Next Article in Journal
Interfacial Electron Redistribution of FeCo2S4/N-S-rGO Boosting Bifunctional Oxygen Electrocatalysis Performance
Next Article in Special Issue
Towards the Sustainable Production of Ultra-Low-Sulfur Fuels through Photocatalytic Oxidation
Previous Article in Journal
Efficient and Eco-Friendly Perspectives for C-H Arylation of Benzothiazole Utilizing Pd Nanoparticle-Decorated Chitosan
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Biogenic Silica Behavior in the Incorporation of Mesoporous Anatase TiO2 for Excellent Photocatalytic Mineralization of Sodium Diclofenac

by
Christian Brice Dantio Nguela
1,2,
Ngomo Horace Manga
1,
Clément Marchal
2,
Aimé Victoire Abega
2,3,
Ndi Julius Nsami
1 and
Didier Robert
2,*
1
Laboratory of Applied Physical and Analytical Chemistry, Department of Inorganic Chemistry, Faculty of Science, University of Yaounde1, Yaoundé P.O. Box 812, Cameroon
2
Institut de Chimie et Procédés pour l’Energie, l’Environnement et la Santé, UMR-CNRS 7515, Université de Strasbourg, 25 rue Becquerel, 67087 Strasbourg, France
3
Department of Chemistry, Faculty of Sciences, University of Douala, Douala P.O. Box 24157, Cameroon
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(9), 1001; https://doi.org/10.3390/catal12091001
Submission received: 5 August 2022 / Revised: 29 August 2022 / Accepted: 30 August 2022 / Published: 5 September 2022

Abstract

:
TiO2/SiO2 composites were synthesized via a simple sol gel method by surface reduction of Ti4+ ions to Ti3+ using titanium isopropoxide as a TiO2 precursor and rice husks (RHA) as a SiO2 source. The silica content and calcination temperature of the materials were evaluated. Thermal, crystallographic and physicochemical aspects suggest that biogenic silica (SiO2) can improve the thermal stability of the anatase phase of TiO2, when the SiO2 content reaches 20%. The N2 adsorption-desorption isotherms showed that the SiO2-modified samples have uniform pore diameters and a large specific surface area. The XPS analysis showed the surface reduction of Ti4+ ions to Ti3+ within the TiO2 network via oxygen vacancies after SiO2 introduction, which is beneficial for the photocatalytic reaction. Photocatalytic degradation of sodium diclofenac (SDFC) shows that TiO2/SiO2 composites have better activity compared to commercial P25. Mesoporous TiO2 composite modified with 20 wt% SiO2 showed better photocatalytic mineralization than P25 (83.7% after 2 h instead of 57.3% for P25). The excellent photocatalytic mineralization of the photocatalysts can be attributed to the high anatase crystallinity exhibited by XRD analysis, high specific surface area, surface hydroxyl groups, and the creation of oxygen vacancy, as well as the presence of Ti3+ ions.

