Next Article in Journal
Numerical Simulation of T Joints Constructed from Hollow Steel Sections
Previous Article in Journal
HertDroid: Android Malware Detection Method with Influential Node Filter and Heterogeneous Graph Transformer
Previous Article in Special Issue
Study of High Transmittance of SiO2/Nb2O5 Multilayer Thin Films Deposited by Plasma-Assisted Reactive Magnetron Sputtering
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis and Characterization of a Photocatalytic Material from TiO2 Nanoparticles Supported on Zeolite Obtained from Ignimbrite Residue Used in Decolorization of Methyl Orange

1
Universidad Nacional de San Agustín de Arequipa UNSA, Av. Independencia s/n, Arequipa 04001, Peru
2
Center for the Development of Advanced Materials and Nanotechnology, Universidad Nacional de Ingeniería, Av. Túpac Amaru 210, Lima 15333, Peru
*
Author to whom correspondence should be addressed.
Appl. Sci. 2024, 14(8), 3146; https://doi.org/10.3390/app14083146
Submission received: 20 December 2023 / Revised: 8 March 2024 / Accepted: 22 March 2024 / Published: 9 April 2024
(This article belongs to the Special Issue Properties, Characterization and Applications of Ceramics Materials)

Abstract

:
In the present work, a TiO2/zeolite photocatalyst was synthesized by dispersing TiO2 nanoparticles obtained through the sol-gel method onto the surface of natural zeolite derived from ignimbrite residue. The zeolite was obtained from an ignimbrite rubble treatment collected from a quarry in Arequipa City, Peru. The research focused on the effect of zeolite on the TiO2 nanoparticles. The synthesized photocatalysts were characterized using various techniques, including field-emission scanning electron microscopy (FE-SEM), energy-dispersive X-ray (EDS), X-ray diffraction (XRD), diffuse reflectance spectroscopy (DRS), and Brunauer–Emmett–Teller surface area analysis (BET). The results revealed that the TiO2/zeolite samples displayed high crystallinity, with TiO2 being present in three phases and zeolite being present in the analcime phase. Furthermore, these samples exhibited a band gap of 3.14 eV and a high surface area compared to that of bare TiO2. Finally, the photocatalytic activity of the TiO2/zeolite composite obtained was evaluated toward the decomposition of 10 ppm and 20 ppm of methyl orange (MO) dye. The TiO2/zeolite samples demonstrated improved photocatalytic activity compared to that of pristine TiO2 under the same experimental conditions. This enhancement is primarily attributed to the increased specific surface area of the TiO2/zeolite samples, making them promising materials for future efficient and sustainable photocatalytic applications.

1. Introduction

Water pollutants, including detergents, industrial chemicals, pesticides, and organic dyes, pose significant environmental threats. These substances are produced by diverse industries [1,2], such as the food, textile, cosmetics, paper, and other sectors [3]. Consequently, efforts have been concentrated on exploring innovative materials and developing efficient wastewater treatment methods to address this environmental concern. Several techniques have been studied for their effectiveness, including adsorption, ion exchange, chemical oxidation, electrolysis, and photocatalytic treatments [4,5].
In recent years, photocatalysis has emerged as a promising and eco-friendly technology. This simple yet potent method has shown a high potential for reducing organic pollution by facilitating an absolute decrease in toxic substances [6]. Several photocatalyst materials have been described, such as ZnO, CuO, TiO2, WO3, Fe2O3, and CuO [7,8,9]; among them, titanium dioxide (TiO2) stands out, especially due to its properties such as low cost, non-toxicity, and high stability in aqueous media and its safe handling [10,11]. However, the high aggregation tendency, low adsorption capacity, and difficult recovery of TiO2 limit its application in water treatment [12,13]; therefore, many researchers have focused on the immobilization of TiO2 on suitable support materials to alleviate the recovery efficiency of TiO2. Among the various support materials, activated carbon [14,15], glass [16], silica [17], graphene [18], and zeolite [19,20] have been used as immobilization substrates to try to improve their separation and recovery in water treatment. Zeolites are promising as a support for TiO2 because they are abundant, readily available, and inexpensive [21]. In addition, they exhibit a high specific surface area, excellent adsorption, and catalysis capacity [21,22]. For these reasons, numerous studies have been conducted to explore the utilization of TiO2/zeolite composites as photocatalytic materials, focusing on optimizing the degradation rate of pollutants in various environmental applications. [23,24].
The synthesis of zeolites involves the use of precursors such as sodium silicate and sodium aluminate together with structuring agents and alkali sources, which can be synthesized from various natural sources that are rich in silica and alumina, known as pozzolans. Examples of these sources include volcanic ash, volcanic tuff, pozzolanic clays, minerals such as diatomite and kaolinite, and rice husk ash. According to the Instituto Geológico Minero y Metalúrgico del Perú, the city of Arequipa is considered a volcanic area where there are different types of volcanic deposits called ignimbrites [25]. The method of extraction and carving of ignimbrite, which is also called “sillar”, generates a large amount of residue, so the raw material is wasted, which contributes to the non-valuation of natural resources. Thus, ignimbrite waste is currently a byproduct that is undervalued and underutilized. Huanca et al. [26] synthesized zeolite from ignimbrite material from the Arequipa area, obtaining a zeolite with high cation exchange capacity and promising applications in the remediation and mitigation of heavy metal contamination in industrial effluents. In addition, a study investigated ignimbrite coated with SiO2-TiO2 for NOx degradation [9].
On the other hand, methyl orange (MO, C14H14N3NaO3S) is an organic chemical compound used as a dye in the textile industry. It can act as a contaminant in water when released. Methyl orange is challenging to biodegrade, and as a result, it may persist in water, potentially causing adverse effects on aquatic ecosystems and water quality. For this reason, removal methods, such as heterogeneous photocatalysis, are sought to effectively decompose it and reduce its environmental impact in water [27].
Accordingly, the present research aims to develop a TiO2/zeolite photocatalytic material using the sol-gel method. Zeolite was synthesized from “sillar” gravel through the hydrothermal method, and it was intended for the zeolite to serve as a support to enhance the photocatalytic discoloration properties of TiO2. The evaluation was conducted using methyl orange as a model of an organic contaminant. To assess the properties of TiO2/zeolite and its precursors, X-ray diffraction (XRD), field-emission scanning electron microscopy (FE-SEM), reflectance (%R) (using a UV–visible spectrophotometer equipped with an integrating sphere), and Brunauer–Emmett–Teller (BET) surface area analyses were performed. Finally, dye decolorization under UV light irradiation was performed by comparing the photocatalytic performance of TiO2/zeolite and pristine TiO2 samples.

