Next Article in Journal
Research on the Simulation Model of Continuous Fiber-Reinforced Composites Printing Track
Next Article in Special Issue
Functionalized GO/Hydroxy-Terminated Polybutadiene Composites with High Anti-Migration and Ablation Resistance Performance
Previous Article in Journal
Biosorbents Based on Biopolymers from Natural Sources and Food Waste to Retain the Methylene Blue Dye from the Aqueous Medium
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Structural Characterization of Titanium–Silica Oxide Using Synchrotron Radiation X-ray Absorption Spectroscopy

by
Arpaporn Teamsinsungvon
1,2,3,
Chaiwat Ruksakulpiwat
1,2,3,
Penphitcha Amonpattaratkit
4 and
Yupaporn Ruksakulpiwat
1,2,3,*
1
School of Polymer Engineering, Institute of Engineering, Suranaree University of Technology, Nakhon Ratchasima 30000, Thailand
2
Center of Excellence on Petrochemical and Materials Technology, Chulalongkorn University, Bangkok 10330, Thailand
3
Research Center for Biocomposite Materials for Medical Industry and Agricultural and Food Industry, Nakhon Ratchasima 30000, Thailand
4
Synchrotron Light Research Institute (SLRI), 111 University Avenue, Muang District, Nakhon Ratchasima 30000, Thailand
*
Author to whom correspondence should be addressed.
Polymers 2022, 14(13), 2729; https://doi.org/10.3390/polym14132729
Submission received: 8 June 2022 / Revised: 29 June 2022 / Accepted: 30 June 2022 / Published: 3 July 2022
(This article belongs to the Special Issue Nanoparticles and Polymers: Preparations and Applications)

Abstract

:
In this study, titania–silica oxides (TixSiy oxides) were successfully prepared via the sol–gel technique. The Ti and Si precursors were titanium (IV), isopropoxide (TTIP), and tetraethylorthosilicate (TEOS), respectively. In this work, the effects of pH and the Ti/Si atomic ratio of titanium–silicon binary oxide (TixSiy) on the structural characteristics of TixSiy oxide are reported. 29Si solid-state NMR and FTIR were used to validate the chemical structure of TixSiy oxide. The structural characteristics of TixSiy oxide were investigated using X-ray diffraction, XRF, Fe-SEM, diffraction particle size analysis, and nitrogen adsorption measurements. By applying X-ray absorption spectroscopy (XAS) obtained from synchrotron light sources, the qualitative characterization of the Ti–O–Si and Ti–O–Ti bonds in Ti–Si oxides was proposed. Some Si atoms in the SiO2 network were replaced by Ti atoms, suggesting that Si–O–Ti bonds were formed as a result of the synthesis accomplished using the sol–gel technique described in this article. Upon increasing the pH to alkaline conditions (pH 9.0 and 10.0), the nanoparticles acquired a more spherical shape, and their size distribution became more uniform, resulting in an acceptable nanostructure. TixSiy oxide nanoparticles were largely spherical in shape, and agglomeration was minimized. However, the Ti50Si50 oxide particles at pH 10.0 become nano-sized and agglomerated. The presence of a significant pre-edge feature in the spectra of Ti50Si50 oxide samples implied that a higher fraction of Ti atoms occupied tetrahedral symmetry locations, as predicted in samples where Ti directly substituted Si. The proportion of Ti atoms in a tetrahedral environment agreed with the value of 1.83 given for the Ti–O bond distance in TixSiy oxides produced at pH 9.0 using extended X-ray absorption fine structure (EXAFS) analysis. Photocatalysis was improved by adding 3% wt TiO2, SiO2, and TixSiy oxide to the PLA film matrix. TiO2 was more effective than Ti50Si50 pH 9.0, Ti50Si50 pH 10.0, Ti50Si50 pH 8.0, and SiO2 in degrading methylene blue (MB). The most effective method to degrade MB was TiO2 > Ti70Si30 > Ti50Si50 > Ti40Si60 > SiO2. Under these conditions, PLA/Ti70Si30 improved the effectiveness of the photocatalytic activity of PLA.

