Next Article in Journal
Susceptibility Assessment of Landslides in the Loess Plateau Based on Machine Learning Models: A Case Study of Xining City
Next Article in Special Issue
The Biosynthesis of Nickel Oxide Nanoparticles: An Eco-Friendly Approach for Azo Dye Decolorization and Industrial Wastewater Treatment
Previous Article in Journal
Is the Urban Landscape Connected? Construction and Optimization of Urban Ecological Networks Based on Morphological Spatial Pattern Analysis
Previous Article in Special Issue
Homogeneous Photosensitized Oxidation for Water Reuse in Cellars: A Study of Different Photosensitizers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Development of a Novel 3D Highly Porous Structure for TiO2 Immobilization and Application in As(III) Oxidation

by
Julio A. Scherer Filho
1,
Belisa A. Marinho
1,2,*,
Fabiola Vignola
1,
Luciana P. Mazur
1,
Sergio Y. G. González
1,
Adriano da Silva
1,
Antônio Augusto Ulson de Souza
1 and
Selene M. A. Guelli Ulson de Souza
1
1
Laboratory of Mass Transfer (LABMASSA), Department of Chemical Engineering (EQA), Federal University of Santa Catarina (UFSC), Florianópolis 88040-900, Brazil
2
Department for Nanostructured Materials, Jožef Stefan Institute, Jamova 39, 1000 Ljubljana, Slovenia
*
Author to whom correspondence should be addressed.
Sustainability 2023, 15(20), 14760; https://doi.org/10.3390/su152014760
Submission received: 10 September 2023 / Revised: 6 October 2023 / Accepted: 9 October 2023 / Published: 11 October 2023

Abstract

:
One of the main drawbacks of the application of photocatalysis for wastewater treatment is the use of dispersed photocatalysts, which are difficult to remove from effluent after the treatment process and may pose additional toxicity to the receiving bodies. As an alternative, immobilized catalysts can be applied; however, this strategy can increase the difficulties in mass and photo transfer. This work presents the development of an inert and highly porous support for TiO2 immobilization. The produced materials have a high surface area and contribute to diminishing the difficulties in mass and phototransfer during photocatalysis. Different types of polymeric materials were tested as support, and a Taguchi experimental design with an L9 arrangement was used to optimize the immobilization process and evaluate the effect of TiO2 content and the use of bidding agents, ultrasound, and thermic treatment. The grey automotive polyurethane foam proved to be the best support, using 5.0% of TiO2 (wt.%) in the immobilization suspension with Triton X as the binding agent and heat treatment during immobilization. At the optimal conditions, it was possible to achieve total As(III) oxidation (below the analytical detection limit) in 240 min, with nearly 100% As(V) present in solution at the end of the reaction (almost no As adsorption on the catalyst surface). In addition, the catalytic bed was able to promote the As(III) complete oxidation in up to five consecutive cycles without significant leaching or deactivation of the immobilized TiO2.

1. Introduction

Photocatalysis is a widely utilized method for treating water and wastewater containing toxic organic pollutants, heavy metals, and microorganisms [1,2]. TiO2 is the most utilized photocatalyst, and even though it appears to be a promising material, in the vast majority of studies, suspensions containing powder TiO2 are utilized. This approach highlights the major drawbacks of photocatalytic processes, including the limitation of UV light penetration due to the turbidity caused by dispersed nanoparticles [3], the difficulty in reusing the catalyst after the water treatment [4], and the need for a final filtration step to avoid the presence of nanoparticles in the final effluent [5]. To overcome these limitations, TiO2 films can be prepared on different types of inert substrates, eliminating the need for a post-filtration step and allowing the catalyst to be reused [6].
On the other hand, the use of TiO2 films increases the mass and photon transfer limitations, making “pollutants molecules-catalyst” contact more difficult and limiting the number of photons that reach the catalyst surface. Therefore, the ideal support should have a high surface area [7] while allowing the radiation to pass through its outer structure [8]. Several types of materials have been tested as catalyst supports, including glass, paper, ceramics, fiberglass, cellulose acetate sheets, alginate and chitosan spheres, and stainless steel, among others [5,9,10]. Likewise, different immobilization methods can be applied, such as chemical/physical vapor deposition, sputtering, electrospinning, dip-coating and sol–gel methods [8,11]. The commercial sponge-type polyurethane (PU) is an interesting candidate since it is highly porous, chemically inert, mechanically stable, flexible and elastic, easily shaped, and readily available at a low cost [12,13]. Furthermore, due to its thermoplastic properties, it is possible to enhance the TiO2 fixation on a PU surface by using a simple thermal treatment method [14,15,16]. Despite these advantages, only a few studies have been carried out using PU as a support. Shoaebargh et al. [15] tested a new hybrid photodegradation-enzymatic process by immobilizing glucose oxidase (GOx) and TiO2 on a PU surface inserted inside a photobioreactor. The authors, however, did not investigate the parameters that influence TiO2 immobilization on PU or the catalyst’s reuse potential.
An in depth study of the parameters that affect the preparation of the catalyst films is fundamental to ensuring proper immobilization as well as the material’s durability. Besides the choice of the appropriate support, the immobilization method, the use of binder additives, and thermic and ultrasound treatment should be considered. The addition of additives, such as Triton X, glutaraldehyde, citric acid, and sodium alginate, may allow the formation of stronger chemical bonds between the carrier and the catalyst or aid in the dispersion of TiO2 nanoparticles in the medium, preventing the formation of agglomerates and producing more homogeneous films. The latter effect can also be achieved by ultrasound treatment of the starting catalyst suspension, while thermic treatment may allow better fixation of the catalyst on the support surface [17,18].
Therefore, the present work aims to evaluate the performance of different 3D highly porous structures as well as the most suitable immobilization method for the production of TiO2 thin films and their application in photocatalysis. To the best of our knowledge, a simultaneous evaluation of the parameters that affect the preparation of the catalyst films has not yet been carried out. This approach is relevant since synergistic effects can also be tracked. The efficiency in the oxidation of As(III) to As(V), catalyst leaching, and photocatalyst reuse were also evaluated. The oxidation of As(III) to As(V) in aqueous solution was chosen as the model reaction since the inorganic As(III) is described as approximately 70 times more toxic than inorganic As(V) [19]. While the two forms of arsenic are toxic, As(V) is more easily removed by precipitation or adsorption processes than As(III). Both aspects covered in the present study, TiO2 immobilization and As removal, are of fundamental importance to the sustainable use of water resources. The TiO2 immobilization process helps to suppress the discharge of nanoparticles in the aquatic environment, minimizing the impacts related to the discharge of nanoparticles and allowing the catalyst to be reused. In addition, arsenic is first on the priority list of the Agency for Toxic Substances and Disease Registry (ATSDR) due to its toxicity combined with the elevated potential its human exposure [20]. The newly developed material can be used as part of an efficient strategy for the removal of toxic pollutants such as arsenic, contributing to the sustainable recovery of polluted environments.