Graphical Abstract

1. Introduction

Pharmaceuticals and personal care products (PPCPs) present in wastewater have potential negative impacts on humans and ecosystems, even when present in trace amounts [1]. Among these contaminants, sodium diclofenac (SDCF) is a non-steroidal anti-inflammatory drug commonly used as an analgesic, anti-arthritic, and anti-rheumatic drug [2]. Its biodegradation and natural attenuation are quite limited [3]. Among the different processes implemented for the removal of SDCF, heterogeneous photocatalysis is considered an effective method to achieve its mineralization [4,5]. During the last two decades, heterogeneous photocatalysis has been demonstrated to be a “green” and effective method for water purification and disinfection [6]. This process is based on the absorption of light by a semiconductor, usually TiO2, to produce electron–hole pairs, reacting at the interface to produce oxidizing species such as hydroxyl or superoxide radical. One of the main advantages of photocatalysis is the capacity to mineralize, without selectivity, the majority of organic compounds.
TiO2 is one of the most widely used photocatalysts, due to its high activity, chemical stability, low toxicity, strong oxidizing power to decompose organic pollutants, and environmental friendliness [7]. It has been widely studied for water purification, hydrogen production, self-cleaning surfaces, etc. [8]. However, TiO2 anatase powders, although having high surface areas, are not stable at high temperatures and easily lose their quantum efficiency and adsorption capacity, by phase transformation and crystallite growth [9], which considerably limits their photocatalytic efficiency [10]. Recently, our research group synthesized the carbon nanotube/TiO2 composite nanomaterial in which Ti3+ ions and oxygen vacancies appeared within the TiO2 band gap as a sub-band (sub-level). This composite showed a consistent improvement in the degradation of methyl orange under visible light irradiation [11]. Another strategy to solve this problem is to combine TiO2 with SiO2 [12,13]. SiO2 is a very important inorganic material used for a wide range of commercial applications such as molecular sieves, catalysts, and in biomedical and electrical applications. Biogenic SiO2 has been studied as potential support due to its high specific surface area, porous structure, thermal stability, ability to effectively decrease aggregation, and transparency in the UV light range [14,15,16]. This biogenic silica could be a by-product of rice production, especially rice husks, with a relative humidity of approximately 20–25% by weight of the total dry weight of the rice [16,17]. According to Ndindeng et al., there are 5 million tons of rice produced per year by sub-Saharan African countries, leading to high waste rice husks discharged into the environment [18]. Due to its low mechanical strength, low nutritional value and low bulk density, its use is limited [19]. However, rice husks contain a large amount of silica, approximately 98% of the total inorganic compounds [20], which can be used as raw material to produce value-added silicon-based materials with interesting texture and morphology. The implementation of such substances from rice husks is very interesting for the fact that they come from biomass, which is a renewable energy source [21]. Over the past decade, this area of research has progressed considerably and expanded, driven by the global emphasis on sustainable and renewable resources. As a chemically inert material, porous silica plays an important role in various applications, including in adsorption and as catalyst support. It can be artificially synthesized with a certain degree of control in its nano-morphology [22], and this usually involves a complicated process. Most of the time, a large amount of amorphous silica is produced by multi-step thermal reduction of raw silica at high temperatures and high pressures under extremely acidic and energy-intensive conditions, which have a high impact on the environment [18]. Therefore, it is important to develop an affordable, sustainable, and environmentally friendly technique for the manufacture of porous silica to meet the growing demand for its widespread applications. Until now, many works have been interested in the synthesis of TiO2/SiO2 composites with different methods such as thermal and solvothermal hydrolysis [23], the sol-gel method [24], the hydrothermal method [25], and spray coating [26]. However, most of these methods use tetraethyl orthosilicate (TEOS) as a silica precursor and require several steps in the syntheses, high temperatures and pressures, and are also expensive.
The present work is based on the synthesis of TiO2 nanoparticles homogeneously and uniformly dispersed on the surface of biogenic silica nanoparticles by a simple sol-gel method, in which the surface reduction of Ti4+ ions to Ti3+ occurred to generate electronic states below the conduction band of TiO2. The Ti3+ ions act as a sub-band gap to reduce the recombination rate of photogenerated electrons/holes pairs. The nanomaterials were characterized using thermogravimetric analysis, X-ray diffraction (XRD), scanning electron microscopy (SEM), UV-Vis diffuse reflectance spectroscopy, X-ray photoelectron spectroscopy (XPS), Brunauer–Emmett–Teller Analysis (BET), Fourier transform infrared (FT-IR), and energy dispersive X-ray analysis (EDX). The photocatalytic performance of the synthesized nanomaterials on the degradation of sodium diclofenac as a model pollutant in an aqueous solution under UV-A irradiation is also reported.

2. Results and Discussion

2.1. Thermal Analysis

Thermogravimetric (TG) and differential scanning calorimetric (DSC) analysis data of TiO2 gels prepared without SiO2 and with 20% SiO2 are shown in Figure 1. The DSC curve shows that the TG exhibit three similar thermal accidents, although they are different from each other in terms of weight changes. The first loss at approximately 100 °C corresponds to the removal of adsorbed water, on pure TiO2 and the 20% T/S composite, respectively. The second loss with a large exothermic peak at 245°C corresponds to the partial elimination of hydroxyl groups. It can be seen that the weight loss of T/S 20% is higher than that of TiO2. This may be due to the presence of more hydroxyl groups introduced by SiO2 that are easier to eliminate at 245 °C [27]. Finally, the last loss corresponds to the transformation of amorphous TiO2 to crystalline TiO2 anatase observed at 381.8 °C for pure TiO2 and increases to 386.1 °C for the T/S 20% composite in agreement with the X-ray diffraction results (see next section).

2.2. Structural Analysis

It can be seen (Figure 2A) that the silica nanoparticles show an amorphous phase with a wide diffraction peak between 15 and 30°, which is in good agreement with the literature [20]. The DRX pattern of pure TiO2 shows the presence of anatase and rutile phases for TiO2 calcined at 650 °C. The diffraction peaks at 25.3°, 38.83°, 48.11°, 53.95°, 62.79°, 69.0°, and 70.46° correspond to the (101), (004), (200), (211), (204), (220), and (215) planes of the anatase, respectively (ICSD:033547). The diffraction peaks at 27.54°, 36.17°, 39.28°, 41.32°, 44.14°, 54.54°, 56.66°, 64.15°, and 69.0° correspond to the (110), (101), (200), (111), (211), (220), (002), (310), and (301) planes of the rutile, respectively. The diagram of TiO2/SiO2 composites shows that silica nanoparticles play a major role in the thermal stability of pure TiO2. When the silica content is less or equal to 5%, a weak rutile phase peak is observed, which disappears by increasing the silica content to 20%. This shows that silica can delay the transformation of the anatase phase to rutile and reduce the size of TiO2 crystallites (Figure 3). The decrease in crystallite size can be attributed to the good dispersion of TiO2 on the silica surface, which prevents any possible agglomeration of the nanoparticles [28].
In Figure 2B, it can be seen that the main phase of the T/S 20% composite calcined at 350 °C is pure anatase. When the temperature increases, the anatase peaks become progressively strong and narrow, and the anatase crystallite size increases from 6.1 nm to 21.03 nm at 750 °C (Figure 3) with the appearance of rutile peaks. This is confirmed by a change in lattice parameters ‘a‘ and ‘c’ of the composites (Table 1), causing a decrease in microstrain and dislocation density as the temperature increases. This implies that the transition temperature of the anatase to rutile is between 650 and 750 °C. The microstrains (δ) and dislocation density (ε) were obtained by Equations (1) and (2), while the average crystallite size (D) of the anatase nanoparticles was obtained by the Debye–Scherrer equation (Equation (3)) using the peak attributed to the anatase crystal planes (101) by the following relations:
D = k λ β c o s θ
ε = β 4 t a n θ
δ = 1 D 2
where λ is the X-ray wavelength, θ is a specific angle, β is the width at half maximum (FWHM) for the anatase peak (101), and k is the constant depending on the shape of the crystallites (k is 0.9 when the particles are spherical).
The results in Table 1 and Figure 2 clearly show that the average crystallite size increases with temperature and decreases with silica content. This clearly indicates that the presence of SiO2 allows good dispersion of TiO2 particles on its surface (see SEM in Section SEM and EDX Analysis), thus giving photocatalytic nanomaterials with a large specific surface area.