Graphical Abstract

1. Introduction

In recent years, binary oxides of Ti and Si have gained a great deal of attention in a wide range of applications, including TiO2–SiO2 [1], Ag–TiO2–SiO2 [2], and the ordered mesoporous TiO2/SBA-15 matrix [3]. Sol–gel processing is an adjustable method that has attained commercial success because it allows for the control of the desired properties of a material during the preparation process. As a result, the sol–gel method has become one of the most preferred methods to produce oxide materials. Titanium dioxide (TiO2), often known as titania, is an important substance. This multifunctional material may be used in coatings [4,5], biomaterials [6], photocatalysts [7], food packaging [8], and chemical sensors [9], to mention a few applications. It possesses a number of desirable characteristics, including chemical resistance, chemical stability [10], photocatalytic activity, outstanding photostability, biocompatibility, and antibacterial activity [11]. Furthermore, titanium dioxide (TiO2) is widely recognized as a large-bandgap semiconductor with photocatalytic activity, whereas silicon dioxide (SiO2), often known as silica, has a well-defined ordered structure, a large surface area, is cost-effective to produce, and is easy to modify [12].
Adding SiO2 to the formulation of a polymer is also known to increase its modulus of elasticity, strength, heat and fire resistance, wear resistance, insulating properties, and other properties [13]. In addition to drug delivery, bioimaging, gene transport, and engineering, SiO2 is employed as a food ingredient. Moreover, the FDA has classified silica as a “generally recognized as safe” (GRAS) substance, making it an attractive candidate for biological activities [14].
Titanium–silicon oxide (TixSiy) is a fascinating material family that has gained a great deal of attention in recent years. The oxide has been widely employed as a catalyst and support for a variety of processes. It is constructed of TiO2 and SiO2, and it not only possesses the advantages of both TiO2 (photocatalytic properties and antimicrobial biomaterials) and SiO2 (high thermal stability and excellent mechanical strength), but it also expands their applications by developing additional catalytic active sites, owing to the chemical link between the two materials [15]. There are several techniques for manufacturing TixSiy oxide, but sol–gel approaches appear to be the most cost-effective. These methods do not necessitate the use of costly solvents. Titanium alkoxide (e.g., titanium isopropoxide) is frequently utilized in sol–gel techniques.
One of the most prevalent methods for producing oxide materials is the sol–gel method. This method allows for the management of textural characteristics, as well as a high degree of mixing in bulk mixed oxides while retaining the oxide material anion-free. Furthermore, the materials produced using this method are not microstructurally ordered, have a large surface area, and show pore size variation [16,17].
The powerful technique of synchrotron X-ray absorption spectroscopy (XAS) using tunable, very intense X-rays from a high-energy electron storage ring has been applied to investigate the structural properties of materials. XAS is particularly attractive because of its ability to deliver electronic structure as well as geometric information. A typical K-edge absorption spectrum is divided into two sections: (i) the X-ray absorption near-edge structure (XANES) (<50 eV) contains bond lengths and angles, as well as information about the three-dimensional structure around the absorbing atom; (ii) the extended X-ray absorption fine structure (EXAFS) region (typically > 50 eV) provides information on the initial coordination shell around the absorbing atom, including coordination numbers and bond lengths [18].
In most of the previous reports, XAS has been employed to determine information on the coordination environment of tetravalent Ti[Ti(IV)] in structurally complex oxide materials and bond distances between Ti–O and Ti–Si atoms. Niltharach et al. [19] used XANES methods to investigate the structural details of sol–gel-produced TiO2 samples with and without the inclusion of Ce. The XANES results also showed that the sample produced under the low hydrolysis condition had a significant number of Ti atoms in forms other than anatase and rutile TiO2. Won Bae Kim et al. [20] used the linear combination of two reference XANES spectra to estimate the pre-edge of the Ti K-edge in order to quantitatively analyze the percentages of Ti–O–Si and Ti–O–Ti bonds. The findings of pre-edge fitting in conjunction with XRD and XPS suggested that monolayer coverage was attained at around 7–10 wt.% Ti loading, where the concentration of Ti in Ti–O–Si was saturated to 0.56 mmol-Ti/g material. Shuji Matsuo et al. [21] determined the local Ti environments in the sol, gel, and xerogels of titanium oxide prepared by a sol–gel method using titanium K-edge XANES. All of the samples could be divided into three groups: the anatase group, the anatase-like structure group, and the weak Ti–Ti interaction group.
A variety of materials were produced by changing the composition of the primary sol and/or the thermal treatment. The nature of the precursor, the molar ratio between silicon and titanium alkoxide, the type of solvent, and the use of modifying agents all had an influence on the final material’s microstructure and hence its properties. Furthermore, while manufacturing TixSiy oxide via sol–gel, the pH of the solution is one of the most essential factors determining TixSiy oxide’s properties. The hydrolysis and condensation behavior of a solution during gel formation, as well as the structure of the binary oxide, is influenced by its pH [22]. In order to maintain the sample in a powder morphology with a large surface area, the gel must be stabilized. The goal of this study was to evaluate the effects of pH and Ti/Si atomic ratio on TixSiy oxide using the sol–gel technique (Stöber method). XANES and EXAFS were used to evaluate Ti–O–Si and Ti–O–Ti connectivity in TixSiy oxides, the local atomic structure, the bond distances between Ti–O and Ti–Si atoms, the coordination number, and the valence state of titanium atoms TiO2 and all mixed oxide samples. In addition, the photocatalytic behavior of all samples was investigated by comparing their degradation of methylene blue (MB) solutions exposed to UV radiation.