2. Materials and Methods

2.1. Chemicals

As(III) solutions were prepared from NaAsO2 (Vetec, Duque de Caxias, Brazil, purity 98%). Nitric acid (HNO3—Neon, Suzano, Brazil, purity 65%) was used to acidify the As(III) standards and samples for atomic absorption spectrometry analysis and material washing. An arsenic stock solution (As2O5, 1.0 g L−1 in 2% HNO3, Sigma Aldrich, Duque de Caxias, Brazil) was used to prepare As(V) standard solutions through appropriate dilutions. Ammonium molybdate tetrahydrate ((NH4)6Mo7O24·4H2O—Qhemis, Jundiaí, Brazil, purity 83%) and ascorbic acid (C6H8O6—Vetec, Duque de Caxias, Brazil, purity 99%) were used as colorimetric reagents to determine As(V) concentration and nickel nitrate hexahydrate (Ni(NO3)2·6H2O—Dinâmica, Indaiatuba, Brazil, purity 99%) was used as a modifier during the As(total) determination. TiO2 Degussa P25 (Evonik, Essen, Germany) powder was used as delivered without further modification or purification.
Triton X-100 ((C2H4O)nC14H22O, Dinâmica, Indaiatuba, Brazil), glutaraldehyde (C5H8O2—Vetec, Duque de Caxias, Brazil, purity 25%), citric acid (C6H8O7·H2O—Nuclear, Diadema, Brazil, purity 99%), and sodium alginate ((NaC6H7O6)n—Vetec, Duque de Caxias, Brazil purity 98%) were used as additives in the TiO2-P25 coating suspensions. Ammonium sulfate (Dinâmica, Indaiatuba, Brazil, purity 99%) and sulfuric acid (Química Moderna, Barueri, Brazil, purity 97%, 1.84 g/cm3) were used to dissolve TiO2 nanoparticles, and hydrogen peroxide (Lafan, Várzea Paulista, Barzil, purity 35% (w/v)) was used as a colorimetric reagent to determine the concentration of TiO2 leached during the oxidation reaction. Iron(III) chloride hexahydrate (Vetec, Duque de Caxias, Brazil, purity 97%), 1,10-Phenanthroline 1-hydrate (Dinâmica, Indaiatuba, Brazil, purity 99%), oxalic acid dehydrate (Neon, Suzano, Brazil, purity 98%), glacial acetic acid (Synth, Diadema, Brazil), and ammonium acetate (Dinâmica, Indaiatuba, Brazil, purity 98%) were used for actinometry. The above-mentioned sulfuric acid and sodium hydroxide (Lafan, Várzea Paulista, Barzil, purity 97%) were used for pH adjustment. All samples were filtered through 0.45 μm cellulose acetate membranes (Filtrilo, Colombo, Brazil) before analysis.

2.2. Analytical Determinations

Total arsenic concentration was determined by flame atomic absorption spectrometry (FAAS, model AA 6300, Shimadzu, Kyoto, Japan), with acetylene-N2O flame, a spectral slit width of 0.7 nm, a wavelength of 197.3 nm, a 25-mA lamp current, and continuous background correction with a deuterium lamp, presenting a limit of detection of 2.67 × 10−2 mM. Standards and diluted samples were prepared using ultrapure water acidified with HNO3 (0.02 mM). Nickel solution (0.1 M) was used as a modifier [21]. The pentavalent arsenic concentration was determined by molecular absorption spectrometry (UV-Vis, model Cirrus 80, Femto, São Paulo, Brazil), with a detection limit of 1.7 × 10−4 mM, through the molybdate blue method [22]. The concentration of leached TiO2 was also determined by molecular absorption spectrometry with a detection limit of 1.0 × 10−2 mM, through the hydrogen peroxide method [23,24]. The photon flow for each reactor configuration was determined by the potassium ferrioxalate actinometry method [25].

2.3. Supports

Cellulose acetate monoliths (CAM, Wacotech GmbH & Co. KG, Herford, Germany—Figure S1a), green synthetic fiber (GSF, Scotch Brite 3M, Ribeirão Preto, Brazil—Figure S1b), blue commercial polyurethane sponge (BCP, Bettanin, São José dos Campos, Brazil—Figure S1c) and a grey automotive polyurethane sponge (GPF, FoamPartner GmbH, Wolfhausen, Switzerland—Figure S1d) were tested as support materials for the TiO2 coating. To determine the mass changes during heating, all materials were submitted to a thermogravimetric analysis (TGA, model Jupiter STA 449 F3, Netzsch, Selb, Germany) in a temperature ranging from 25 °C to 1000 °C. Then, to determine the glass transition temperature, a differential scanning calorimetry (DSC, model Jade, Perkin Elmer, Shelton, CT, USA) was carried out, with a heating rate of 10 °C min−1.
The electron microscopy analyses (SEM) were performed at the Central Laboratory of Electronic Microscopy (LCME-UFSC, Florianópolis, Brazil) with the JEOL JSM-6390LV microscope (Thermo Cientific, Brno, Czech Republic). All samples were analyzed with an acceleration voltage of 10 kV. For the chemical characterization of the samples, an elementary analysis was performed by energy dispersive X-ray spectrometry (EDS).