2. Materials and Methods

2.1. Materials

All chemicals utilized were of analytical grade and employed without additional purification. Hydrochloric acid (HCl, 37%), titanium (IV) Isopropoxide (C12H28O4Ti, 97%), absolute ethyl alcohol (C2H5OH, 99.5%), and sodium hydroxide (NaOH, ≥99.9999%) were used for the synthesis of TiO2 nanoparticles. Anatase-grade TiO2 with 99.7% purity from Sigma Aldrich (Merck Peruana S.A., Lima, Perú) was acquired for comparison. Ignimbrite remainders were purchased from the Añashuayco quarries from Arequipa. Ultrapure and distilled water were used in all processes.

2.2. Obtaining Zeolite from Ignimbrite Residue

The natural ignimbrite remainders used in this study were obtained from Añashuayco quarry, Arequipa GD (latitude: −16.3592070, length: −71.6085700). The ignimbrite remainders were reduced to smaller parts as required. Then, through a milling process in a planetary ball mill (MP100-Retsch, Retsch, Düsseldorf, Germany) at 440 rpm for 30 min, the powder obtained was sieved to homogenize the granulometry with a 200 µm mesh size.
The synthesis of zeolite was realized through hydrothermal treatment using 120 mL of NaOH-3M with 15 g of milling ignimbrite residue. The mixture was placed in a stainless-steel autoclave and finally placed in a furnace at 180 °C for 24 h.
After this time, it was cooled at room temperature in the autoclave. The resulting solid was washed with distilled water until a pH of 7 was reached, and this was carried out in 10 replicates of 300 mL of distilled water through sonication and centrifugation. Finally, the resulting zeolite was dried in a furnace at 100 °C for 12 h.

2.3. Synthesis of TiO2 Nanoparticles and the TiO2/Zeolite Material

Titanium (IV) isopropoxide served as the titanium source when preparing the TiO2/Zeolite photocatalyst. The titanium precursor solution was prepared by following the method by Hosseini et al. [28]. Initially, 10 mL of titanium (IV) isopropoxide was added dropwise to 30 mL of absolute ethyl alcohol and stirred for 60 min at room temperature. Subsequently, 1 g of synthesized zeolite was added, and the mixture was continuously stirred at 70 °C for 30 min.
After that, a hydrolysis catalyst solution was prepared by stirring 3 mL of HCl in 150 mL of ultrapure water for 10 min. This solution was then gradually added to the mixture of titanium (IV) isopropoxide, ethanol, and zeolite, with continuous stirring at 70 °C for 80 min. The mixture was dried at 100 °C for 16 h to evaporate the solvent. Finally, the resulting dry powders were calcined at 300 °C for 1 h to produce TiO2/zeolite compounds, which primarily exhibited the anatase phase of TiO2 [29].
On the other hand, to obtain pure TiO2 for the tests, the same steps as those of the previously described sol-gel method were followed, excluding the addition of zeolite in the procedure.

2.4. Characterization

Various characterization tests were performed to compare the features among the zeolite, TiO2, and TiO2/zeolite samples. In the case of pure TiO2, two types were used: those that were commercially obtained from Sigma-Aldrich and a sample synthesized according to what was described above. The crystal structure and identification of the phases in the obtained samples were analyzed using an X-ray diffractometer (XRD, Aeris Research, Malvern PANalytical, Malvern, UK) equipped with a CuKα radiation source (λ = 1.5406 Å). Morphological analyses were conducted through scanning electron microscopy (Scios 2, ThermoFisher Scientific, Waltham, MA, USA), and energy-dispersive X-ray spectroscopy (EDS, Thermo Scientific, UltraDry) was utilized for elemental chemical analysis with a working range of 20–25 keV.
To estimate the energy of the band gap of semiconductors, it is crucial to highlight the effectiveness of contemporary methodologies, such as the photoacoustic methodology, in meticulously deciphering the distinct contributions of scattering and absorption. These advanced techniques, which are grounded in the Kubelka–Munk theory, provide a sophisticated perspective on the intricacies of semiconductors, facilitating a more detailed exploration of their energy gap characteristics [30,31,32]. In this study, the optical properties of the samples were tested using a UV-vis DRS (diffuse reflectance spectroscopy) spectrophotometer (Perkin Elmer, Waltham, MA, USA) equipped with an integration sphere in the wavelength range of 200–700 nm. The energy band gaps (Eg) were determined using the Kubelka–Munk theory and Tauc plot analysis, as per Equations (1) and (2), respectively [33].
F R = K S
[ F R   x   h ν ] 1 / p = A ( h ν E g )
where K and S are the absorption and scattering coefficients, is the photon energy, Eg is the optical band gap energy, and A is a proportionality constant independent of the photon energy [34]. To determine the band gap energy (Eg), a linear fit to a plot of [ F R x   h ν ] 1 / p versus must be obtained through extrapolation to zero. In the case of an indirect band gap, such as that of TiO2, a value of p = 2 will be used [33,34].
The textural properties of the samples were assessed through nitrogen isotherms at 77 K using a Micromeritics Gemini VII (Micromeritics, Norcross, GA, USA), where each sample (approximately 0.2 g) underwent a preliminary degassing at 200 °C for 2 h. The specific surface area was determined using the multipoint BET calculation method by following the BET expression in its linear form as outlined by Müller [35].

2.5. Photocatalytic Activity

The photocatalytic activity of the TiO2/zeolite composite was evaluated by decolorizing aqueous solutions containing methyl orange (MO) at concentrations of 10 and 20 ppm. In each test, 0.6 g of the photocatalyst was dispersed in 200 mL of the corresponding MO solution. Subsequently, the suspension was illuminated with a 300 W OSRAM Ultravitalux lamp, where 70 Wm−2 was measured in the UV-A intensity range. The experiments were realized at room temperature. During irradiation, 10 mL of the treated solution was collected every 10 min and centrifuged at 6000 rpm for 4 min. Then, the supernatant was removed, and another centrifugation was performed at 6000 rpm for 2 min, resulting in a clear solution. The remaining concentration of methyl orange was analyzed using a Lambda 365 UV-vis spectrophotometer (PerkinElmer, Waltham, MA, USA) at 507 nm. A comparison of the photocatalytic efficiency of the TiO2/zeolite composite and pure TiO2 was carried out. The methyl orange removal efficiency in percentage degradation (%Ef fdeg) was calculated with Equation (3) [27]:
% E f f d e g = C 0 C t C 0 × 100 %
where C 0 and C t represent the initial concentration of the MO solution and the concentration of MO analyzed after light irradiation during a specific time interval, respectively.

