Y.; Li, C.; Bao, J.; Wang, X.; Yu, W.; Shao, L. Degradation of Azo Dyes with Different Functional Groups in Simulated Wastewater by Electrocoagulation. Water 2022, 14, 123. [Google Scholar] [CrossRef]
  10. Lanfredi, S.; Nobre, M.A.L.; Poon, P.S.; Matos, J. Hybrid Material Based on an Amorphous-Carbon Matrix and ZnO/Zn for the Solar Photocatalytic Degradation of Basic Blue 41. Molecules 2019, 25, 96. [Google Scholar] [CrossRef] [PubMed]
  11. Nouri, L.; Hemidouche, S.; Boudjemaa, A.; Kaouah, F.; Sadaoui, Z.; Bachari, K. Elaboration and characterization of photobiocomposite beads, based on titanium (IV) oxide and sodium alginate biopolymer, for basic blue 41 adsorption/photocatalytic degradation. Int. J. Biol. Macromol. 2020, 151, 66–84. [Google Scholar] [CrossRef]
  12. Ghosh, S.K. (Ed.) Waste Water Recycling and Management; Springer: Singapore, 2019. [Google Scholar]
  13. Aziz, F.F.A.; Jalil, A.A.; Triwahyono, S.; Mohamed, M. Controllable structure of fibrous SiO2–ZSM-5 support decorated with TiO2 catalysts for enhanced photodegradation of paracetamol. Appl. Surf. Sci. 2018, 455, 84–95. [Google Scholar] [CrossRef]
  14. Abdu, M.; Tibebu, S.; Babaee, S.; Worku, A.; Msagati, T.A.M.; Nure, J.F. Optimization of photocatalytic degradation of Eriochrome Black T from aqueous solution using TiO2-biochar composite. Results Eng. 2025, 25, 104036. [Google Scholar] [CrossRef]
  15. Byrne, C.; Subramanian, G.; Pillai, S.C. Recent advances in photocatalysis for environmental applications. J. Environ. Chem. Eng. 2018, 6, 3531–3555. [Google Scholar] [CrossRef]
  16. Boochakiat, S.; Inceesungvorn, B.; Nattestad, A.; Chen, J. Bismuth-Based Oxide Photocatalysts for Selective Oxidation Transformations of Organic Compounds. ChemNanoMat 2023, 9, e202300140. [Google Scholar] [CrossRef]
  17. Tho, N.T.M.; Khanh, D.N.N.; Thang, N.Q.; Lee, Y.-I.; Phuong, N.T.K. Novel reduced graphene oxide/ZnBi2O4 hybrid photocatalyst for visible light degradation of 2,4-dichlorophenoxyacetic acid. Environ. Sci. Pollut. Res. 2020, 27, 11127–11137. [Google Scholar] [CrossRef] [PubMed]
  18. Burkov, V.I.; Gorelik, V.S.; Egorysheva, A.V.; Kargin, Y.F. Laser Raman Spectroscopy of Crystals with the Structure of Sillenite. J. Russ. Laser Res. 2001, 22, 243–267. [Google Scholar] [CrossRef]
  19. Lima, A.F.; Lalic, M.V. First-principles study of the BiMO4 antisite defect in the Bi12MO20 (M=Si, Ge, Ti) sillenite compounds. J. Phys. Condens. Matter. 2013, 25, 495505. [Google Scholar] [CrossRef]
  20. Chen, J.; Zhan, J.; Zhang, Y.; Tang, Y. Construction of a novel ZnCo2O4/Bi2O3 heterojunction photocatalyst with enhanced visible light photocatalytic activity. Chin. Chem. Lett. 2019, 30, 735–738. [Google Scholar] [CrossRef]
  21. Jin, X.; Dasheng, G.; Shuang, C.; Xiaohua, W.; Ningning, L. Enhanced visible light photocatalytic activity of AgI/TiO2 composite fabricated by a grinding method. Water Qual. Res. J. 2019, 54, 257–264. [Google Scholar] [CrossRef]
  22. Madji, S.; Belmedani, M.; Mekatel, E.; Trari, M. Co-precipitation synthesis and characterization of the sillenite Bi12CoO20: Application to photoactivity for water treatment on the heterojunction with ZnO. J. Mater. Sci. Mater. Electron. 2023, 34, 2093. [Google Scholar] [CrossRef]