Catalyst Preparation

Before use, all materials used as support were immersed in ultrapure water solution with alkaline detergent (Derquim LM 01, Panreac Química, Barcelona, Spain) and sonicated in an ultrasonic device (USC 2500—Unique) for 1 h at 50 kHz. The materials were then washed with ultrapure water in abundance and then dried in an air recirculation oven (MA035—Marconi) at 50 °C for 24 h.
The photocatalyst aqueous suspensions containing 2 wt.% of TiO2-P25 and ultrapure water were prepared and sonicated for 10 min at 50 kHz in order to better disperse the particles. A catalyst layer was deposited by immersing the polymeric structures into the aqueous suspension at a withdrawal rate of 0.5 mm s−1, assuring a thin and uniform film on each substrate surface. The samples were dried at 50 °C for 1 h in the air recirculation oven. The quantity of catalyst deposited was controlled by weighing the coated supports in a precision balance (model ABS204-S, Mettler Toledo, Barueri, Brazil).
The structures were then immersed in a 100 mL ultrapure water solution through a nylon strip under 60 rpm mechanic shaking (model NL-01-01-A, New Lab, Piracicaba, Brazil) for 5 min. The samples were newly dried at 50 °C for 1 h and weighed after cooling in order to verify the amount of leached catalyst.

2.4. Chemical Agents

To improve the catalytic immobilization process, different types of additives were tested. Triton X (TX) was used as a dispersant to increase the mass of immobilized catalyst on the substrate surface [16]. Glutaraldehyde (GL) and citric acid (CA) were used as cross-linking agents [17,18]. Different suspensions were prepared, containing 2 wt.% of TiO2-P25, 2 wt.% of the selected additive in ultrapure water. Then, the same immobilization procedure and leaching test previously applied were also used to evaluate the additives.

2.5. Experimental Design

Due to the large number of variables that affect catalytic immobilization, an experimental design was applied to optimize the process. Four influencing parameters (control factors) were chosen for the statistical analysis: (i) catalyst load in the coating suspension; (ii) type of binder agent; (iii) amount of binder agent; and (iv) ultrasound dispersion of the catalyst.
To determine the effects of control factors on the TiO2 immobilization, three levels were considered for each factor, as shown in Table 1. Therefore, to decrease the number of experiments and obtain the optimal conditions, the Taguchi method was chosen [26]. The total degree of freedom for four factors at three levels each is 8. Therefore, the L9 orthogonal array with 9 experimental positions was selected, and experiments were designed accordingly (Table 2). The experimental runs were determined by Statistica 12 software and performed in duplicate to minimize the noise factor.
The TiO2 immobilization on the supports was carried out using the previously described method but applying the ultrasonic cavitation technique with a high-intensity ultrasonic horn (20 kHz, Sonic Dismembrator Model 500, Fischer Scientific, Waltham, MA, USA). The drying methods were defined based on the additives’ boiling points and on previous published studies [16,17,18].
To facilitate the data analysis, a result factor was utilized (Response A, Equation (1), Table 1). Response A was based on the combination of five indicators: (i) the TiO2 % immobilized on the support (xi); (ii) the As(III) % oxidized after 60 min of photocatalytic reaction (gi, As(V) %); (iii) the TiO2 % remaining immobilized after the reaction (zi); (iv) the As(III) % oxidized after reuse assay (ki, As(V) %); and (v) the TiO2 % remaining immobilized after reuse (hi).
The Taguchi method involves performing analysis of means (ANOM) on raw data of output responses. ANOM indicates the significant factors, with their contribution to the output response (Δ), and the configuration of factor levels that maximize factor A. The range of mean values (Δ) was calculated for each factor; a large range implies a high influence on factor A. In addition, the statistical analysis of variance (ANOVA) was also performed to reveal which factor was statistically significant for Response A. From the predicted results, the condition for the maximum Response A was obtained. An experimental run was performed at the predicted condition to validate the predicted optimum Response A.

2.6. Photochemical Reactors

2.6.1. Batch Reactor with External Lighting

The As(III) oxidation reactions, during the determination of the best coating method, were carried out in a batch reactor with external light (Figure S2). The setup consisted of a 150 mL capacity reservoir with magnetic stirring (New Lab, Piracicaba, Brazil). The radiation source was the same UVA lamp (Philips Actinic BL TL TL/10 1FM/10X25CC, 6 W, λmax = 365 nm), fixed 13 cm above the solution level in the reservoir. The potassium ferrioxalate actinometry assays indicated a photonic flux of 7.80 × 10−8 Einstein s−1.

2.6.2. Annular Reactor

The catalyst reuse tests were performed in a lab scale annular reactor (Figure S3) during consecutive As(III) oxidation cycles. The total treated and illuminated volume were 1.5 L and ~700 mL, respectively. Table S1 summarizes the main characteristics of the reactor. Previous assays of potassium ferrioxalate actinometry indicated a photonic flux of 1.67 × 10−6 Einstein s−1.

2.7. As(III) Oxidation Experimental Procedure

The initial As(III) concentration used in all experiments was 0.27 mM. In the annular reactor, the jacket temperature was maintained at 20 °C, pH 8.0, pump flow of 1.6 L min−1, and the reservoir was stirred at 60 rpm. A total volume of 1.5 L was recirculated in the system. Experiments in the batch reactor with external light were carried out with 100 mL of As(III) solution, pH controlled at 8.0, agitation at 60 rpm, and without temperature control. In both reactors, the samples were collected after the addition of each reagent, before and after the lamp connection, and at predetermined times.