2. Materials and Methods

2.1. Materials

Tetraethylorthosilicate (TEOS, 98%, AR-grade) and titanium (IV) isopropoxide (TTIP, 98%, AR-grade) were purchased from Acros (Geel, Belgium). Absolute ethanol (C2H5OH, AR-grade), hydrochloric acid (HCl, AR-grade), and ammonium hydroxide (NH4OH, AR-grade) were supplied by Carlo Erba Reagents (Emmendingen, Germany). PLA4043D was provided by NatureWorks LLC (Minnetonka, MN, USA).

2.2. Synthesis of Titania–Silica Binary Oxide (TixSiy Oxide), Silica, and Titanium Dioxide

Sol–gel techniques were used to prepare titanium–silicon oxide (TixSiy) utilizing a modified Stöber procedure, including the simultaneous hydrolysis and condensation of TEOS [20]. Figure 1 illustrates the experimental technique for producing TixSiy oxide particles. TEOS and TTIP were used without purification.
The total moles of Ti-alkoxide and Si-alkoxide were determined to be 0.12 mol. For the pre-hydrolysis step, the mixture of two alkoxides was added into a five-neck round-bottom flask containing 150 cm3 of ethanol for 1 h. A condenser was used in the condensation step to prevent the solvent from vaporizing out of the process. Stirring raised the temperature of the reactor to 75–80 °C. After an hour of mixing, a yellowish translucent sol was observed after adding TTIP and 25 ml of ethanol to the reactor drop by drop for 1 h. DI water was added to the sol, which was then aged at room temperature for 2 h at 80 °C. The aged gel was then refined in a centrifugal separator for 10 min at 6000 rpm. Finally, the solid particles were calcined for 3 h at 450 °C. A reference pure SiO2 sample was prepared in NH4OH in the same manner as described above but without the addition of TTIP [23]. In addition, reference pure TiO2 was generated in the same maner as TixSiy oxide in acid [24]. Table 1 and Table 2 reveal the composition of the sol–gel liquid solution used to produce TixSiy oxide.

2.3. Characterization of Titanium–Silicon Oxide

2.3.1. 29Si Solid-State Nuclear Magnetic Resonance Spectroscopy (29Si Solid-State NMR)

The 29Si solid-state NMR spectra were used to evaluate the structure of the silicate phase in SiO2 and TixSiy oxide. 29Si NMR spectra were recorded on a Bruker Avance III HD 500 MHz spectrometer (Billerica, MA, USA).

2.3.2. Fourier Transform Infrared Spectroscopy (FT-IR)

Infrared spectra of TixSiy oxide, TiO2, and SiO2 were recorded by Fourier transform infrared spectroscopy (FTIR) (Bruker, Tensor 27, Billerica, MA, USA) using attenuated total reflectance (ATR) equipped with a platinum diamond crystal (TYPE A225/QL). The spectra were recorded at wavenumbers from 400 to 4000 cm−1 with 4 cm−1 resolution through the accumulation of 64 scans. All samples of TixSiy oxide, TiO2, and SiO2 were dried in an oven at 70 °C for 4 h before testing.

2.3.3. Field Emission Scanning Electron Microscopy (FE-SEM)

FE-SEM images of TixSiy oxide, TiO2, and SiO2 were examined by a field-emission scanning electron microscope (FESEM-EDS, Carl Zeiss Auriga, Oberkochen, Germany). The film specimens were coated with carbon prior to the investigation. An acceleration voltage of 3 kV was used to collect SEM images of the samples.

2.3.4. X-ray Diffraction (XRD)

The diffractograms of TixSiy oxide, TiO2, and SiO2 were recorded via powder X-ray diffraction (Bruker, Model D2 phaser, Billerica, MA, USA), with CuKα radiation, scanning from 10° to 90° at a rate of 0.05°/s, with a current of 35 mV and 35 mA.