3. Results and Discussion

3.1. XRD Analysis

Figure 1 shows the XRD patterns of the ignimbrite precursor material, zeolite, TiO2/zeolite, and pure TiO2 obtained with the sol-gel method. The ignimbrite pattern reveals that its main identified crystalline phase is anorthoclase, followed by cristobalite and albite. Anorthoclase is a feldspar composed of aluminum silicates with varying percentages of potassium, sodium, and calcium [36,37]. Furthermore, the obtained peaks are primarily located between angles of 15°and 40° (2θ), which is consistent with zeolite synthesis studies by Rajakrishnamoorthy et al. [38] and Verrecchia et al. [39]. The open halo that was observed suggests thermodynamically metastable amorphous aluminosilicate structures with high pozzolanic activity [40].
The zeolite pattern shows that the peaks of ignimbrite decreased, whereas new peaks appeared. This reduction in initial crystalline phases may be attributed to the dissolution of precursors with an alkaline solution, contributing to the increase in silicate and aluminate solutions present in these [41]. This assumption is supported by the study by Ojha et al. [42], where a reduction in the intensity of quartz and mullite was observed after the synthesis process using NaOH. The zeolite spectrum indicated characteristic peaks at 2θ = 15.7°, 18.1°, 25.8°, and 30.4° corresponding to the analcime-type zeolite structure (JCPDS no. 76-0904) [43].
Finally, the XRD patterns of the TiO2 and TiO2/zeolite catalysts evidence characteristic peaks for the anatase phase (JCPDS no. 96-901-5930) that were observed at approximately 25.3°, 37.7°, 48.0° 54.1°, and 62.6° [44]; a rutile phase (JCPDS no. 01-078-1510) was detected in a small proportion [45]. No significant changes were observed in the TiO2 peaks after incorporating zeolite. However, in the TiO2/zeolite compound, two additional peaks corresponding to analcime zeolite were observed. Therefore, the TiO2/zeolite photocatalytic material exhibits the structure of the TiO2 component in the anatase phase along with peaks corresponding to the analytical structure of zeolite.

3.2. SEM Analysis

The samples were analyzed using scanning electron microscopy to examine their surface morphology, interactions, particle size, and elemental chemical composition. The weight percentages of the significant oxides are shown in Table 1, and silicon dioxide (SiO2) was the predominant component in the ignimbrite, zeolite, and TiO2/zeolite samples. The term “others” in Table 1 includes MgO, K2O, CaO, and Fe2O3 oxides.

3.2.1. Ignimbrite

SEM micrographs of the precursor material (ignimbrite), which was previously milled and passed through a 200 µm mesh, are shown in Figure 2a. The figure reveals agglomerated particles with a varied morphology, suggesting a particle size below 100 µm, which was consistent with the expected milling size. The EDS chemical composition analysis is shown in Figure 2c. In this sample, large amounts of silicon, aluminum, and oxygen were revealed, demonstrating the siliceous nature of ignimbrite, which favors zeolite synthesis. The spectrum also shows signals related to CaO and Fe2O3, which could negatively affect the synthesis process [46].

3.2.2. Zeolite

SEM micrographs of the synthesized zeolite are shown in Figure 3a,b. The measured size of the synthesized zeolites ranges from 9 µm to 46 µm, with an average dimension of approximately 45 µm. This average size notably exceeds the measurements reported by earlier researchers [47,48]. Furthermore, the particles exhibit an octahedral three-dimensional crystalline structure, a characteristic of analcime-type zeolite [49]. The EDS spectrum in Figure 3c and the oxide weight percentages listed in Table 1 for the obtained zeolite reveal that SiO2 is the predominant component (54.72%), which is consistent with the starting material, whereas the oxides MgO, K2O, CaO, and Fe2O3 comprise 11.35%.

3.2.3. Synthesized TiO2 Nanoparticles

SEM micrographs of the TiO2 synthesized with the sol-gel method are presented in Figure 4a,b. These images reveal agglomerated TiO2 nanoparticles, which, according to the measurements, exhibit a size range from 18 nm to 45 nm. In addition, it should be noted that in Figure 4c and Table 1, titanium oxide is the predominant component (99.52%).

3.2.4. TiO2/Zeolite Photocatalytic Material

The SEM micrographs in Figure 5a,b exhibit the morphology of the TiO2/zeolite photocatalyst synthesized via the sol-gel method. The images confirm the presence of zeolite and TiO2 nanoparticles, aligning with the EDS analysis and consistent with findings from prior studies [13,50,51]. The diameter of the TiO2 nanoparticles on the zeolite surface was determined to be in the range from 15 nm to 47 nm. Furthermore, the TiO2/zeolite composite demonstrates enhanced material dispersion within its structure compared to that in the pure TiO2 sample. The EDS spectrum (Figure 5c) and Table 1 reveal that SiO2 and TiO2 are the predominant components, constituting the majority, while the remaining 0.90% comprises various other oxides.

3.3. UV–Visible Spectroscopy

The determination of the band gap is a crucial characterization for photocatalytic materials, as it plays a significant role in the photocatalytic reaction—specifically, in light absorption and the generation of electron–hole pairs [52]. According to the Kubelka–Munk model and the extrapolation of the linear adjustment to zero, the corresponding Tauc plot was created for commercial pure TiO2 (Figure 6a), synthesized TiO2 (Figure 6b), and TiO2/zeolite (Figure 6c) with Equations (1) and (2). The results, which are summarized in Table 2, reveal that the calculated band gap for TiO2/zeolite is 3.14 eV, which is smaller than the values obtained for commercial TiO2 (3.29 eV) and synthesized TiO2 (3.18 eV) and is in proximity to the reported values [53,54,55]. These values imply activation only with a UV light source [56]. In addition, the band gap of TiO2/zeolite implies heightened light absorption, which leads to an increased generation of excited charge carriers and, consequently, enhances photocatalytic activity.

3.4. Physisorption Test Parameters

The BET surface area, pore volume, mean pore diameter, and particle size constitute interconnected characteristics that are essential for optimizing the photocatalytic efficiency of a material [57]. In this case study, these parameters were assessed through physisorption testing, and the findings are detailed in Table 3. The results indicated that the TiO2/zeolite material exhibited the highest surface area (169.079 ± 0.802 m2/g), surpassing those of the synthesized TiO2 (86.951 ± 0.346 m2/g), the commercial TiO2 from Sigma Aldrich (80.571 ± 0.278 m2/g), and zeolite analcime (1.740 ± 0.005 m2/g). The zeolite analcime exhibited the lowest surface area value, which was characteristic of this structure type due to its high-temperature synthesis [49]. On the other hand, the value obtained for TiO2/zeolite was approximately double that of the synthesized pure TiO2. The high surface area and porous structure obtained for this material favor the adsorption, desorption, and diffusion of reagents, resulting in high photocatalytic activity [58]. Furthermore, the other measured parameters, such as the pore volume and pore size, in the TiO2/zeolite composite also exhibit the highest values compared to those of the other samples, as indicated in Table 3.