3. Results and Discussion

3.1. Effect of Polymeric Support Type in the TiO2 Immobilization

To determine the effect of the support material in the TiO2 immobilization, all polymeric structures were equally cut (1 × 1 × 1 cm) and evaluated over three parameters: (i) load of TiO2 supported; (ii) thermic stability; and (iii) glass transition temperature. Regarding the amount of catalyst immobilized on the support surface and the amount that remained fixed after immersion in ultrapure water, it is possible to observe from Figure 1 that the grey automotive polyurethane sponge (GPF) foam was the material capable of retaining a larger mass of TiO2 adhered in the formed film. In addition, the film formed on this material surface was also the one that presented lower catalyst leaching.
To evaluate the material’s thermic stability, their degradation temperature was determined by thermogravimetric analyses (TGA). It can be noticed by Figure 2 that the material’s thermic stability increases in the following order: green synthetic fiber (GSF) < cellulose acetate monoliths (CAM) < blue commercial polyurethane sponge (BCP) < grey automotive polyurethane sponge (GPF), presenting the beginning of mass loss at around 146, 178, 185, and 253 °C, respectively. These results indicated that the GPF structure is the material that could better withstand some type of thermal treatment for the fixation of the catalyst.
Finally, DSC analysis was performed to obtain qualitative and quantitative information on physical and chemical changes involving endothermic processes (heat absorption), exothermic processes (release of heat), or changes in heat capacity. Therefore, the main objective of the analysis was to determine the material’s glass transition temperature, which is the temperature where polymer molecules turn from the solid phase to a vitreous state. The glass transition temperature is easily identified by a step in the baseline of the measurement curve [27]. In this condition, the material’s internal energy is still not sufficient for melting but can still modify some of the material’s properties, such as heat capacity, coefficient of thermal expansivity, and viscoelasticity [28]. By discovering the glass transition temperature, controlled heat treatment can be applied in the structure coated with TiO2, leading to deeper penetration of the catalyst into the polymer structure and an increase in catalyst adhesion after cooling. From Figure 3, it is possible to observe an initial endothermic behavior (melting process) related to the dehydration process, followed by an exothermic behavior (decomposition process). The valley observed during CAM analysis (Figure 3a) indicates that the crystallization process for this material is in the range 175–200 °C [29]. Furthermore, no changes in the baseline for the CAM, GSF, and BCP structures were observed. Thus, these three materials do not present a glass transition temperature, or, at least, it was not identified. By analyzing the endothermic peaks, it can be concluded that the melting of the materials occurs at approximately 183, 275, and 306 °C, respectively. Nevertheless, for the GPF, it was possible to notice the baseline change, making evident the material glass transition range, which occurs between approximately 215 and 237 °C with an average value of 226 °C.
Due to the higher thermic stability and the identification of the glass transition temperature of the gray polymeric foam (GPF), it was chosen to be used in the next experiments. In addition, this foam has higher and more dispersed pores, which minimize photon transfer limitations. Furthermore, the GPF also presents: (i) high surface area; (ii) stability against degradation by strong oxidative radicals; (iii) no toxicity, chemical, or mechanical stability; (iv) high durability; (v) hydrophobic nature; (vi) low cost and availability [12,13,14,15].

3.2. Coating Method Optimization

For the experimental design process, the TiO2 immobilization tests were performed using GPF structures with dimensions of 2 × 2 × 2 cm and a batch reactor with an external UVA lamp for the As(III) photooxidation reaction. As the objective was to optimize the process of catalyst coating, the Taguchi method with the maximize the means (direct observation) characteristic was chosen. With this analysis, the responses versus the levels of different factors can be observed directly from a broken line plot. The mean value of Response A for the corresponding control factors at each level was calculated according to the assignment of the experiment.
The experiment sequence was conducted in a random order to avoid the incorporation of systematic errors. The raw results are shown in the supplementary material (Table S2), along with the equations for normalizing the output variables used in the experimental design. At each experimental level, two samples were prepared to minimize the noise factor. The normalized parameters mean and the resultant Response A are shown in Table 3 and were used to predict the optimum configuration to maximize Response A. The mean of the correspondent Response A at different levels of the control parameters is summarized in Table S3 (ANOM). The corresponding diagrams of direct observation analysis for the individual effects of various parameters at each level are shown in Figure 4.
From Figure 4, it is possible to observe that the optimal coating condition should be achieved using 5% of TiO2 (wt.%) in the coating suspension, Triton X (TX) as the additive with a concentration of 3% (wt.%), and without the use of the ultrasonic horn. The difference in Response A using different TiO2 concentrations in the coating suspension can be explained by the mass transfer resistance, which decreases when there is a larger mass of catalyst in the suspension. For the same reason, Triton X presents the best results, as it aids in the dispersion of the catalyst nanoparticles in the coating suspension and thus increases the chance of contact between the TiO2 and the support [30]. The fact that the 3% concentration of Triton X has been more efficient can be explained by the solubility of the additive, since at higher concentrations it does not dissolve and precipitates at the bottom of the coating suspension recipient with a part of the catalyst. The ultrasonic cavitation at high intensity (70%) may negatively affect Response A since the energy supplied by the ultrasonic horn can contribute to the removal of the adhered catalyst particles from the support surface. Since no significant difference between the processes without horn (WH) and with ultrasonic horn at 30% intensity were noticed, it was cheaper and more convenient to not use it.
The difference between the maximal and minimum values for Response A at each factor is defined as the extreme difference, which directly reflects the effect of each factor. The larger the extreme difference, the more significant the controlling factor is. As can be seen from Table S3, the highest difference is related to horn intensity, followed by additive concentration, TiO2 concentration, and additive type. Analysis of variance (ANOVA) was performed to determine the significance of each coating parameter. Statistically, the F-test provides the information, at a certain confidence level, for which the results are significantly different. As can be seen in Table 4, all the variables (factors) have a large F-value, indicating that the variation in each process parameter implies a significant change in the coating results.

Confirmation Experiment

The optimal Response A value predicted by the Taguchi analysis is shown in Table 5. As can be seen from the orthogonal matrix L9 shown in Table 3, there is no experiment arrangement that involves the optimal configuration found by the statistical analysis. Therefore, it is necessary to perform an additional experiment with this configuration to verify if the response obtained is indeed the optimal one. The confirmation experiment resulted in an experimental Response A (45.13%) similar to the predicted one by the Taguchi method (45.49%), with a difference of only 0.36%. With the validation of the developed method, the optimal levels of the factors that maximize Response A were utilized to prepare enough supported catalyst to be used in the lab-scale annular photoreactor.