  23. Ferro-García, M.A.; Rivera-Utrilla, J.; Bautista-Toledo, I.; Moreno-Castilla, C. Adsorption of Humic Substances on Activated Carbon from Aqueous Solutions and Their Effect on the Removal of Cr(III) Ions. Langmuir 1998, 14, 1880–1886. [Google Scholar] [CrossRef]
  24. Nicholas, P. Cheremisinoff. 2—Thermal Analysis. In Polymer Characterization: Laboratory Techniques and Analysis; Noyes Publications: Park Ridge, NJ, USA, 1996; pp. 17–24. [Google Scholar] [CrossRef]
  25. Zhu, N.; Qian, F.; Xu, X.; Wang, M.; Teng, Q. Thermogravimetric Experiment of Urea at Constant Temperatures. Materials 2021, 14, 6190. [Google Scholar] [CrossRef]
  26. Baaloudj, O.; Nasrallah, N.; Kenfoud, H.; Algethami, F.; Modwi, A.; Guesmi, A.; Assadi, A.A.; Khezami, L. Application of Bi12ZnO20 Sillenite as an Efficient Photocatalyst for Wastewater Treatment: Removal of Both Organic and Inorganic Compounds. Materials 2021, 14, 5409. [Google Scholar] [CrossRef]
  27. Wang, X.-L.; Xiao, Y.; Lv, Z.-J.; Yu, H.; Yang, Y.; Dong, X.-T. A novel 2D nanosheets self-assembly camellia-like ordered mesoporous Bi12ZnO20 catalyst with excellent photocatalytic property. J. Alloys Compd. 2020, 835, 155409. [Google Scholar] [CrossRef]
  28. Jagadeesh, C.; Sailaja, B.B.V. Photocatalytic Degradation of Rohdamine B Dye using CoBi2O4 Nanocatalyst and Effect of Various Operational Parameters. Cikitusi J. 2019, 6, 97–107. [Google Scholar]
  29. Baaloudj, O.; Nasrallah, N.; Bouallouche, R.; Kenfoud, H.; Khezami, L.; Assadi, A.A. High efficient Cefixime removal from water by the sillenite Bi12TiO20: Photocatalytic mechanism and degradation pathway. J. Clean. Prod. 2022, 330, 129934. [Google Scholar] [CrossRef]
  30. Zheng, H.; Chen, G.; Zhang, A.; Tan, Z.; Wang, R.; Wang, H.; Mei, Y.; Zhang, X.; Ran, J. Enhanced photocatalytic activity of Bi24O31Br10 microsheets constructing heterojunction with AgI for Hg0 removal. Sep. Purif. Technol. 2021, 262, 118296. [Google Scholar] [CrossRef]
  31. Hemmi, A.; Belmedani, M.; Mekatel, E.; Djaballah, A.M.; Sadaoui, Z.; Brahimi, B.; Trari, M. Characterization of MgAl2O4 semiconductor prepared by co-precipitation route: Application as photocatalyst for tetracycline degradation. J. Mater. Sci. Mater. Electron. 2025, 36, 217. [Google Scholar] [CrossRef]
  32. Manzoor, Q.; Farrukh, M.A.; Sajid, A. Optimization of lead (II) and chromium (VI) adsorption using graphene oxide/ZnO/chitosan nanocomposite by response surface methodology. Appl. Surf. Sci. 2024, 655, 159544. [Google Scholar] [CrossRef]
  33. Gu, M.; Hao, L.; Wang, Y.; Li, X.; Chen, Y.; Li, W.; Jiang, L. The selective heavy metal ions adsorption of zinc oxide nanoparticles from dental wastewater. Chem. Phys. 2020, 534, 110750. [Google Scholar] [CrossRef]
  34. Singh, V.; Ahmed, G.; Vedika, S.; Kumar, P.; Chaturvedi, S.K.; Rai, S.N.; Vamanu, E.; Kumar, A. Toxic heavy metal ions contamination in water and their sustainable reduction by eco-friendly methods: Isotherms, thermodynamics and kinetics study. Sci. Rep. 2024, 14, 7595. [Google Scholar] [CrossRef] [PubMed]
  35. Benrighi, Y.; Nasrallah, N.; Chaabane, T.; Sivasankar, V.; Darchen, A.; Baaloudj, O. Characterization of CoCr2O4 semiconductor: A prominent photocatalyst in the degradation of basic blue 41 from wastewater. Opt. Mater. 2021, 122, 111819. [Google Scholar] [CrossRef]