3.5. Photocatalytic Performance Results

To evaluate the photocatalytic activity, the photodegradation of methyl orange dye was measured at concentrations of 10 and 20 ppm in aqueous solutions under UV light. The photocatalytic activity results for 10 and 20 ppm are shown in Figure 7, with Ct/Co representing the ratio of the concentration at time “t” to the initial concentration of methyl orange over time in minutes. The highest photocatalytic efficiency during the first 60 min under UV light irradiation was observed for the TiO2/zeolite sample compared to pure TiO2 in the anatase phase for both concentrations. The percentage of degradation efficiency was calculated according to Equation (3), and the results are shown in Table 4. For a concentration of 10 ppm, the TiO2/zeolite synthesized from ignimbrite residue as a source of zeolite degraded 94.71% of the dye in the first 60 min, while the synthesized TiO2 degraded 83.93%, which was 10.78% lower. For the case of 20 ppm, the TiO2/zeolite photocatalyst (71.36%) also demonstrated higher efficiency than that of pure TiO2 (56.78%). The pH remained constant at 3.5 for all cases.
The photodegradation of MO with the various photocatalysts was fitted to the pseudo-first-order kinetic model [59]:
ln C o C t = k t
where Ct, C0, and k represent the residual MO concentration, initial MO concentration, and pseudo-first-order rate constant, respectively. Figure 8 displays a graph representing the relationship between ln(C0/Ct) and time for various concentrations of MO when utili-zing TiO2/zeolite and pristine TiO2 as photocatalysts. A robust fit for the first-order model was achieved, with a correlation coefficient (R2) exceeding 0.90 in all cases, as depicted in Table 5. Additionally, the rate constant was determined from these graphs, and the results are presented in Table 5. As shown in the figure and table, TiO2/zeolite exhibited the highest rate constants for the photodegradation of MO both for the concentration of 10 ppm (0.0436 min−1) and for that of 20 ppm (0.0207 min−1), and they were higher than those obtained for pure TiO2 (0.0286 min−1 and 0.0139 min−1 for 10 and 20 ppm, respectively). The enhanced efficiency of the TiO2/zeolite photocatalytic material was attributed to the presence of zeolite, which facilitated the adsorption of the dye and the interaction with TiO2 to break down these organic dyes, which agreed with other studies that were carried out [13,15,21]. The synergy of TiO2 nanoparticles with zeolite boosts the photocatalytic activity by leveraging zeolites’ high surface area and many active sites for enhanced adsorption of methyl orange [13]. This combination prevents nanoparticle aggregation and amplifies light adsorption. Furthermore, zeolites act as electron sinks, diminishing electron–hole recombination and prolonging electron lifetime [13,21]. Consequently, this enhances the overall photocatalytic activity of the material.

3.5.1. Effect of Catalyst Loading

In photocatalysis, the degradation efficiency is affected by the photocatalyst dose. To examine the effect of the catalyst dose on the photodegradation of MO for 30 min, the amount of TiO2/zeolite was varied from 0.2 to 1 g at a constant MO concentration of 10 ppm. As shown in Figure 9, as the amount of photocatalyst increased from 0.2 to 1 g, the MO degradation efficiency increased (0%, 45%, 58%, 88%, and 95% for 0.2, 0.4, 0.6, 0.8, and 1 g, respectively). This occurs because a higher catalyst concentration offers additional active sites for redox reactions. In other words, irradiation produces more charge carriers to participate in redox reactions [60].

3.5.2. The Reusability Study

The study of photocatalytic applications places significant importance on reusability. This aspect is crucial, as it enables us to determine whether a photocatalyst can be effectively reused in wastewater treatment. Given this, to evaluate the reusability of the TiO2/zeolite composite, three consecutive tests of degradation of MO were conducted under the same reaction conditions. As shown in Table 6, the TiO2/zeolite photocatalysts retain their ability to degrade MO solutions at 10 ppm even after three uses, showing a slight difference from the initial experiment. However, for 20 ppm solutions of MO, the sample presents a considerable decrease in its degradation efficiency after two cycles. The loss might be attributed to the catalyst leaching that inevitably occurs during the recovery and washing process, as well as the blocking of the active site by the surface coverage with the dye and its degradation intermediates [61].

3.5.3. Comparative Analysis

The TiO2/zeolite photocatalyst was compared with previously reported photocatalysts for the photodegradation of organic dyes, and the results are noted in Table 7. It should be noted that while the presented photocatalyst is comparable and even improved with respect to those reported in the literature, direct comparisons are challenging due to differences in dye type and its initial concentration, light source, catalyst morphology, irradiation time, and source of zeolite. In this research, the source of zeolite was ignimbrite, which is a good source of zeolites used in the synthesis of photocatalytic materials. Therefore, these considerations suggest that the TiO2/zeolite presented here is a novel and suitable material for the photodegradation of methyl orange.

4. Conclusions

In summary, a TiO2/zeolite photocatalytic material was synthesized through the sol-gel method to facilitate the degradation of organic dyes. The zeolite, which was sourced from ignimbrite residues from the Añashuayco quarry in Arequipa, was obtained in its analcime type, while the TiO2 was obtained in its anatase form, which is preferred for its photocatalytic activity. Furthermore, the synthesized pure TiO2 showed a band gap and physisorption parameters like those of commercial TiO2. Meanwhile, the TiO2/zeolite photocatalytic material exhibited superior physisorption properties and dye degradation, achieving a photocatalytic efficiency of 94.71% and 83.93% for solutions of 10 and 20 ppm, respectively, in 60 min. This occurred due to its high specific surface area (169.0788 m2/g), larger pore volume (0.04 cm3/g), and band gap of 3.14 eV. On the other hand, the TiO2/zeolite photocatalytic material can be reused in photocatalytic applications with slight changes in its performance. However, at higher concentrations, this reusability decreases. Zeolite improved the photodegradation performance of the TiO2/zeolite material due to its influence on the surface morphology and interaction, favoring the uniform dispersion of TiO2 nanoparticles and providing more active sites for the adsorption of contaminants. Thus, this TiO2/zeolite photocatalytic material based on zeolite obtained from ignimbrite residue effectively removes organic dyes such as methyl orange from water, demonstrating high efficiency and potential applications in wastewater treatment.

Author Contributions

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

Funding

This research was financed by the Universidad Nacional de San Agustín de Arequipa-Perú with the project: “Valorización del Ripio de Sillar de las Canteras de Arequipa usadas como carga para la obtención de recubrimientos fotocatalíticos TiO2/Ignimbrita con propiedades Auto-Limpiantes y Descontaminante” (contract number: IBAIB-05-2018-UNSA) and through the collaboration of the Center for the Development of Advanced Materials and Nanotechnology of the Universidad Nacional de Ingeniería, Peru. P.G.R. thanks the Council for Science and Technology (CONCYTEC) and PROCIENCIA program for the postdoctoral fellowship (contract number: 059-2021-PROCIENCIA).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author. The data are not publicly available due to privacy.