3.3. As(III) Oxidation by GPF-TiO2/UVA-Irradiated Annular Reactor

The As(III) photocatalytic oxidation by TiO2 can be driven by three different mechanisms through successive one-electron steps: (i) direct oxidation by TiO2 photogenerated hole ( h V B + ), (ii) indirect oxidation by OH generated by the reaction of H2O, and h V B + , and (iii) indirect oxidation by O 2 / H O 2 generated by the reaction of O2 and e C B [31]. Figure 5 shows a schematic diagram for the photocatalytic oxidation of As(III) by TiO2 heterogeneous photocatalysis.
After the optimization of the TiO2 immobilization process, the use of GPF-TiO2 in the annular reactor system was evaluated to promote As(III) oxidation. In this case, the total mass of TiO2 immobilized on the support was 293 mg (Figure S4). From Figure 6, it is possible to observe the complete oxidation of As(III) after 240 min of photocatalytic reaction during the first use of GPF-TiO2. Although this result seems much slower than the ones reported by other authors dealing with As(III) photooxidation, the vast majority of available studies use TiO2 in suspension. Furthermore, the different operational conditions and lack of standardization of the reported results make the direct comparison of experiments difficult. As an example, using the TiO2@Fe3O4 photocatalyst, Xiao et al. [32] achieved complete As(III) oxidation in only 4 min of photocatalytic reaction. However, since the authors used 10 mg of catalyst in suspension in 40 mL of treated solution with an As(III) initial concentration of 10 mg L−1, the direct comparison with the results obtained in the present article is very imprecise.
Since this is a catalytic process, an important parameter is the catalyst’s potential to be reused. Thus, the GPF-TiO2 structures (Figure S4) were reused for six consecutive oxidation cycles using fresh solutions of 0.27 mM As(III) with a pH of 8 at 20 °C. After each experiment, the support was immersed in ultrapure water, sonicated for 5 min at 50 kHz, dried at 50 °C for 1 h, and weighed after cooling. As expected, the kinetic constant slightly decreased (Table S4) as the support was reused, from 7.0 × 10−3 min−1 in the 1st cycle to 6.1 × 10−3 min−1 in the 5th cycle. Still, during the first five cycles, complete As(III) oxidation was successfully achieved in up to 300 min of reactions. Similar results were reported by Fausey et al. [33] in the oxidation of 1 mg L−1 of As(III) by immobilized TiO2 nanoparticles in electrospun nanofibers modified with reduced graphene (rGO-TiO2@fibers).
During the 6th reuse cycle, however, the complete oxidation of As(III) was not achieved, and 0.022 mM (or 8.0%) of As(III) was remaining in solution after 300 min of reaction. It was also observed an increase in the As(V) adsorption over the consecutive reuse cycles. This behavior was probably related to the increase in the amount of leached TiO2 over the cycles, causing small flaws in the film and eventually favoring adsorption. While in the 1st cycle, the As(Total) adsorbed was only 0.01 mM, at the end of the 6th cycle, this value was increased to 0.13 mM. The loss of TiO2 per leaching was practically constant for each reuse, resulting in similar concentrations. The total TiO2 mass lost in the 6 oxidation cycles was 15 mg, which leaves 278 mg (~95%) of catalyst immobilized on the support surface. During the 6 reaction cycles, a total solution volume of 9 L was treated, thus the total TiO2 concentration in the final effluent was 1.7 mg L−1. According to Kahru and Dubourguier [34], although the toxicity of TiO2 depends on the organism type, these nanoparticles are harmful to the environment at concentrations higher than 10 mg L−1. Therefore, the amount leached from GPF-TiO2 is smaller than the limit that poses a risk to aquatic systems, evidencing the contribution of the present work to a more sustainable process.

GPF-TiO2 Characterization

To investigate the surface of the GPF structures, the support was evaluated through scanning electron microscopy (SEM) before TiO2 deposition (sample Y), after TiO2 deposition (sample A), and after six reuse cycles of As(III) oxidation in the annular reactor system (sample A6). In sample Y (Figure 7a), it is possible to observe a smooth surface of the initial GPF structure. In sample A (Figure 7b), it is possible to visualize the TiO2 particles well adhered to the surface, forming a uniform film. Since the heat treatment reached the glass transition temperature of the polymer, the catalyst particles seem fused to the surface of the material, which is in accordance with the low leaching rate during the reuse tests. Finally, this effect can also be noted in sample A6 (Figure 7c), since it shows a very similar surface to that of unused TiO2 film (sample A, Figure 7b). The EDS analysis corroborates this observation since the Ti peak was detected in both A and A6 samples (Figure 7e,f, respectively). In sample Y (Figure 7d), as expected, it was not possible to detect the presence of a catalyst. In addition, in sample A6 (Figure 7f), it is possible to identify a small peak related to the presence of the arsenic element, which was already expected since a small percentage of the pollutant was adsorbed on the surface of the support after 6 oxidation cycles.

4. Conclusions

The oxidation of As(III) species was successfully achieved using the annular UVA photoreactor packed with the highly porous 3D GPF structure coated with TiO2, allowing complete oxidation of the contaminant in 240 min with only 3.8% of As(total) adsorption. The optimization of the TiO2 immobilization method through the Taguchi L9 orthogonal array experimental design allowed the preparation of a material with good film stability. In fact, the prepared support could be used for five reaction cycles, promoting As(III) complete oxidation in 300 min without significant leaching of the immobilized TiO2. Therefore, the GPF support prepared through the developed methodology reached the proposed objectives and can be a viable alternative for the oxidation of As(III) in aqueous medium, mainly due to the low cost of production and the reusability.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/su152014760/s1, Figure S1. Materials used as a support; (a) cellulose acetate monoliths—CAM; (b) green synthetic fiber—GSF; (c) blue commercial polyurethane sponge—BCP; (d) grey polyurethane sponge—GPF. Figure S2. Scheme (a) and photo (b) of the batch reactor with external UVA lamp. Figure S3. Scheme (a) and photo (b) of the annular photoreactor illuminated by UVA lamp. Figure S4. Sample A applied in the photocatalytic oxidation of As(III) in an annular photoreactor illuminated by UVA lamp (As (III) = 0.27 mM, pH 8.0 and 20 °C). Table S1. Photoreactor geometrical characteristics. Table S2. Non-normalized output variables response and experimental randomized sequence. Table S3. Average of Response A for each factorial level in particular (ANOM). Table S4. Pseudo first-order kinetic constants with the corresponding determination coefficient (R2) for the reuse cycles of sample A in As(III) oxidation ([As (III)] = 0.27 mM) by UVA/GPF-TiO2 annular reactor system at pH 8.0 and 20 °C.

Author Contributions

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

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data supporting will be provided if requested.