  36. Mahmoodi, N.M.; Abdi, J. Nanoporous metal-organic framework (MOF-199): Synthesis, characterization and photocatalytic degradation of Basic Blue 41. Microchem. J. 2019, 144, 436–442. [Google Scholar] [CrossRef]
  37. Mahmoodi, N.M.; Keshavarzi, S.; Ghezelbash, M. Synthesis of nanoparticle and modelling of its photocatalytic dye degradation ability from colored wastewater. J. Environ. Chem. Eng. 2017, 5, 3684–3689. [Google Scholar] [CrossRef]
  38. Kalpaklı, Y. The effect of in-situ Pr6O11 phase formation on photocatalytic Performance: Mono azo dye degradation. Inorg. Chem. Commun. 2024, 159, 111788. [Google Scholar] [CrossRef]
  39. Mahmoodi, N.M.; Abdi, J.; Oveisi, M.; Alinia Asli, M.; Vossoughi, M. Metal-organic framework (MIL-100 (Fe)): Synthesis, detailed photocatalytic dye degradation ability in colored textile wastewater and recycling. Mater. Res. Bull. 2018, 100, 357–366. [Google Scholar] [CrossRef]
  40. Emam, H.E.; Ahmed, H.B.; Gomaa, E.; Helal, M.H.; Abdelhameed, R.M. Recyclable photocatalyst composites based on Ag3VO4 and Ag2WO4 @MOF@cotton for effective discoloration of dye in visible light. Cellulose 2020, 27, 7139–7155. [Google Scholar] [CrossRef]
  41. Apostolopoulou, A.; Mahajan, S.; Sharma, R.; Stathatos, E. Novel development of nanocrystalline kesterite Cu2ZnSnS4 thin film with high photocatalytic activity under visible light illumination. J. Phys. Chem. Solids 2018, 112, 37–42. [Google Scholar] [CrossRef]
  42. Kafian, S.; Sadeghzadeh-Attar, A. Photocatalytic degradation of Basic Blue 41 dye under visible light over SrTiO3/Ag3PO4 hetero-nanostructures. Int. J. Appl. Ceram. Technol. 2022, 19, 3347–3357. [Google Scholar] [CrossRef]
  43. Abhayasimha, K.C.; Rao, C.S.; Nair, V. Combination of ensemble machine learning models in photocatalytic studies using nano TiO2-Ligninbased biochar. Chemosphere 2024, 352, 141326. [Google Scholar] [CrossRef]
Figure 1. Crystalline structure of sillenite Bi12MO20 (generated using Vesta software (version 3.5.8)).
Figure 1. Crystalline structure of sillenite Bi12MO20 (generated using Vesta software (version 3.5.8)).
Processes 13 01342 g001
Figure 2. Experimental procedure for synthesis of sillenite Bi12ZnO20.
Figure 2. Experimental procedure for synthesis of sillenite Bi12ZnO20.
Processes 13 01342 g002
Figure 3. Thermal analysis of sillenite Bi12ZnO20.
Figure 3. Thermal analysis of sillenite Bi12ZnO20.
Processes 13 01342 g003
Figure 4. X-ray diffraction patterns of the nanomaterials AgI (a), Bi12ZnO20 (b), and Bi12ZnO20/AgI (c).
Figure 4. X-ray diffraction patterns of the nanomaterials AgI (a), Bi12ZnO20 (b), and Bi12ZnO20/AgI (c).
Processes 13 01342 g004
Figure 5. SEM images of nanomaterials: Bi12ZnO20 (a), AgI (b), and Bi12ZnO20/AgI (c).
Figure 5. SEM images of nanomaterials: Bi12ZnO20 (a), AgI (b), and Bi12ZnO20/AgI (c).
Processes 13 01342 g005
Figure 6. Raman spectra of BZO, AgI and BZO/AgI.
Figure 6. Raman spectra of BZO, AgI and BZO/AgI.
Processes 13 01342 g006
Figure 7. BET surface analysis of AgI, Bi12ZnO20, and Bi12ZnO20/AgI.
Figure 7. BET surface analysis of AgI, Bi12ZnO20, and Bi12ZnO20/AgI.