2.3.5. X-ray Absorption Near-Edge Structure Spectroscopy (XANES) and Extended X-ray Absorption Fine Structure Spectroscopy (EXAFS)

All TiO2 and SiO2 standard compounds and TixSiy oxide samples were analyzed by XANES and EXAFS spectroscopy at the Ti K-edge at beamline 8 of the electron storage ring (using an electron energy of 1.2 GeV, a bending magnet, beam current of 80–150 mA, and 1.1 to 1.7 × 1011 photons s−1) at the Synchrotron Light Research Institute (SLRI), Nakhon Ratchasima, Thailand [25]. Finely ground, homogenized powder of each sample was spread as a thin film (area 2.0 cm × 0.5 cm) and carefully dispersed with a spatula to yield a homogeneous particle distribution and to avoid hole effects on Kapton tape (Lanmar Inc., Northbrook, IL, USA) mounted on a sample holder. At least five spectra were obtained for each standard compound and oxide sample. All XANES and EXAFS spectra were measured in the transmission mode with ionization chamber detectors. For the acquisition of all spectra, a Ge (220) double crystal monochromator with an energy resolution (ΔE/E) of 2 × 10−4 was used. An energy range of 4936–5824 eV with an energy step of 2, 0.2, 0.05k eV was used for Ti K-edge spectra. The photon energy was calibrated against the K-edge of Ti foil at 4966 ± 0.2 eV. Finally, the normalized XANES and EXAFS data were processed and analyzed after background subtraction in the pre-edge and post-edge regions using a software package including (i) ATHENA (Chicago, IL, USA) for XAS data processing, (ii) ARTEMIS (Chicago, IL, USA) for EXAFS data analysis using theoretical standards from FEFF, and (iii) HEPHAESTUS (Chicago, IL, USA) software for a collection of beamline utilities based on tables of atomic absorption data. This package is based on the IFEFFIT library of numerical and XAS algorithms and is written in the Perl programming language using the Perl/Tk graphics toolkit [26].

2.3.6. X-ray Fluorescence (XRF)

The chemical compositions of TixSiy oxide, TiO2, and SiO2 were investigated by energy-dispersive XRF (EDS Oxford Instrument ED 2000, Abingdon, UK) with a Rh X-ray tube with a vacuum medium.

2.3.7. Particle Size Distribution

The particle size distribution of TixSiy oxide, TiO2, and SiO2 was evaluated by a diffraction particle size analyzer (DPSA) (Malvern Instruments, model Mastersizer 2000, Great Malvern, England). The TixSiy oxide, TiO2, and SiO2 were dispersed in absolute ethanol and analyzed by a He-Ne laser. The average particle size distribution was investigated from the standard volume percentiles at 10, 50, and 90%. The average volume-weighted diameter was used to define the average particle size.

2.3.8. Specific Surface Area (BET)

The specific surface areas of TixSiy oxide, TiO2, and SiO2 were investigated by a nitrogen adsorption analyzer (BELSORP MINI II, model Bel-Japan, Osaka, Japan). The samples of of TixSiy oxide, TiO2, and SiO2 were degassed at 300 °C for 3 h before measurement. The Brunauer, Emmett, Teller (BET) and Barrett, Joyner, Halenda (BJH) methods were used to calculate the specific surface area and pore size distribution, respectively.

2.3.9. Photocatalytic Activity Characterization

PLA, PLA/TiO2, PLA/SiO2, and PLA/TixSiy oxide composites with 3 wt.% of TixSiy oxide were prepared by the solvent-casting method. The polymer was dissolved in solvent (10% w/v) by stirring. Firstly, TixSiy oxide was dispersed in chloroform with ultrasonic treatment for 1 day. After this, PLA was added to the TixSiy oxide and strictly stirred for 4 days. The dispersions of PLA composites were additionally ultrasonically treated for 1 h using a frequency of 42 kHz, four times per day. The treated dispersions were gently poured onto Petri dishes, and the solvent was evaporated at room temperature. The films were dried to a constant mass, dried at room temperature for 24 h, and kept in the oven at 40 °C for 4 h. Film thickness was measured by using a micrometer in eight replicates for each sample, from which an average value was attained. Pure PLA and nanocomposite films had a uniform thickness of 200–250 µm.
The photocatalytic activity of PLA composite films was evaluated in terms of the degradation of methylene blue (MB) under UV light according to Chinese standard GB/T 23762-2009. PLA composite films were placed in a flask and then 200 mL of MB solution (10 mg/L) was added. The flask was placed on a mechanical shaker at 50 rpm in a UV chamber with 4 UV lamps (LP Hg lamps, 8 watts, main light emission at 245 nm). The schematic design of the investigational setup for the photocatalytic process is presented in Figure 2. Then, 4 mL MB solution was collected every 60 min and analyzed using a UV–vis spectrophotometer (Cary300, Agilent Technology, CA, USA). In order to maintain the volume of MB solution in the flask, the samples were returned after each measurement. The maximum absorbance of MB occurred at 664 nm (Figure 3). The spectrometer was calibrated with a solution of MB at 1mg/L, 3mg/L, 5mg/L, 7mg/L, and 10mg/L concentrations, respectively. The calibration curve of methylene blue aqueous solutions is shown in Figure 4. In order to precisely predict the decomposition of MB under UV light, the PLA and composite films, as well as the solution, were stored in a black box without any photocatalyst. The concentration of MB was also collected every 60 min to estimate the absorption of MB.