Acknowledgments

The authors are grateful to LABMASSA—Mass Transfer Laboratory, POSENQ-UFSC—for the available laboratorial infrastructure. This study was supported in part by the Coordenação de Aperfeiçoamento de Pessoal de Nível Superior-Brasil (CAPES)-Finance Code 001.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gopalakrishnan, G.; Jeyakumar, R.B.; Somanathan, A. Challenges and Emerging Trends in Advanced Oxidation Technologies and Integration of Advanced Oxidation Processes with Biological Processes for Wastewater Treatment. Sustainability 2023, 15, 4235. [Google Scholar] [CrossRef]
  2. Marinho, B.A.; De Liz, M.V.; Tiburtius, E.R.L.; Nagata, N.; Peralta-Zamora, P. TiO2 and ZnO mediated photocatalytic degradation of E2 and EE2 estrogens. Photochem. Photobiol. Sci. 2013, 12, 678–683. [Google Scholar] [CrossRef] [PubMed]
  3. Chairungsri, W.; Pholchan, P.; Sumitsawan, S.; Chimupala, Y.; Kijjanapanich, P. Photocatalytic Degradation of Textile Dyeing Wastewater Using Titanium Dioxide on a Fixed Substrate: Optimization of Process Parameters and Continuous Reactor Tests. Sustainability 2023, 15, 12418. [Google Scholar] [CrossRef]
  4. Jones, M.; Li, K.; Chen, M.; Chen, L.; Zhao, S.; Xue, W.; Han, Y. Synthesis and Application of a Fe3O4/Ag3PO4/g-C3N4 Magnetic Composite Photocatalyst for Sulfonamide Antibiotics Degradation. Sustainability 2023, 15, 13279. [Google Scholar] [CrossRef]
  5. Hoang, N.T.T.; Nguyen, D.D. Improving the Degradation Kinetics of Industrial Dyes with Chitosan/TiO2/Glycerol Films for the Sustainable Recovery of Chitosan from Waste Streams. Sustainability 2023, 15, 6979. [Google Scholar] [CrossRef]
  6. Espíndola, J.C.; Cristóvão, R.O.; Mendes, A.; Boaventura, R.A.R.; Vilar, V.J.P. Photocatalytic membrane reactor performance towards oxytetracycline removal from synthetic and real matrices: Suspended vs immobilized TiO2-P25. Chem. Eng. J. 2019, 378, 122114. [Google Scholar] [CrossRef]
  7. de Bittencourt, M.A.; Novack, A.M.; Filho, J.A.S.; Mazur, L.P.; Marinho, B.A.; da Silva, A.; de Souza, A.A.U.; de Souza, S.M.A.G.U. Application of FeCl3 and TiO2-coated algae as innovative biophotocatalysts for Cr(VI) removal from aqueous solution: A process intensification strategy. J. Clean. Prod. 2020, 268, 122164. [Google Scholar] [CrossRef]
  8. da Costa Filho, B.M.; Araujo, A.L.P.; Padrão, S.P.; Boaventura, R.A.R.; Dias, M.M.; Lopes, J.C.B.; Vilar, V.J.P. Effect of catalyst coated surface, illumination mechanism and light source in heterogeneous TiO2 photocatalysis using a mili-photoreactor for n-decane oxidation at gas phase. Chem. Eng. J. 2019, 366, 560–568. [Google Scholar] [CrossRef]
  9. Wan, W.; Li, Y.; Bai, S.; Yang, X.; Chi, M.; Shi, Y.; Liu, C.; Zhang, P. Three-Dimensional Porous ZnO-Supported Carbon Fiber Aerogel with Synergistic Effects of Adsorption and Photocatalysis for Organics Removal. Sustainability 2023, 15, 13088. [Google Scholar] [CrossRef]
  10. Alardhi, S.M.; Abdalsalam, A.H.; Ati, A.A.; Abdulkareem, M.H.; Ramadhan, A.A.; Taki, M.M.; Abbas, Z.Y. Fabrication of polyaniline/zinc oxide nanocomposites: Synthesis, characterization and adsorption of methylene orange. Polym. Bull. 2023, 1–27. [Google Scholar] [CrossRef]
  11. Espíndola, J.C.; Cristóvão, R.O.; Santos, S.G.S.; Boaventura, R.A.R.; Dias, M.M.; Lopes, J.C.B.; Vilar, V.J.P. Intensification of heterogeneous TiO2 photocatalysis using the NETmix mili-photoreactor under microscale illumination for oxytetracycline oxidation. Sci. Total Environ. 2019, 681, 467–474. [Google Scholar] [CrossRef] [PubMed]
  12. Nuthalapati, K.; Sheng, Y.-J.; Tsao, H.-K. Ag NPs-coated polyurethane sponge as a water filter for removal of toxic metal ions at high concentrations. Chemosphere 2023, 343, 140266. [Google Scholar] [CrossRef] [PubMed]
  13. Akay, O.; Altinkok, C.; Acik, G.; Yuce, H.; Ege, G.K. A bio-based and non-toxic polyurethane film derived from Luffa cylindrica cellulose and ʟ-Lysine diisocyanate ethyl ester. Eur. Polym. J. 2021, 161, 110856. [Google Scholar] [CrossRef]
  14. Magalhães, F.; Moura, F.C.C.; Lago, R.M. TiO2/LDPE composites: A new floating photocatalyst for solar degradation of organic contaminants. Desalination 2011, 276, 266–271. [Google Scholar] [CrossRef]
  15. Shoaebargh, S.; Karimi, A. RSM modeling and optimization of glucose oxidase immobilization on TiO2/polyurethane: Feasibility study of AO7 decolorization. J. Environ. Chem. Eng. 2014, 2, 1741–1747. [Google Scholar] [CrossRef]
  16. Liu, J.; Li, Y.; Arumugam, S.; Tudor, J.; Beeby, S. Investigation of Low Temperature Processed Titanium Dioxide (TiO2) Films for Printed Dye Sensitized Solar Cells (DSSCs) for Large Area Flexible Applications. Mater. Today Proc. 2018, 5, 13846–13854. [Google Scholar] [CrossRef]
  17. Comí, M.; Lligadas, G.; Ronda, J.C.; Galià, M.; Cádiz, V. Adaptive bio-based polyurethane elastomers engineered by ionic hydrogen bonding interactions. Eur. Polym. J. 2017, 91, 408–419. [Google Scholar] [CrossRef]
  18. da Rosa Schio, R.; da Rosa, B.C.; Gonçalves, J.O.; Pinto, L.A.A.; Mallmann, E.S.; Dotto, G.L. Synthesis of a bio–based polyurethane/chitosan composite foam using ricinoleic acid for the adsorption of Food Red 17 dye. Int. J. Biol. Macromol. 2019, 121, 373–380. [Google Scholar] [CrossRef]
  19. Carneiro, M.A.; Pintor, A.M.A.; Boaventura, R.A.R.; Botelho, C.M.S. Efficient removal of arsenic from aqueous solution by continuous adsorption onto iron-coated cork granulates. J. Hazard. Mater. 2022, 432, 128657. [Google Scholar] [CrossRef]
  20. ATSDR. Substance Priority List. Available online: https://www.atsdr.cdc.gov/spl/index.html#2022spl (accessed on 6 October 2023).
  21. Water Environment Federation. Standard Methods for the Examination of Water and Wastewater; Water Environment Federation: Alexandria, WV, USA, 1999. [Google Scholar]
  22. Lenoble, V.; Deluchat, V.; Serpaud, B.; Bollinger, J.C. Arsenite oxidation and arsenate determination by the molybdene blue method. Talanta 2003, 61, 267–276. [Google Scholar] [CrossRef]
  23. Jackson, N.B.; Wang, C.M.; Luo, Z.; Schwitzgebel, J.; Ekerdt, J.G.; Brock, J.R.; Heller, A. Attachment of TiO2 Powders to Hollow Glass Microbeads: Activity of the TiO2—Coated Beads in the Photoassisted Oxidation of Ethanol to Acetaldehyde. J. Electrochem. Soc. 1991, 138, 3660–3664. [Google Scholar] [CrossRef]
  24. Ashley, S.E.Q. Colorimetric Determination of Traces of Metals. By E. B. Sandell. J. Phys. Chem. 2002, 49, 263–264. [Google Scholar] [CrossRef]
  25. Kuhn, H.J.; Braslavsky, S.E.; Schmidt, R. Chemical actinometry (IUPAC technical report). Pure Appl. Chem. 2004, 76, 2105–2146. [Google Scholar] [CrossRef]
  26. Taguchi, G.; Konishi, S. Taguchi Methods, Orthogonal Arrays and Linear Graphs, Tools for Quality American Supplier Institute; ASI Press: Fremont, CA, USA, 1987; pp. 8–35. [Google Scholar]
  27. Šudomová, L.; Weissmannová, H.D.; Steinmetz, Z.; Řezáčová, V.; Kučerík, J. A differential scanning calorimetry (DSC) approach for assessing the quality of polyethylene terephthalate (PET) waste for physical recycling: A proof-of-concept study. J. Therm. Anal. Calorim. 2023, 148, 10843–10855. [Google Scholar] [CrossRef]