Processes 13 01342 g007
Figure 8. Optical properties of semiconductor nanomaterials: AgI, Bi12ZnO20, and Bi12ZnO20/AgI.
Figure 8. Optical properties of semiconductor nanomaterials: AgI, Bi12ZnO20, and Bi12ZnO20/AgI.
Processes 13 01342 g008
Figure 9. Measurement of point of zero charge of Bi12ZnO20/AgI.
Figure 9. Measurement of point of zero charge of Bi12ZnO20/AgI.
Processes 13 01342 g009
Figure 10. The adsorption efficiency of BB41 on Bi12ZnO20/AgI under experimental conditions (pH~5.3, dose = 1 g L−1 and [BB41] = 10 mg L−1).
Figure 10. The adsorption efficiency of BB41 on Bi12ZnO20/AgI under experimental conditions (pH~5.3, dose = 1 g L−1 and [BB41] = 10 mg L−1).
Processes 13 01342 g010
Figure 11. Effect of mass ratio of Bi12ZnO20/AgI heterosystem on photodegradation of BB41 (pH~5.3, [BB41] = 10 mg L−1, dose = 1 g L−1, irradiation time = 4 h).
Figure 11. Effect of mass ratio of Bi12ZnO20/AgI heterosystem on photodegradation of BB41 (pH~5.3, [BB41] = 10 mg L−1, dose = 1 g L−1, irradiation time = 4 h).
Processes 13 01342 g011
Figure 12. Impact of pH on absorption spectrum of BB41 molecule ([BB41] = 10 mg L−1, dose = 1 g L−1).
Figure 12. Impact of pH on absorption spectrum of BB41 molecule ([BB41] = 10 mg L−1, dose = 1 g L−1).
Processes 13 01342 g012
Figure 13. Influence of photocatalyst dose on photocatalytic degradation of BB41 ([BB41] = 10 mg L−1, pH = 8).
Figure 13. Influence of photocatalyst dose on photocatalytic degradation of BB41 ([BB41] = 10 mg L−1, pH = 8).
Processes 13 01342 g013
Figure 14. (a) Influence of initial BB41 concentration on decolorization rate (pH = 8, photocatalyst dose = 0.75 g L−1); (b) UV-Vis absorption spectrum of photocatalytic degradation of BB41 by Bi12ZnO20/AgI heterosystem under solar irradiation as function of irradiation time ([BB41] = 10 mg L−1, pH~8, dose = 0.75 g L−1).
Figure 14. (a) Influence of initial BB41 concentration on decolorization rate (pH = 8, photocatalyst dose = 0.75 g L−1); (b) UV-Vis absorption spectrum of photocatalytic degradation of BB41 by Bi12ZnO20/AgI heterosystem under solar irradiation as function of irradiation time ([BB41] = 10 mg L−1, pH~8, dose = 0.75 g L−1).
Processes 13 01342 g014
Figure 15. Linear representation of pseudo-first-order kinetic model (a) and pseudo-second-order kinetic model (b) for photocatalytic degradation of BB41.
Figure 15. Linear representation of pseudo-first-order kinetic model (a) and pseudo-second-order kinetic model (b) for photocatalytic degradation of BB41.
Processes 13 01342 g015
Figure 16. Mineralization analysis of BB41 under solar irradiation ([BB41] = 10 mg L−1, pH~8, photocatalyst dose = 0.75 g L−1).
Figure 16. Mineralization analysis of BB41 under solar irradiation ([BB41] = 10 mg L−1, pH~8, photocatalyst dose = 0.75 g L−1).
Processes 13 01342 g016
Figure 17. Recycling efficiency of Bi12ZnO20/AgI heterosystem in photodegradation of BB41.
Figure 17. Recycling efficiency of Bi12ZnO20/AgI heterosystem in photodegradation of BB41.
Processes 13 01342 g017
Figure 18. Characterization of heterosystem after recycling: XRD diffractogram (a), SEM image (b), EDX spectrum (c).
Figure 18. Characterization of heterosystem after recycling: XRD diffractogram (a), SEM image (b), EDX spectrum (c).
Processes 13 01342 g018
Table 1. Structural parameters of Bi12ZnO20 sillenite.
Table 1. Structural parameters of Bi12ZnO20 sillenite.