2.3. Mesoporous Structure

Figure 4 shows the adsorption–desorption isotherms of N2 for SiO2, TiO2, and T/S-X calcinated at 650 °C. The type IV isotherms show an H3-type hysteresis loop corresponding to capillary condensation on the mesopores. This confirms the mesoporous characteristic of the materials, which possesses significant importance in the adsorption of organic molecules. Furthermore, this result suggests that SiO2 effectively preserves the TiO2 mesostructures, especially when SiO2 content reaches 20%. Moreover, the pore volume increases with the SiO2 content of composite materials, whereas the pore diameter decreases. These results suggest an increase in surface availability of the target pollutant, leading to an improvement in adsorption capacity. The effect of silica on the specific surface area shows that the surface area of the photocatalysts composite increases with SiO2 content, which further demonstrates the importance of SiO2 in TiO2 dispersion (Table 2).

SEM and EDX Analysis

It can be seen in Figure 5 that SiO2 is similar to highly porous cloud-like platelets stacked more or less densely (Figure 5A). Figure 5B shows that pure TiO2 shows obvious agglomerated particles with denser and larger particles, which leads to a decrease in surface area in agreement with the BET result. In contrast, the 20% T/S composite has a relatively uniform TiO2 nanoparticle distribution with less aggregation and a porous microstructure (Figure 5C). This material exhibits the characteristics of a large 2D porous plate covered with a small bead. It is noted that said platelets are bonded in-plane with a strong decrease in platelet stacking. This result indicates that silica can improve the regularity of mesopores, inhibiting the growth of nanocrystallites and agglomeration of TiO2 nanoparticles. The chemical composition of T/S20% was determined by EDX analysis (Figure 5D). The results show that the main elements on its surface area are O, Si, and Ti. The Si element confirms the existence of SiO2 in the structure of TiO2. The mass percentages of the different elements are 37%, 1.5%, and 61.5%, respectively, in the composite (see composition in Figure 5D) for the elements O, Si, and Ti.

2.4. Surface Property

The IR spectrum of SiO2 is given in Figure 6A, and the bands at 1063 and 799 cm−1 correspond to the characteristic asymmetric and symmetric stretching vibrations of the Si-O-Si bond, respectively. The bands at 958 and 450 cm−1 are attributed to the stretching vibrations of Si-O-H and Si-O bonds, while the bands at 1632 and 3414 cm−1 correspond, respectively, to the bending vibration of the Si-O-Si bond and stretching vibrations of the O-H bond adsorbed water [29].
Comparing the IR spectrum (Figure 6A) of SiO2 to the IR spectra (Figure 6B,C) of samples with different SiO2 contents and varying the calcination temperature, a weak band at approximately 3650 cm−1 is observed, which corresponds to the O-H bonds of the coordination water. The broad bands at approximately 1630 and 3400 cm−1 correspond to the deformation and stretching vibrations, respectively, of the O-H bond of adsorbed, bound, or free water. The characteristic bands at 2985 and 2894 cm−1 correspond to the stretching vibrations of the -CH2 bond. The bands at approximately 1070 cm−1 are attributed to the asymmetric vibration of Si-O-Si bonds in SiO2 [30]. It can be noted that the number of hydroxyl bonds and the surface Si-O-Si bonds increases as the SiO2 content increases (Figure 6B), and the O-H bonds in the T/S 20% sample remain constant until the calcination temperature reaches 650 °C. These results suggest that silica can effectively limit the loss of surface hydroxyl groups during calcination, thus the hydroxyl groups on the surface of the samples can be preserved, which is beneficial for the photocatalytic reaction in that these groups can increase molecule adsorption via coulombic attractions. This can be confirmed by the band at approximately 3400 cm−1, which becomes broader. The decrease in the intensity of hydroxyl groups with increasing temperature (Figure 6C), can be caused by the aggregation of TiO2 nanoparticles, which release the surface hydroxyl groups from SiO2 and then become easily volatile at this temperature, according to the DRX analysis.