Conflicts of Interest

The authors declare no conflicts of interest. The funders had no role in the design of the study, in the collection, analyses, or interpretation of data, in the writing of the manuscript, or in the decision to publish the results.

References

  1. Crini, G.; Lichtfouse, E. Wastewater Treatment: An Overview. In Green Adsorbents for Pollutant Removal; Springer: Berlin, Germany, 2018; pp. 1–21. [Google Scholar] [CrossRef]
  2. Abdellah, M.H.; Nosier, S.A.; El-Shazly, A.H.; Mubarak, A.A. Photocatalytic Decolorization of Methylene Blue Using TiO2/UV System Enhanced by Air Sparging. Alex. Eng. J. 2018, 57, 3727–3735. [Google Scholar] [CrossRef]
  3. Khataee, A.R.; Pons, M.N.; Zahraa, O. Photocatalytic Degradation of Three Azo Dyes Using Immobilized TiO2 Nanoparticles on Glass Plates Activated by UV Light Irradiation: Influence of Dye Molecular Structure. J. Hazard. Mater. 2009, 168, 451–457. [Google Scholar] [CrossRef] [PubMed]
  4. Peng, H.; Guo, J. Removal of chromium from wastewater by membrane filtration, chemical precipitation, ion exchange, adsorption electrocoagulation, electrochemical reduction, electrodialysis, electrodeionization, photocatalysis and nanotechnology: A review. Environ. Chem. Lett. 2020, 18, 2055–2068. [Google Scholar] [CrossRef]
  5. Ahmad, A.; Mohd-Setapar, S.H.; Chuong, C.S.; Khatoon, A.; Wani, W.A.; Kumar, R.; Rafatullah, M. Recent advances in new generation dye removal technologies: Novel search for approaches to reprocess wastewater. RSC Adv. 2015, 5, 30801–30818. [Google Scholar] [CrossRef]
  6. Ameta, R.; Solanki, M.S.; Benjamin, S.; Ameta, S.C. Photocatalysis. In Advanced Oxidation Processes for Wastewater Treatment: Emerging Green Chemical Technology; Elsevier Inc.: Amsterdam, The Netherlands, 2018; pp. 135–175. [Google Scholar] [CrossRef]
  7. Ramos, P.G.; Espinoza, J.; Sánchez, L.A.; Rodriguez, J.M. Enhanced photocatalytic degradation of Rhodamine B employing transition metal (Fe, Cu, Co) doped ZnO/rGO nanostructures synthesized by electrospinning-hydrothermal technique. J. Alloys Compd. 2023, 966, 171559. [Google Scholar] [CrossRef]
  8. Ângelo, J.; Magalhães, P.; Andrade, L.; Mendes, A. Characterization of TiO2-based semiconductors for photocatalysis by electrochemical impedance spectroscopy. Appl. Surf. Sci. 2016, 387, 183–189. [Google Scholar] [CrossRef]
  9. Florez, D.M.M.G.; Ale, R.B.G.; Idme, A.F.H.; Alarcon, L.A.L.; Huallpa, E.A.; Castro, Y.; Apestegui, P.G.R.; Rodriguez, J.M.R. SiO2-TiO2 Films Supported on Ignimbrite by Spray Coating for the Photocatalytic Degradation of NOx Gas and Methyl Orange Dye. Int. J. Photoenergy 2020, 2020, 4756952. [Google Scholar] [CrossRef]
  10. Nakata, K.; Fujishima, A. TiO2 Photocatalysis: Design and Applications. J. Photochem. Photobiol. C Photochem. Rev. 2012, 13, 169–189. [Google Scholar] [CrossRef]
  11. Guo, Q.; Zhou, C.; Ma, Z.; Yang, X. Fundamentals of TiO2 Photocatalysis: Concepts, Mechanisms, and Challenges. Adv. Mater. 2019, 31, 1901997. [Google Scholar] [CrossRef]
  12. Sarkar, A.K.; Saha, A.; Tarafder, A.; Panda, A.B.; Pal, S. Efficient Removal of Toxic Dyes via Simultaneous Adsorption and Solar Light Driven Photodegradation Using Recyclable Functionalized Amylopectin-TiO2-Au Nanocomposite. ACS Sustain. Chem. Eng. 2016, 4, 1679–1688. [Google Scholar] [CrossRef]
  13. Zhang, G.; Song, A.; Duan, Y.; Zheng, S. Enhanced Photocatalytic Activity of TiO2/Zeolite Composite for Abatement of Pollutants. Microporous Mesoporous Mater. 2018, 255, 61–68. [Google Scholar] [CrossRef]
  14. Fu, X.; Yang, H.; Lu, G.; Tu, Y.; Wu, J. Improved performance of surface functionalized TiO2/activated carbon for adsorption–photocatalytic reduction of Cr(VI) in aqueous solution. Mater. Sci. Semicond. Process 2015, 39, 362–370. [Google Scholar] [CrossRef]
  15. Martins, A.C.; Cazetta, A.L.; Pezoti, O.; Souza, J.R.B.; Zhang, T.; Pilau, E.J.; Asefa, T.; Almeida, V.C. Sol-gel synthesis of new TiO2/activated carbon photocatalyst and its application for degradation of tetracycline. Ceram. Int. 2017, 43, 4411–4418. [Google Scholar] [CrossRef]
  16. Bouarioua, A.; Zerdaoui, M. Photocatalytic activities of TiO2 layers immobilized on glass substrates by dip-coating technique toward the decolorization of methyl orange as a model organic pollutant. J. Environ. Chem. Eng. 2017, 5, 1565–1574. [Google Scholar] [CrossRef]
  17. Kuo, H.P.; Yao, S.W.; Huang, A.N.; Hsu, W.Y. Photocatalytic degradation of toluene in a staged fluidized bed reactor using TiO2/silica gel. Korean J. Chem. Eng. 2017, 34, 73–80. [Google Scholar] [CrossRef]
  18. Kusiak-Nejman, E.; Morawski, A.W. TiO2/graphene-based nanocomposites for water treatment: A brief overview of charge carrier transfer, antimicrobial and photocatalytic performance. Appl. Catal. B 2019, 253, 179–186. [Google Scholar] [CrossRef]
  19. Liu, X.; Liu, Y.; Lu, S.; Guo, W.; Xi, B. Performance and mechanism into TiO2/Zeolite composites for sulfadiazine adsorption and photodegradation. Chem. Eng. J. 2018, 350, 131–147. [Google Scholar] [CrossRef]
  20. Badvi, K.; Javanbakht, V. Enhanced photocatalytic degradation of dye contaminants with TiO2 immobilized on ZSM-5 zeolite modified with nickel nanoparticles. J. Clean. Prod. 2021, 280, 124518. [Google Scholar] [CrossRef]
  21. Setthaya, N.; Chindaprasirt, P.; Yin, S.; Pimraksa, K. TiO2-zeolite photocatalysts made of metakaolin and rice husk ash for removal of methylene blue dye. Powder Technol. 2017, 313, 417–426. [Google Scholar] [CrossRef]
  22. Chong, M.N.; Tneu, Z.Y.; Poh, P.E.; Jin, B.; Aryal, R. Synthesis, characterisation and application of TiO2–zeolite nanocomposites for the advanced treatment of industrial dye wastewater. J. Taiwan. Inst. Chem. Eng. 2015, 50, 288–296. [Google Scholar] [CrossRef]
  23. Aziztyana, A.P.; Wardhani, S.; Prananto, Y.P.; Purwonugroho, D.; Darjito. Optimisation of Methyl Orange Photodegradation Using TiO2-Zeolite Photocatalyst and H2O2 in Acid Condition. IOP Conf. Ser. Mater. Sci. Eng. 2019, 546, 042047. [Google Scholar] [CrossRef]
  24. Rahman, A.; Nurjayadi, M.; Wartilah, R.; Kusrini, E.; Prasetyanto, E.A.; Degermenci, V. Enhanced Activity of TiO2/Natural Zeolite Composite for Degradation of Methyl Orange under Visible Light Irradiation. Int. J. Technol. 2018, 9, 1159–1167. [Google Scholar] [CrossRef]
  25. Lebti, P.P.; Thouret, J.C.; Wörner, G.; Fornari, M. Neogene and Quaternary ignimbrites in the area of Arequipa, Southern Peru: Stratigraphical and petrological correlations. J. Volcanol. Geotherm. Res. 2006, 154, 251–275. [Google Scholar] [CrossRef]
  26. Huanca, P.K.; Corrales, L.T.; Paredes, B.; Almirón, J.J.; Gonzales, D.P. Estudio de la modificación hidrotermal de un mineral Ignimbrítico para obtener zeolita sintética de alta capacidad de intercambio catiónico. Rev. Bol. Quim 2019, 36, 173–180. [Google Scholar] [CrossRef]