Peak Position
(°2th)
FWHM (°2th)Crystal Size (nm)Average Crystal Size (nm)Density
(g cm−3)
24.6760.13361.157539.03
27.6450.13859.431
30.3440.14158.559
32.8420.14855.776
52.2720.19445.729
53.9070.19446.011
55.5090.20344.142
Table 2. Comparison of Raman peaks of Bi12ZnO20 sillenite with those reported in literature.
Table 2. Comparison of Raman peaks of Bi12ZnO20 sillenite with those reported in literature.
Bi12ZnO20 Raman Peaks Observed in This StudyRaman Peaks of Zinc sillenite Bi12.66Zn0.33O19.33 (Ref. [31])Raman Peaks of Sil-Lenite Bi12ZnO20 ([26]) Raman Peaks of BZnO Sillenite ([18])
52-56
80788083
-9192-
126122123127
136136137141
162162162166
-208-209
--181-
--207-
254250251257
307310305311
380380372377
- 446444434
527527527526
630623666622
Table 3. Kinetic parameters of BB41 photodegradation using Bi12ZnO20/AgI heterosystem.
Table 3. Kinetic parameters of BB41 photodegradation using Bi12ZnO20/AgI heterosystem.
C0 (mg L−1)Pseudo-First OrderPseudo-Second Order
R2k1(min−1)R2k2 (L mg−1 min−1)
100.84080.01230.87880.0218
200.96710.00270.94430.0002
300.97330.00320.95320.0002
400.95540.00250.94580.00009
500.96470.00060.87650.00002
Table 4. Pseudo-first-order kinetic parameters of the BB41 photodegradation reaction.
Table 4. Pseudo-first-order kinetic parameters of the BB41 photodegradation reaction.
C0 (mg L−1)R2kapp (min−1)t1/2 (min)r0 (mg L−1 min−1)
100.84080.012356.3530.123
200.96710.0027256.7210.054
300.97330.0032216.6080.096
400.95540.0025277.2590.1
500.96470.00061155.2450.03
Table 5. Comparison of BB41 photodegradation by different photocatalysts.
Table 5. Comparison of BB41 photodegradation by different photocatalysts.
PhotocatalystLight Source[BB41]pHDose Degradation Rate (%)Reference
(mg L−1)(g L−1)
BZO/AgISolar irradiation1080.7598This study
Cu2ZnSnS44 fluorescentlamps (4 W)10free-97.5[41]
SrTiO3/Ag3PO4Irradiation visible20free599[42]
C/ZnO/ZnLumière UV12.5free0.196[10]
TiO2- biochar based on ligninLumière UV406.10.7596.72[43]
(8 W)
TiO2/CaAlgSolar irradiation307196[11]
CoCr2O4Tungsten lamp (200 W)1050.599[35]
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

Madji, S.; Belmedani, M.; Mekatel, E.; Zouaoui, S.; Lebouachera, S.E.I. The Development of a New Bi12ZnO20/AgI Heterosystem for the Degradation of Dye-Contaminated Water by Photocatalysis Under Solar Irradiation: Synthesis, Characterization and Kinetics. Processes 2025, 13, 1342. https://doi.org/10.3390/pr13051342

AMA Style

Madji S, Belmedani M, Mekatel E, Zouaoui S, Lebouachera SEI. The Development of a New Bi12ZnO20/AgI Heterosystem for the Degradation of Dye-Contaminated Water by Photocatalysis Under Solar Irradiation: Synthesis, Characterization and Kinetics. Processes. 2025; 13(5):1342. https://doi.org/10.3390/pr13051342

Chicago/Turabian Style

Madji, Serine, Mohamed Belmedani, Elhadj Mekatel, Sarra Zouaoui, and Seif El Islam Lebouachera. 2025. "The Development of a New Bi12ZnO20/AgI Heterosystem for the Degradation of Dye-Contaminated Water by Photocatalysis Under Solar Irradiation: Synthesis, Characterization and Kinetics" Processes 13, no. 5: 1342. https://doi.org/10.3390/pr13051342

APA Style

Madji, S., Belmedani, M., Mekatel, E., Zouaoui, S., & Lebouachera, S. E. I. (2025). The Development of a New Bi12ZnO20/AgI Heterosystem for the Degradation of Dye-Contaminated Water by Photocatalysis Under Solar Irradiation: Synthesis, Characterization and Kinetics. Processes, 13(5), 1342. https://doi.org/10.3390/pr13051342

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