2.5. Surface Composition and Element Surface Oxidation State Analysis

To determine the surface composition as well as the oxidation state of the constituent elements of each material, XPS analysis was performed. This analysis clarified the existence, or not, of the Ti-O-Si bond in the SiO2-modified TiO2 mesoporous. The XPS analysis of the T/S 20% composite calcined at 650 °C (Figure 7A) shows the presence of three main peaks, notably Ti 2p (456.33 eV), O 1s (528.84 eV), and Si 2p (102.84 eV), confirming that the TiO2 nanoparticles are well coated on the surface of the silica nanoparticles.
Figure 7B shows the deconvolution of the Ti 2p region, where the bands centered at 463.50 and 458.00 eV are assigned to Ti4+ 2p1/2 and Ti4+ 2p3/2, respectively. On the other hand, the peaks at 461.03 and 455.50 eV are attributed to Ti3+ 2p1/2 and Ti3+ 2p3/2, respectively, which may illustrate that Ti3+ ions are introduced due to the surface reduction of Ti4+ ions in the TiO2/SiO2 composite at high temperatures. Figure 7C shows the deconvolution of the O 1s peak, which can be divided into four peaks at 533.52, 531.00, 529.54, and 526.50 eV, which correspond to Si-O-Si, Ti-OH, and O22− (O-Ti-O) bonds and oxygen vacancies, respectively. On this last point, it should be noted that with the increase in temperature, vacancies were created within TiO2 according to the following equation:
Ti4+–O–Ti4+ →→→ Ti3+–□–Ti3+ + ½ O2
Figure 7D displays the Si 2p peak observed at 103.4 eV, indicating the existence of Si-O-Si bonding in the composite. Therefore, the results obtained from the XPS spectra are consistent with the infrared spectroscopy analysis.

2.6. Photocatalytic Test for SDCF Mineralization

The photocatalytic efficiency of TiO2/SiO2 composites, TiO2–P25, and pure TiO2 was evaluated by degradation and mineralization of sodium diclofenac in an aqueous solution under UV-A radiation (60 W/m2, λmax = 368 nm). It can be seen from Figure 8A that the composite containing 20% of SiO2 shows higher photocatalytic activity and also better mineralization than TiO2–P25 (83.68% mineralization efficiency while TiO2–P25 has only 57.33% after 120 min). These results reveal that silica can effectively enhance the photocatalytic activity of TiO2 even at high temperatures. It can be seen that the calcination temperature has a significant effect on photocatalytic activity. As the calcination temperature increases up to 650 °C, the 20% T/S composite always shows better photocatalytic activity and photocatalytic mineralization (Figure 8C) due to its large surface area, high crystallinity, and the presence of abundant hydroxyl groups on its surface. After calcination at 750 °C, the photocatalytic activity decreases considerably due to the appearance of the rutile phase which is less active than the anatase phase. On the other hand, the increase in the size of the TiO2 nanoparticles leads to a decrease in the surface contact with the pollutant and surface hydroxyl groups, which causes a decrease in the interactions with the target molecule. However, it still shows better photocatalytic mineralization than TiO2–P25, with 74.56% mineralization after 120 min.
Good photocatalytic efficiency of the TiO2/SiO2 composites can be attributed to several factors. First, the size of the TiO2 nanoparticles within the composites was smaller than that of pure TiO2 calcined at 650 °C. The decrease in particle size can shorten the time for the electron/hole pairs to reach the sample surface, which effectively reduces recombination charges and increases the rate of reduction or oxidation, thus improving the overall photocatalytic performance of the composite nanomaterial [31]. In addition, the absence of the rutile phase in the composite at 650 °C promotes high photocatalytic activity.
Second, it is well known that the photocatalytic reaction is a surface reaction; thus, a large surface area is generally beneficial for photocatalytic reactions as it can provide more adsorption sites and photocatalytic reaction centers [16].
Third, TiO2 combined with SiO2 could effectively retain its mesoporous structure even at high calcination temperatures, which promotes the good diffusion of organic molecules into its structure, and results in high photocatalytic activity. Moreover, it has been reported that organic molecules have easier adsorption on the surface of TiO2 in the presence of silica [32].
Finally, the composites have more surface hydroxyl groups compared to pure TiO2 as FTIR analysis confirmed (see Section 2.4). It is widely believed that surface hydroxyl groups can capture photo-induced holes and produce active hydroxyl radicals and prevent electron–hole recombination at the same time [16].

3. Materials and Methods

3.1. Materials

All reagents used were of analytical grade and used without further purification: Titanetetraisopropoxide Ti(OC3H7)4 (Aldrich, 97%), ethanol (SdS-France, 99.9%), H2SO4 (prolabo, 98%), NaOH (prolabo 98%), HCl (Prolabo, 37%), sodium Diclofenac (Aldrich 98%), and TiO2-P25 (average size 20 nm, purity 97%, surface area 50 m²·g−1 and 80% anatase, 20% rutile, Evonik industries). The rice husk samples used in this study were collected in the northwest Cameroon region in the locality of Ndop (6°00′00″ North, 10°25′00″ East), more precisely in the Ngo-ketunjia department.