  28. Técnicas de Caracterização de Polímeros. 2004. Available online: www.artliber.com.br (accessed on 10 September 2023).
  29. Capitain, C.; Ross-Jones, J.; Möhring, S.; Tippkötter, N. Differential scanning calorimetry for quantification of polymer biodegradability in compost. Int. Biodeterior. Biodegrad. 2020, 149, 104914. [Google Scholar] [CrossRef]
  30. Nam, J.G.; Lee, E.S.; Jung, W.C.; Park, Y.J.; Sohn, B.H.; Park, S.C.; Kim, J.S.; Bae, J.Y. Photovoltaic enhancement of dye-sensitized solar cell prepared from [TiO2/ethyl cellulose/terpineol] paste employing TRITONTM X-based surfactant with carboxylic acid group in the oxyethylene chain end. Mater. Chem. Phys. 2009, 116, 46–51. [Google Scholar] [CrossRef]
  31. Meichtry, J.M.; Levy, I.K.; Mohamed, H.H.; Dillert, R.; Bahnemann, D.W.; Litter, M.I. Mechanistic Features of the TiO2 Heterogeneous Photocatalysis of Arsenic and Uranyl Nitrate in Aqueous Suspensions Studied by the Stopped-Flow Technique. ChemPhysChem 2016, 17, 885–892. [Google Scholar] [CrossRef]
  32. Xiao, M.; Li, R.; Yin, J.; Yang, J.; Hu, X.; Xiao, H.; Wang, W.; Yang, T. Enhanced photocatalytic oxidation of As(III) by TiO2 modified with Fe3O4 through Ti—O—Fe interface bonds. Colloids Surfaces A Physicochem. Eng. Asp. 2022, 651, 129678. [Google Scholar] [CrossRef]
  33. Fausey, C.L.; Zucker, I.; Shaulsky, E.; Zimmerman, J.B.; Elimelech, M. Removal of arsenic with reduced graphene oxide-TiO2-enabled nanofibrous mats. Chem. Eng. J. 2019, 375, 122040. [Google Scholar] [CrossRef]
  34. Kahru, A.; Dubourguier, H.C. From ecotoxicology to nanoecotoxicology. Toxicology 2010, 269, 105–119. [Google Scholar] [CrossRef]
Figure 1. Mass of TiO2 immobilized on the surface of the analyzed materials, before () and after washing ().
Figure 1. Mass of TiO2 immobilized on the surface of the analyzed materials, before () and after washing ().
Sustainability 15 14760 g001
Figure 2. Thermogravimetric analysis (TGA) of (a) CAM (); (b) GSF (); (c) BCP (); (d) GPF ().
Figure 2. Thermogravimetric analysis (TGA) of (a) CAM (); (b) GSF (); (c) BCP (); (d) GPF ().
Sustainability 15 14760 g002
Figure 3. Differential Scanning Calorimetry (DSC) of (a) CAM (); (b) GSF (); (c) BCP (); (d) GPF ().
Figure 3. Differential Scanning Calorimetry (DSC) of (a) CAM (); (b) GSF (); (c) BCP (); (d) GPF ().
Sustainability 15 14760 g003
Figure 4. Maximization of the mean for Response A: initial concentration of TiO2 (); additive type (); additive concentration (); horn intensity (); mean 41.03 (˗); standard error ± 2 × 0.1626 (---).
Figure 4. Maximization of the mean for Response A: initial concentration of TiO2 (); additive type (); additive concentration (); horn intensity (); mean 41.03 (˗); standard error ± 2 × 0.1626 (---).
Sustainability 15 14760 g004
Figure 5. TiO2 photoactivation and As(III) oxidation mechanisms.
Figure 5. TiO2 photoactivation and As(III) oxidation mechanisms.
Sustainability 15 14760 g005
Figure 6. Reuse performance of TiO2 immobilized structure in the As(III) oxidation ([As (III)] = 0.27 mM) by UVA/GPF-TiO2 annular reactor system at pH 8.0 and 20 °C; 1st cycle (); 2nd cycle (); 3rd cycle (); 4th cycle (); 5th cycle (); 6th cycle ().
Figure 6. Reuse performance of TiO2 immobilized structure in the As(III) oxidation ([As (III)] = 0.27 mM) by UVA/GPF-TiO2 annular reactor system at pH 8.0 and 20 °C; 1st cycle (); 2nd cycle (); 3rd cycle (); 4th cycle (); 5th cycle (); 6th cycle ().
Sustainability 15 14760 g006
Figure 7. Top view SEM images of: (a) sample Y; (b) GPF-TiO2; (c) GPF-TiO2 after 6 reuse cycles of As(III) oxidation in the annular reactor system. EDS spectra of: (d) sample Y; (e) GPF-TiO2; (f) GPF-TiO2 after 6 reuse cycles of As(III) oxidation in the annular reactor system.
Figure 7. Top view SEM images of: (a) sample Y; (b) GPF-TiO2; (c) GPF-TiO2 after 6 reuse cycles of As(III) oxidation in the annular reactor system. EDS spectra of: (d) sample Y; (e) GPF-TiO2; (f) GPF-TiO2 after 6 reuse cycles of As(III) oxidation in the annular reactor system.
Sustainability 15 14760 g007
Table 1. Factors, levels, and response utilized in the design of experiments.
Table 1. Factors, levels, and response utilized in the design of experiments.
Influencing Parameters
FactorLevels
123
Initial TiO2 concentration (wt.%)0.11.05.0
AdditiveTriton X (TX)Citric acid (CA)Glutaraldehyde (GL)
Additive concentration (wt.%)135
Ultrasonic horn (intensity/time)0 *30% (5 min)70% (5 min)
Response
% of TiO2 immobilized (xi)Response A
% As(V) after reaction (gi) A i = x i + g i + z i + k i + h i 5 (1)
% de TiO2 immobilized after reaction (zi)
% As(V) after reuse (ki)
% de TiO2 immobilized after reuse reaction (hi)
* Without ultrasonic treatment.
Table 2. Design of experiments according to Taguchi L9 orthogonal array.
Table 2. Design of experiments according to Taguchi L9 orthogonal array.
TrailsInitial TiO2 Concentration (wt.%)AdditiveAdditive Concentration (wt.%)Ultrasonic Horn (Intensity %)
10.1 TX1 0
20.1 CA3 30
30.1 GL5 70
41.0 TX3 70
51.0 CA5 0
61.0 GL1 30
75.0TX5 30
85.0CA1 70
95.0GL3 0
Table 3. Mean values of normalized experimental results and the respective factor A.
Table 3. Mean values of normalized experimental results and the respective factor A.
Trials% of TiO2 Immobilized% As(V) after Reaction% de TiO2 Immobilized after Reaction% As(V) after Reuse% de TiO2 Immobilized after Reuse ReactionResponse A
xigizikihiAi
18.661.9699.801.0699.4142.18
24.781.6899.460.8198.4141.03
314.018.5379.004.5875.6736.35
41.821.5999.862.2399.6641.03
50.511.1399.171.3298.8240.19
60.5710.5596.278.3494.9842.14
70.639.0899.656.5499.2843.03
80.234.2493.205.6091.5238.96
90.4814.5598.0614.6594.7044.37
Table 4. ANOVA. (SQ) sum of squares; (df) degrees of freedom; (SMQ) sum of the mean squares; (F) F-test.
Table 4. ANOVA. (SQ) sum of squares; (df) degrees of freedom; (SMQ) sum of the mean squares; (F) F-test.
EffectSQdfSMQFPorcentage ContributionRank
Initial TiO2 concentration (wt.%)15.514527.757397.751417.393
Additive type12.321926.161077.636013.814
Additive concentration (wt.%)15.712027.856098.995417.612
Ultrasonic horn (% intensity)45.6707222.83540287.754351.191
Residual0.714290.07936
Table 5. Results of the confirmation together with the mean of the optimal Factor A predicted by the Statistica software.
Table 5. Results of the confirmation together with the mean of the optimal Factor A predicted by the Statistica software.
Confirmation ExperimentLevels
Initial TiO2 concentration (w/w)5
AdditiveTX
Additive concentration (w/w)3
Ultrasonic horn (% intensity)WH
Output factorsResponse
12
% of TiO2 immobilized0.750.69
% As(V) after reaction12.5513.25
% de TiO2 immobilized after reaction99.9599.90
% As(V) after reuse12.1512.25
% de TiO2 immobilized after reaction99.9099.85
Factor A (%)45.0645.19
Factor A (%) mean45.13
Results predicted by Statistica
Factor A45.49
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