  27. Waghchaure, R.H.; Adole, V.A.; Jagdale, B.S. Photocatalytic degradation of methylene blue, rhodamine B, methyl orange and Eriochrome black T dyes by modified ZnO nanocatalysts: A concise review. Inorg. Chem. Commun. 2022, 143, 109764. [Google Scholar] [CrossRef]
  28. Hosseini, M.S.; Sadeghi, M.T.; Khazaei, M. Wettability alteration from superhydrophobic to superhydrophilic via synthesized stable nano-coating. Surf. Coat. Technol. 2017, 326, 79–86. [Google Scholar] [CrossRef]
  29. Mosquera, E.; Rosas, N.; Debut, A.; Guerrero, V.H. Síntesis y caracterización de nanopartículas de dióxido de titanio obtenidas por el método de sol-gel. Rev. Politécnica 2015, 36, 7. [Google Scholar]
  30. Lamastra, F.R.; Grilli, M.L.; Leahu, G.; Belardini, A.; Li Voti, R.; Sibilia, C.; Salvatori, D.; Cacciotti, I.; Nanni, F. Photoacoustic Spectroscopy Investigation of Zinc Oxide/Diatom Frustules Hybrid Powders. Int. J. Thermophys. 2018, 39, 110. [Google Scholar] [CrossRef]
  31. Lamastra, F.R.; Grilli, M.L.; Leahu, G.; Belardini, A.; Li Voti, R.; Sibilia, C.; Salvatori, D.; Cacciotti, I.; Nanni, F. Diatom frustules decorated with zinc oxide nanoparticles for enhanced optical properties. Nanotechnology 2017, 28, 375704. [Google Scholar] [CrossRef]
  32. Li Voti, R.; Leahu, G.; Sibilia, C.; Matassa, R.; Familiari, G.; Cerra, S.; Salamone, T.A.; Fratoddi, I. Photoacoustics for listening to metal nanoparticle super-aggregates. Nanoscale Adv. 2021, 3, 4692–4701. [Google Scholar] [CrossRef]
  33. Landi, S.; Segundo, I.R.; Freitas, E.; Vasilevskiy, M.; Carneiro, J.; Tavares, C.J. Use and misuse of the Kubelka-Munk function to obtain the band gap energy from diffuse reflectance measurements. Solid. State Commun. 2022, 341, 114573. [Google Scholar] [CrossRef]
  34. López, R.; Gómez, R. Band-gap energy estimation from diffuse reflectance measurements on sol–gel and commercial TiO2: A comparative study. J. Sol-Gel Sci. Technol. 2012, 61, 1–7. [Google Scholar] [CrossRef]
  35. Müller, B.R. Preparation and characterization of K2CO3-doped powdered activated carbon for effective in-vitro adsorption of deoxynivalenol. Bioresour. Technol. 2021, 15, 100703. [Google Scholar] [CrossRef]
  36. Klima, K.M.; Schollbach, K.; Brouwers, H.J.H.; Yu, Q. Enhancing the thermal performance of Class F fly ash-based geopolymer by sodalite. Constr. Build. Mater. 2022, 314, 125574. [Google Scholar] [CrossRef]
  37. Cruz Sánchez, E.; Torres, M.E.; Diaz, C.; Saito, F. Effects of grinding of the feldspar in the sintering using a planetary ball mill. J. Mater. Process. Technol. 2004, 152, 284–290. [Google Scholar] [CrossRef]
  38. Rajakrishnamoorthy, P.; Karthikeyan, D.; Saravanan, C.G. Emission reduction technique applied in SI engines exhaust by using zsm5 zeolite as catalysts synthesized from coal fly ash. Mater. Today Proc. 2020, 22, 499–506. [Google Scholar] [CrossRef]
  39. Verrecchia, G.; Cafiero, L.; De Caprariis, B.; Dell’Era, A.; Pettiti, I.; Tuffi, R.; Scarsella, M. Study of the parameters of zeolites synthesis from coal fly ash in order to optimize their CO2 adsorption. Fuel 2020, 276, 118041. [Google Scholar] [CrossRef]
  40. Churata, R.; Almirón, J.; Vargas, M.; Tupayachy-Quispe, D.; Torres-Almirón, J.; Ortiz-Valdivia, Y.; Velasco, F. Study of Geopolymer Composites Based on Volcanic Ash, Fly Ash, Pozzolan, Metakaolin and Mining Tailing. Buildings 2022, 12, 1118. [Google Scholar] [CrossRef]
  41. Vichaphund, S.; Aht-Ong, D.; Sricharoenchaikul, V.; Atong, D. Characteristic of fly ash derived-zeolite and its catalytic performance for fast pyrolysis of Jatropha waste. Environ. Technol. 2014, 35, 2254–2261. [Google Scholar] [CrossRef]
  42. Ojha, K.; Pradhan, N.C.; Samanta, A.N. Zeolite from fly ash: Synthesis and characterization. Bull. Mater. Sci. 2004, 27, 555–564. [Google Scholar] [CrossRef]
  43. Galindo Valbuena, H.M.; Medina, A.F.; Vargas, J.C.; Hernández Fandiño, O. Synthesis of zeolites Na-A, Na-X, and analcime from crushed stone waste and their applications in heavy metal removal in aqueous media. Chem. Eng. Res. Des. 2023, 197, 159–172. [Google Scholar] [CrossRef]
  44. Jung, Y.J.; Kim, W.G.; Yoon, Y.; Kang, J.W.; Hong, Y.M.; Kim, H.W. Removal of amoxicillin by UV and UV/H2O2 processes. Sci. Total Environ. 2012, 420, 160–167. [Google Scholar] [CrossRef] [PubMed]
  45. Znad, H.; Abbas, K.; Hena, S.; Awual, M.R. Synthesis a novel multilamellar mesoporous TiO2/ZSM-5 for photo-catalytic degradation of methyl orange dye in aqueous media. J. Environ. Chem. Eng. 2018, 6, 218–227. [Google Scholar] [CrossRef]
  46. Rodríguez Valdivia, M. Hydrothermal synthesis of zeolites from a volcanic ash by alkaline treatment and their potential application in the removal of NH4+, Pb2+, Zn2+ and Mn2+. Matéria 2022, 27, e13132. [Google Scholar] [CrossRef]
  47. Novembre, D.; Gimeno, D. Synthesis and characterization of analcime (ANA) zeolite using a kaolinitic rock. Sci. Rep. 2021, 11, 13373. [Google Scholar] [CrossRef] [PubMed]
  48. Vigil de la Villa Mencía, R.; Goiti, E.; Ocejo, M.; Giménez, R.G. Synthesis of zeolite type analcime from industrial wastes. Microporous Mesoporous Mater. 2020, 293, 109817. [Google Scholar] [CrossRef]
  49. Setthaya, N.; Pindi, C.; Chindaprasirt, P.; Pimraksa, K. Synthesis of faujasite and analcime using of rice husk ashand metakaolin. Adv. Mater. Res. 2013, 770, 209–212. [Google Scholar] [CrossRef]
  50. Jin, Y.; Qi, T.; Ge, Y.; Chen, J.; Liang, L.; Ju, J.; Zhao, J. Ultrasensitive electrochemical determination of phosphate in water by using hydrophilic TiO2 modified glassy carbon electrodes. Anal. Methods 2021, 13, 996–1002. [Google Scholar] [CrossRef] [PubMed]
  51. Jansson, I.; Suárez, S.; Garcia-Garcia, F.J.; Sánchez, B. Zeolite–TiO2 hybrid composites for pollutant degradation in gas phase. Appl. Catal. B 2015, 178, 100–107. [Google Scholar] [CrossRef]
  52. Katal, R.; Masudy-Panah, S.; Tanhaei, M.; Farahani, M.H.D.A.; Jiangyong, H. A review on the synthesis of the various types of anatase TiO2 facets and their applications for photocatalysis. Chem. Eng. J. 2020, 384, 123384. [Google Scholar] [CrossRef]
  53. Tayade, R.J.; Surolia, P.K.; Kulkarni, R.G.; Jasra, R.V. Photocatalytic degradation of dyes and organic contaminants in water using nanocrystalline anatase and rutile TiO2. Sci. Technol. Adv. Mater. 2007, 8, 455–462. [Google Scholar] [CrossRef]
  54. Dette, C.; Pérez-Osorio, M.A.; Kley, C.S.; Punke, P.; Patrick, C.E.; Jacobson, P.; Giustino, F.; Jung, S.J.; Kern, K. TiO2 Anatase with a Bandgap in the Visible Region. Nano Lett. 2014, 14, 6533–6538. [Google Scholar] [CrossRef] [PubMed]