3.2. Extraction of Nanosilica from Rice Husks

First, 150 g of rice husks previously washed with distilled water were refluxed with 500 mL (0.5 M) HCl at 70 °C for 2 h to remove metal impurities. After the complete reaction, the rice husks were washed several times with distilled water to remove HCl, and dried at 110 °C for 24 h. Then, the treated rice husks were calcinated at 800 °C with a heating rate of 10 °C/min for 2 h, to remove the carbonaceous material. White ash (silica) of high purity was obtained. The resulting ash was magnetically stirred with a 250 mL solution of 1 M NaOH for 2 h and filtered. The obtained solution was neutralized with 0.5 M sulfuric acid until the pH was equal to 2, after which a nanosilica gel was formed. This gel was then washed with distilled water until pH 7 was reached and then dried in an oven at 70 °C for 48 h to form a silica nanoparticle powder as shown in the following reactions:
White ash + 2 NaOH → → Na2SiO3 + H2O
Na2SiO3 + H2SO4 → → SiO2 + Na2SO4 + H2O

Synthesis of TiO2/SiO2 Nanophotocatalysts

TiO2/SiO2 composites were synthesized using the sol-gel method for the reaction between titanium dioxide and silica. First, 15 mL of anhydrous ethanol was mixed with a certain amount of nanosilica (solution A). Meanwhile, 5 mL of Ti(OC3H7)4 was dissolved in 15 mL of anhydrous ethanol and 30 mL of distilled water under stirring (solution B). Then, solution A was added dropwise into solution B under vigorous stirring at room temperature. The weight percentage of silica added varied from 5 to 50% relative to TiO2. The reaction solution was kept under stirring for 5 h and then aged at 25 °C for 24 h. Finally, the resulting gel was dried at 110 °C for 24 h, calcined at a certain temperature in air for 2 h with a heating rate of 5 °C/min, and labeled TS x–T, where x represented the theoretical molar percentage of silica in the catalyst and T indicated the calcination temperature (in the range of 350–750 °C).

3.3. Characterizations

TG/DSC analyses were performed in an air atmosphere with a heating rate of 10 °C/min from room temperature to 800 °C on a LABSYS evo instrument. X-ray diffraction patterns were recorded with a Ragaku Miniflex II X-ray diffractometer (Cu Kα radiation, λ = 0.154056 nm). Crystallite size was estimated by applying the Debye-Scherrer equation to the (101) anatase peak. SEM analyses were performed on an ion beam scanning electron microscope (JEOL 6700F equipped with a field emission gun with an extract potential of 2.5 kV) combined with an EDX analyzer. The FT-IR spectra of the materials were recorded at room temperature in the wavelength range of 400 to 4000 cm−1 with a Thermo Fisher Nicolet S10 spectrophotometer. Textural properties were obtained using an ASAP-2420 analyzer from Micromeritics. The specific surface area (SBET) was calculated from the Brunauer–Emmett–Teller (BET) equation from the physisorption of N2 at 77 K. X-ray photoelectron spectroscopy (XPS) measurements were performed on an ultra-high vacuum (UVH) spectrometer equipped with a VSW Class WA hemispheric electron analyzer. The X-ray source used was a double Al Kᾳ X-ray (1486.6 eV) aluminum anode as incident radiation. The general high-resolution spectra were recorded in constant energy mode (100 and 20 eV, respectively). In order to correct for the shift in binding energy due to electrostatic charge, the internal reference used is the C1s peak at 284.9 eV. characteristic of sp2 hybridized C. The background is subtracted according to the Shirley method.

3.4. Photocatalytic Measurement

First, 100 mg of the photocatalyst (0.5 g/L) was added to the sodium diclofenac solution (200 mL, 10 mg/L) in a 400 mL Pyrex beaker. The suspension was stirred in the dark for 30 min to reach the adsorption–desorption equilibrium. Then, the solution was irradiated with a low-power UV-A lamp (60 W/m2, λmax = 368 nm). The distance between the light source and the suspension was 20 cm. At regular irradiation intervals, approximately 10 mL of suspension was taken with a syringe, then filtered with 0.22 μm filters and analyzed by a UV-vis spectrophotometer of the brand ZUZI Spectrophometer model 4211/50 at 276 nm, and the Total Organic Carbon concentration (TOC) was evaluated with a Shimadzu model TOC-L apparatus.

4. Conclusions

A simple sol-gel method was developed for the synthesis of anatase TiO2/SiO2 composites via surface reduction of Ti4+ ions to Ti3+ using rice husks as the source of the nanosilica. The results show that the biogenic silica increases the specific surface area of the materials and acts as good support that provides enough space for better dispersion of TiO2 on the surface to avoid any possible agglomeration of the nanoparticles. The composites showed clear thermal stability and significant transformation of the anatase phase to rutile between 650 and 750 °C. In addition, the TiO2/SiO2 20% composite showed anatase thermal stability up to 650 °C and excellent photocatalytic mineralization compared to P25 (83.68% after 2 h instead of 57.33% for P25).