Scherer Filho, J.A.; Marinho, B.A.; Vignola, F.; Mazur, L.P.; González, S.Y.G.; da Silva, A.; Ulson de Souza, A.A.; Guelli Ulson de Souza, S.M.A. Development of a Novel 3D Highly Porous Structure for TiO2 Immobilization and Application in As(III) Oxidation. Sustainability 2023, 15, 14760. https://doi.org/10.3390/su152014760

AMA Style

Scherer Filho JA, Marinho BA, Vignola F, Mazur LP, González SYG, da Silva A, Ulson de Souza AA, Guelli Ulson de Souza SMA. Development of a Novel 3D Highly Porous Structure for TiO2 Immobilization and Application in As(III) Oxidation. Sustainability. 2023; 15(20):14760. https://doi.org/10.3390/su152014760

Chicago/Turabian Style

Scherer Filho, Julio A., Belisa A. Marinho, Fabiola Vignola, Luciana P. Mazur, Sergio Y. G. González, Adriano da Silva, Antônio Augusto Ulson de Souza, and Selene M. A. Guelli Ulson de Souza. 2023. "Development of a Novel 3D Highly Porous Structure for TiO2 Immobilization and Application in As(III) Oxidation" Sustainability 15, no. 20: 14760. https://doi.org/10.3390/su152014760

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