  55. Shi, D.; Guo, Z.; Bedford, N. Nanotitanium Oxide as a Photocatalytic Material and its Application. In Nanomaterials and Devices; Elsevier: Amsterdam, The Netherlands, 2015; pp. 161–174. [Google Scholar] [CrossRef]
  56. Dincer, I.; Zamfirescu, C. Hydrogen Production by Photonic Energy. In Sustainable Hydrogen Production; Elsevier: New York, NY, USA, 2016; pp. 309–391. [Google Scholar] [CrossRef]
  57. Matsukawa, Y.; Hirata, S.; Inada, M.; Enomoto, N.; Hojo, J.; Hayashi, K. Kinetic effects of polymorphs and surface areas on adsorption and photocatalytic decomposition of acetaldehyde on titania. Chem. Eng. J. 2020, 397, 125422. [Google Scholar] [CrossRef]
  58. Guo, B.; Shen, H.; Shu, K.; Zeng, Y.; Ning, W. The study of the relationship between pore structure and photocatalysis of mesoporous TiO2. J. Chem. Sci. 2009, 121, 317–321. [Google Scholar] [CrossRef]
  59. Labhane, P.K.; Patle, L.B.; Huse, V.R.; Sonawane, G.H.; Sonawane, S.H. Synthesis of reduced graphene oxide sheets decorated by zinc oxide nanoparticles: Crystallographic, optical, morphological and photocatalytic study. Chem. Phys. Lett. 2016, 661, 13–19. [Google Scholar] [CrossRef]
  60. Sulaiman, F.; Sari, D.K.; Kustiningsih, I. The influence of ozone on the photocatalytic degradation of phenol using TiO2 photocatalyst supported by Bayah natural zeolite. AIP Conf. Proc. 2017, 1840, 110014. [Google Scholar] [CrossRef]
  61. Hunge, Y.M.; Yadav, A.A.; Kang, S.W.; Kim, H. Photocatalytic degradation of tetracycline antibiotics using hydrothermally synthesized two-dimensional molybdenum disulfide/titanium dioxide composites. J. Colloid Interface Sci. 2022, 606, 454–463. [Google Scholar] [CrossRef]
Figure 1. X-ray diffraction patterns for TiO2, TiO2/zeolite, zeolite, and ignimbrite samples.
Figure 1. X-ray diffraction patterns for TiO2, TiO2/zeolite, zeolite, and ignimbrite samples.
Applsci 14 03146 g001
Figure 2. SEM micrograph of (a) milling ignimbrite and (b) the analyzed area of ignimbrite. (c) EDS of the area analyzed in (b).
Figure 2. SEM micrograph of (a) milling ignimbrite and (b) the analyzed area of ignimbrite. (c) EDS of the area analyzed in (b).
Applsci 14 03146 g002
Figure 3. SEM micrographs of the synthesized zeolites at (a) low and (b) high magnification. (c) EDS of the whole area of the synthesized zeolite.
Figure 3. SEM micrographs of the synthesized zeolites at (a) low and (b) high magnification. (c) EDS of the whole area of the synthesized zeolite.
Applsci 14 03146 g003
Figure 4. SEM micrograph of (a) the synthesized TiO2 nanoparticles and (b) the analyzed nanoparticle area. (c) EDS of the entire area analyzed in (b).
Figure 4. SEM micrograph of (a) the synthesized TiO2 nanoparticles and (b) the analyzed nanoparticle area. (c) EDS of the entire area analyzed in (b).
Applsci 14 03146 g004
Figure 5. SEM micrograph of (a) the TiO2/zeolite photocatalytic material and (b) the analyzed nanoparticle area. (c) EDS of the entire area analyzed in (b).
Figure 5. SEM micrograph of (a) the TiO2/zeolite photocatalytic material and (b) the analyzed nanoparticle area. (c) EDS of the entire area analyzed in (b).
Applsci 14 03146 g005
Figure 6. Indirect band gaps calculated for (a) TiO2 from Sigma Aldrich, (b) synthesized TiO2, and (c) TiO2/zeolite. The red line represents the Tauc plot image, and the tangent blue line indicates the estimated band gap energy from the linear fit.
Figure 6. Indirect band gaps calculated for (a) TiO2 from Sigma Aldrich, (b) synthesized TiO2, and (c) TiO2/zeolite. The red line represents the Tauc plot image, and the tangent blue line indicates the estimated band gap energy from the linear fit.
Applsci 14 03146 g006
Figure 7. Photodegradation of methyl orange using pure TiO2 and TiO2/zeolite material at concentrations of (a) 10 ppm and (b) 20 ppm.
Figure 7. Photodegradation of methyl orange using pure TiO2 and TiO2/zeolite material at concentrations of (a) 10 ppm and (b) 20 ppm.
Applsci 14 03146 g007
Figure 8. ln(C0/Ct) as a function of time for the photodegradation of methyl orange at concentrations of (a) 10 ppm and (b) 20 ppm for TiO2/zeolite and pure TiO2.
Figure 8. ln(C0/Ct) as a function of time for the photodegradation of methyl orange at concentrations of (a) 10 ppm and (b) 20 ppm for TiO2/zeolite and pure TiO2.
Applsci 14 03146 g008
Figure 9. Photodegradation efficiency of MO using various amounts of TiO2/zeolite for 30 min.
Figure 9. Photodegradation efficiency of MO using various amounts of TiO2/zeolite for 30 min.
Applsci 14 03146 g009
Table 1. Weight percentages of oxides in the samples.
Table 1. Weight percentages of oxides in the samples.
Weight Percentage %
OxidesIgnimbriteZeoliteTiO2TiO2/Zeolite
SiO274.9754.72----73.55
Al2O313.5219.60----5.77
Na2O4.3812.98----1.06
TiO20.180.5299.5217.54
Cl0.370.830.481.18
Others6.5811.35----0.90
Total100100100100
Table 2. The calculated band gaps of the samples.
Table 2. The calculated band gaps of the samples.
SampleBand Gap (eV)
TiO2/zeolite3.14
TiO2—Sigma Aldrich3.29
Synthesized TiO2 3.18
Table 3. The textural properties of all samples.
Table 3. The textural properties of all samples.
SampleSurface Area
[m2/g]
Pore Volume
[cm3/g]
Average Pore
Diameter
[nm]
Zeolite analcime1.740 ± 0.0050.0004892.1426
TiO2—Sigma Aldrich80.571 ± 0.2780.0223792.1420
Synthesized TiO2 86.951 ± 0.3460.0259582.1505
TiO2/zeolite169.079 ± 0.8020.0420612.1411
Table 4. Comparison of the degradation efficiency for the samples.
Table 4. Comparison of the degradation efficiency for the samples.
SampleConcentration
[ppm]
%Ef fdeg
TiO2/zeolite1094.71
Pure synthesized TiO2 1083.93
TiO2/zeolite2071.36
Pure synthesized TiO2 2056.78
Table 5. Kinetic parameters in the photocatalytic experiments.
Table 5. Kinetic parameters in the photocatalytic experiments.
SampleConcentration
[ppm]
k (min−1)R2
TiO2/zeolite100.04360.963
Pure synthesized TiO2 100.02860.917
TiO2/zeolite200.02070.994
Pure synthesized TiO2 200.01390.994
Table 6. Reutilization of TiO2/zeolite for the photodegradation of methyl orange.
Table 6. Reutilization of TiO2/zeolite for the photodegradation of methyl orange.
SampleDegradation Efficiency (%)Concentration
TiO2/zeolite (First Test)92.8010 ppm
TiO2/zeolite (Second Test)88.15
TiO2/zeolite (Third Test)87.64
TiO2/zeolite (First Test)70.2020 ppm
TiO2/zeolite (Second Test)67.69
TiO2/zeolite (Third Test)18.06
Table 7. Comparison of the photocatalytic performance for different TiO2/zeolite materials.
Table 7. Comparison of the photocatalytic performance for different TiO2/zeolite materials.
SampleDegradation/Time TestConcentration
TiO2/zeolite (this work)94.71%/60 min10 ppm
TiO2/zeolite [23]77.68%/60 min10 ppm
TiO2/SDS [50]90%/300 min30 ppm
TiO2 layers on glass substrates—3 layers [16]97.6%/180 min30 ppm
TiO2/zeolite (this work)71.36%/60 min20 ppm
TiO2/zeolite [23]68.61%/60 min20 ppm
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