Author Contributions

Conceptualization, D.R. and C.B.D.N.; methodology, C.B.D.N.; software, C.M.; validation, N.J.N., formal analysis, C.M.; investigation, C.B.D.N.; data curation, A.V.A.; writing—original draft preparation, C.B.D.N.; writing—review and editing, D.R.; supervision, D.R. and N.H.M.; Project administration, D.R.; funding acquisition, D.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Campus France “Eiffel” grant number 978617A.

Acknowledgments

The authors of this article would like to thank Campus France for the grant “Eiffel” awarded to Christian Brice DANTIO NGUELA.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Papageorgiou, M.; Kosma, C.; Lambropoulou, D. Seasonal occurrence, removal, mass loading and environmental risk assessment of 55 pharmaceuticals and personal care products in a municipal wastewater treatment plant in Central Greece. Sci. Total Environ. 2016, 543, 547–569. [Google Scholar] [CrossRef] [PubMed]
  2. Zhang, Y.; Geißen, S.-U.; Gal, C. Carbamazepine and diclofenac: Removal in wastewater treatment plants and occurrence in water bodies. Chemosphere 2008, 73, 1151–1161. [Google Scholar] [CrossRef] [PubMed]
  3. Pal, A.; Gin, K.Y.-H.; Lin, A.Y.-C.; Reinhard, M. Impacts of emerging organic contaminants on freshwater resources: Review of recent occurrences, sources, fate and effects. Sci. Total Environ. 2010, 408, 6062–6069. [Google Scholar] [CrossRef] [PubMed]
  4. An, J.; Zhou, Q. Degradation of some typical pharmaceuticals and personal care products with copper-plating iron doped Cu2O under visible light irradiation. J. Environ. Sci. 2012, 24, 827–833. [Google Scholar] [CrossRef]
  5. Rosales, E.; Diaz, S.; Pazos, M.; Sanromán, M.A. Comprehensive strategy for the degradation of anti-inflammatory drug diclofenac by different advanced oxidation processes. Sep. Purif. Technol. 2019, 208, 130–141. [Google Scholar] [CrossRef]
  6. Daghrir, R.; Drogui, P.; Robert, D. Modified TiO2 for Environmental Photocatalytic Applications: A Review. Ind. Eng. Chem. Res. 2013, 52, 3581–3599. [Google Scholar] [CrossRef]
  7. Nakata, K.; Fujishima, A. TiO2 photocatalysis: Design and applications. J. Photochem. Photobiol. C Photochem. Rev. 2012, 13, 169–189. [Google Scholar] [CrossRef]
  8. Wang, Y.; Su, Y.; Qiao, L.; Liu, L.; Su, Q.; Zhu, C.; Liu, X. Synthesis of one-dimensional TiO2/V2O5 branched heterostructures and their visible light photocatalytic activity towards Rhodamine B. Nanotechnology 2011, 22, 225702. [Google Scholar] [CrossRef]
  9. Lee, J.-W.; Kong, S.; Kim, W.-S.; Kim, J. Preparation and characterization of SiO2/TiO2 core-shell particles with controlled shell thickness. Mater. Chem. Phys. 2007, 106, 39–44. [Google Scholar] [CrossRef]
  10. Wu, L.; Zhou, Y.; Nie, W.; Song, L.; Chen, P. Synthesis of highly monodispersed teardrop-shaped core–shell SiO2/TiO2 nanoparticles and their photocatalytic activities. Appl. Surf. Sci. 2015, 351, 320–326. [Google Scholar] [CrossRef]
  11. Abega, A.V.; Ngomo, H.M.; Nongwe, I.; Mukaya, H.E.; Sone, P.-M.A.K.; Mbianda, X.Y. Easy and convenient synthesis of CNT/TiO2 nanohybrid by in-surface oxidation of Ti3+ ions and application in the photocatalytic degradation of organic contaminants in water. Synth. Met. 2019, 251, 1–14. [Google Scholar] [CrossRef]
  12. Kapridaki, C.; Pinho, L.; Mosquera, M.J.; Maravelaki-Kalaitzaki, P. Producing photoactive, transparent and hydrophobic SiO2-crystalline TiO2 nanocomposites at ambient conditions with application as self-cleaning coatings. Appl. Catal. B Environ. 2014, 156–157, 416–427. [Google Scholar] [CrossRef]
  13. Wang, J.; Sun, S.; Ding, H.; Chen, W.; Liang, Y. Preparation of a composite photocatalyst with enhanced photocatalytic activity: Smaller TiO2 carried on SiO2 microsphere. Appl. Surf. Sci. 2019, 493, 146–156. [Google Scholar] [CrossRef]
  14. Ren, C.; Qiu, W.; Chen, Y. Physicochemical properties and photocatalytic activity of the TiO2/SiO2 prepared by precipitation method. Sep. Purif. Technol. 2013, 107, 264–272. [Google Scholar] [CrossRef]