Huayna, G.; Laura, A.; Churata, R.; Lazo, L.; Guzmán, R.; Ramos, P.G.; Rodriguez, J.M. Synthesis and Characterization of a Photocatalytic Material from TiO2 Nanoparticles Supported on Zeolite Obtained from Ignimbrite Residue Used in Decolorization of Methyl Orange. Appl. Sci. 2024, 14, 3146. https://doi.org/10.3390/app14083146

AMA Style

Huayna G, Laura A, Churata R, Lazo L, Guzmán R, Ramos PG, Rodriguez JM. Synthesis and Characterization of a Photocatalytic Material from TiO2 Nanoparticles Supported on Zeolite Obtained from Ignimbrite Residue Used in Decolorization of Methyl Orange. Applied Sciences. 2024; 14(8):3146. https://doi.org/10.3390/app14083146

Chicago/Turabian Style

Huayna, Gianina, Antonio Laura, Rossibel Churata, Luis Lazo, Rivalino Guzmán, Pierre G. Ramos, and Juan M. Rodriguez. 2024. "Synthesis and Characterization of a Photocatalytic Material from TiO2 Nanoparticles Supported on Zeolite Obtained from Ignimbrite Residue Used in Decolorization of Methyl Orange" Applied Sciences 14, no. 8: 3146. https://doi.org/10.3390/app14083146

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