  15. Tang, X.; Feng, Q.; Liu, K.; Tan, Y. Synthesis and characterization of a novel nanofibrous TiO2/SiO2 composite with enhanced photocatalytic activity. Mater. Lett. 2016, 183, 175–178. [Google Scholar] [CrossRef]
  16. He, C.; Tian, B.; Zhang, J. Thermally stable SiO2-doped mesoporous anatase TiO2 with large surface area and excellent photocatalytic activity. J. Colloid Interface Sci. 2010, 344, 382–389. [Google Scholar] [CrossRef] [PubMed]
  17. Yu, J.; Zhang, J.; He, J.; Liu, Z.; Yu, Z. Combinations of mild physical or chemical pretreatment with biological pretreatment for enzymatic hydrolysis of rice hull. Bioresour. Technol. 2009, 100, 903–908. [Google Scholar] [CrossRef]
  18. Ndindeng, S.A.; Wopereis, M.; Sanyang, S.; Futakuchi, K. Evaluation of fan-assisted rice husk fuelled gasifier cookstoves for application in sub-Sahara Africa. Renew. Energy 2019, 139, 924–935. [Google Scholar] [CrossRef]
  19. Alosaimi, E.H.; Alsohaimi, I.H.; Dahan, T.E.; Chen, Q.; Younes, A.A.; El-Gammal, B.; Melhi, S. Photocatalytic Degradation of Methylene Blue and Antibacterial Activity of Mesoporous TiO2-SBA-15 Nanocomposite Based on Rice Husk. Adsorpt. Sci. Technol. 2021, 2021, 9290644. [Google Scholar] [CrossRef]
  20. Efremova, S.V. Rice hull as a renewable raw material and its processing routes. Russ. J. Gen. Chem. 2012, 82, 999–1005. [Google Scholar] [CrossRef]
  21. Sonobe, T.; Worasuwannarak, N. Kinetic analyses of biomass pyrolysis using the distributed activation energy model. Fuel 2008, 87, 414–421. [Google Scholar] [CrossRef]
  22. Habeeb, G.A.; Mahmud, H.B. Study on properties of rice husk ash and its use as cement replacement material. Mater. Res. 2010, 13, 185–190. [Google Scholar] [CrossRef]
  23. Hong, S.-S.; Lee, M.S.; Park, S.S.; Lee, G.-D. Synthesis of nanosized TiO2/SiO2 particles in the microemulsion and their photocatalytic activity on the decomposition of p-nitrophenol. Catal. Today 2003, 87, 99–105. [Google Scholar] [CrossRef]
  24. Mahanta, U.; Khandelwal, M.; Deshpande, A.S. TiO2@SiO2 nanoparticles for methylene blue removal and photocatalytic degradation under natural sunlight and low-power UV light. Appl. Surf. Sci. 2022, 576, 151745. [Google Scholar] [CrossRef]
  25. Kominami, H.; Yukishita, K.; Kimura, T.; Matsubara, M.; Hashimoto, K.; Kera, Y.; Ohtani, B. Direct Sol-vothermal Formation of Nanocrystalline TiO2 on Porous SiO2 Adsorbent and Photocatalytic Removal of Nitrogen Oxides in Air over TiO2–SiO2 Composites. Top. Catal. 2008, 47, 155–161. [Google Scholar] [CrossRef]
  26. Yang, Z.-Y.; Shen, G.-Y.; He, Y.-P.; Liu, X.-X.; Yang, S.-J. Preparation of TiO2/SiO2 composite oxide and its photocatalytic degradation of rhodamine B. J. Porous Mater. 2016, 23, 589–599. [Google Scholar] [CrossRef]
  27. Siddiqa, A.; Sabir, S.; Hussain, S.T.; Muhammad, B. Highly active mesoporous SiO2-TiO2 based nanocomposites for photocatalytic degradation of textile dyes and phenol. Eur. J. Chem. 2013, 4, 388–395. [Google Scholar] [CrossRef]
  28. 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]
  29. Cheng, J.; Zhao, L.; Li, G.; Li, F.; Yao, M. B/Co/Fe tridoped TiO2/SiO2 composite films for improved photocatalytic degradation of organic pollutants under visible light. Inorg. Chem. Commun. 2020, 119, 108089. [Google Scholar] [CrossRef]
  30. Yener, H.B.; Helvacı, Ş.Ş. Effect of synthesis temperature on the structural properties and photocatalytic activity of TiO2/SiO2 composites synthesized using rice husk ash as a SiO2 source. Sep. Purif. Technol. 2015, 140, 84–93. [Google Scholar] [CrossRef]
  31. Huang, C.; Bai, H.; Huang, Y.; Liu, S.; Yen, S.; Tseng, Y. Synthesis of Neutral SiO2/TiO2 Hydrosol and Its Application as Antireflective Self-Cleaning Thin Film. Int. J. Photoenergy 2012, 2012, 620764. [Google Scholar] [CrossRef]
  32. Hoffmann, M.R.; Martin, S.T.; Choi, W.; Bahnemann, D.W. Environmental Applications of Semiconductor Photocatalysis. Chem. Rev. 1995, 95, 69–96. [Google Scholar] [CrossRef]
Figure 1. ATG (A) and DSC (B) curves of pure TiO2 and 20% T/S composite.
Figure 1. ATG (A) and DSC (B) curves of pure TiO2 and 20% T/S composite.
Catalysts 12 01001 g001
Figure 2. X-ray diffractogram of samples (A): SiO2, T/S–X at 650°C (B): TS 20%–T.
Figure 2. X-ray diffractogram of samples (A): SiO2, T/S–X at 650°C (B): TS 20%–T.
Catalysts 12 01001 g002
Figure 3. Anatase crystallite sizes of samples (A): T/S 20%–T (T = Temperature); (B): T/S–X (X = wt% SiO2) calcined at 650 °C.
Figure 3. Anatase crystallite sizes of samples (A): T/S 20%–T (T = Temperature); (B): T/S–X (X = wt% SiO2) calcined at 650 °C.
Catalysts 12 01001 g003
Figure 4. Nitrogen adsorption–desorption isotherm curves for (A) SiO2, (B) TiO2, (C) T/S 5%, (D) T/S 20%, (E) T/S 30%, (F) T/S 40%, (G) T/S 50%.
Figure 4. Nitrogen adsorption–desorption isotherm curves for (A) SiO2, (B) TiO2, (C) T/S 5%, (D) T/S 20%, (E) T/S 30%, (F) T/S 40%, (G) T/S 50%.
Catalysts 12 01001 g004aCatalysts 12 01001 g004b
Figure 5. Scanning electron microscopy (SEM) of (A) SiO2, (B) TiO2, (C) T/S 20%, and (D) EDX-T/S 20%.
Figure 5. Scanning electron microscopy (SEM) of (A) SiO2, (B) TiO2, (C) T/S 20%, and (D) EDX-T/S 20%.
Catalysts 12 01001 g005
Figure 6. IR spectrum of (A) silica, the composites (B) T/S–X; (C) T/S–20%–T.
Figure 6. IR spectrum of (A) silica, the composites (B) T/S–X; (C) T/S–20%–T.
Catalysts 12 01001 g006aCatalysts 12 01001 g006b
Figure 7. XPS spectrum of T/S 20% (A), Ti 2p (B), O 1s (C), and Si 2p (D) composites.
Figure 7. XPS spectrum of T/S 20% (A), Ti 2p (B), O 1s (C), and Si 2p (D) composites.
Catalysts 12 01001 g007
Figure 8. Photocatalytic degradation mineralization of SDCF in the presence of catalysts with different SiO2 content compared to pure TiO2 and commercial P25 (A,B) and in the presence of TS 20%–T catalyst calcined at different temperatures compared to pure TiO2 and commercial P25 (C,D).
Figure 8. Photocatalytic degradation mineralization of SDCF in the presence of catalysts with different SiO2 content compared to pure TiO2 and commercial P25 (A,B) and in the presence of TS 20%–T catalyst calcined at different temperatures compared to pure TiO2 and commercial P25 (C,D).
Catalysts 12 01001 g008
Table 1. Crystallographic parameters of the synthesized materials.
Table 1. Crystallographic parameters of the synthesized materials.
Samples2θ (°)a = bcc/a RatioAnatase Nanoparticles Size (nm)Microstrain × 10−3Dislocation Density × 10−3
T/S 20% 45025.373.78509.51402.5146.3723.9724.61
T/S 20% 55025.323.78589.50742.5118.8517.3512.78
T/S 20% 65025.343.78429.51462.51415.489.904.17
T/S 20% 75025.313.78509.51402.51421.037.312.26
TiO2 pur 65025.353.78509.51402.51426.035.881.48
Table 2. Effect of silica on the specific surface area (SBET) of materials compared to pure TiO2.
Table 2. Effect of silica on the specific surface area (SBET) of materials compared to pure TiO2.
SamplesSiO2TiO2T/S5%T/S 20%T/S 30%T/S 40%T/S 50%
specific surface area (m2/g)455.228.9925.4197.72144.29171.75205.69
Pores volume (cm3/g)0.7180.0470.0710.230.290.330.35
Pores diameter (nm)9.6424.11213.8911.199.9910.038.94
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Dantio Nguela, C.B.; Manga, N.H.; Marchal, C.; Abega, A.V.; Nsami, N.J.; Robert, D. Effect of Biogenic Silica Behavior in the Incorporation of Mesoporous Anatase TiO2 for Excellent Photocatalytic Mineralization of Sodium Diclofenac. Catalysts 2022, 12, 1001. https://doi.org/10.3390/catal12091001

AMA Style

Dantio Nguela CB, Manga NH, Marchal C, Abega AV, Nsami NJ, Robert D. Effect of Biogenic Silica Behavior in the Incorporation of Mesoporous Anatase TiO2 for Excellent Photocatalytic Mineralization of Sodium Diclofenac. Catalysts. 2022; 12(9):1001. https://doi.org/10.3390/catal12091001

Chicago/Turabian Style

Dantio Nguela, Christian Brice, Ngomo Horace Manga, Clément Marchal, Aimé Victoire Abega, Ndi Julius Nsami, and Didier Robert. 2022. "Effect of Biogenic Silica Behavior in the Incorporation of Mesoporous Anatase TiO2 for Excellent Photocatalytic Mineralization of Sodium Diclofenac" Catalysts 12, no. 9: 1001. https://doi.org/10.3390/catal12091001

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