3. Results

3.1. Effect of pH

29Si Solid-state NMR spectroscopy was used to investigate the structures of TixSiy oxide (particularly for 1:1 atomic ratio) and SiO2 in order to determine the development of the framework of silicon atoms in these materials. The structural analysis of 29Si NMR is illustrated in Figure 5, which presents the spectra of the silica and the Ti50Si50 oxide. In general, the 29Si NMR spectra of oxides show three peaks for different environments for Si: isolated silanol groups (SiO)3Si–OH, Q3; geminal silanols (SiO)2Si–(OH)2, Q2; and silicon in the siloxane binding environment without hydroxyl groups (SiO)4Si, Q4 (Qn represents a SiO4 unit, with n being the number of bridging in oxygen atoms) [27,28]. According to related references, the three resonance signals of pure SiO2 (synthesized by sol–gel method) appearing at −93.9, −103.3, and −110.2 ppm can be attributed to the configurations of Q2, Q3, and Q4, respectively. The spectrum of silica (Figure 5a) exhibits a chemical shift of −110 ppm, which is attributed to the Q4 units, and a minor peak appears at −103 ppm that would conform to the Q3 units corresponding to the siloxane (Si–O–Si) and silanol (≡Si–OH) groups of silica. The strength of Q4 structural units [Si(SiO)4] represents a three-dimensional network of silica, as evidenced by the intensity of the peaks in Figure 5a. There would also be a very small proportion of Si nuclei directly bonded to the hydroxyl groups, influenced by the chemical shift of the Q3 structural unit [Si(SiO)3OH] [17]. Additionally, Figure 5b indicates that silica was chemically combined with titanium oxide for the formation of the Ti50Si50 oxide, causing Ti atoms to be disrupted in the three-dimensional silica network and the microstructure surrounding the silicon atoms. When comparing the peak at approximately -103 ppm to the quantity of siloxane (Si–O–Si) and silanol (≡Si–OH) groups in pure SiO2, a very substantial increase in Q3 units (mono-substitution) is detected, which might be due to the existence of significant Si–O–Ti bonds in this sample, signaling atomic mixing. The presence of Ti–O–Si–O–Ti–O groups may have caused the appearance of Q2 units, which might be linked to disubstituted silica in tetrahedra [29]. A similar observation was reported by Pabón, Retuert, and Quijada (2007) [17]. Obviously, Ti atoms replaced some Si atoms in the SiO2 network, indicating that, as a consequence of the synthesis performed by the sol–gel method defined in this work, a large number of Si-O–Ti bonds were created, in agreement with the results of the FTIR and XRD.
FTIR spectra of TiO2, SiO2, and Ti50Si50 oxides synthesized at various pH values are shown in Figure 6. TiO2 and all Ti50Si50 oxides displayed a prominent band at 400 to 700 cm−1 in their FTIR spectra, which corresponded to the bending and stretching mode of Ti–O–Ti and was characteristic of well-ordered TiO6 octahedrons [30].
Strong bands near 3440 and 1630 cm−1 can be seen in Figure 6, which are assigned to the streching and deformation vibration of the hydrocyl groups present (TiO2–OH) on the surfaces of TiO2 and all mixed oxide samples. The intensive band of OH-group asymmetrical and symmetrical stretching vibrations at 3440 cm−1 and O–H deformation vibration at 1630 cm−1 could confirm the large amount of water molecules [31]. However, the OH-group vibration bands were substantially weaker as the calcination temperatures increased [32]. The peaks at 440, 801, and 1052 cm−1 corresponded to the rocking, symmetric, and asymmetric stretching vibrations of Si–O–Si of silica, respectively [3]. However, TiO2 also presents a peak at 1000-1200 cm−1, which is assigned to the deformation vibration of Ti–O–Ti of TiO2 [33]. The peak at 949 cm−1 was commonly attributed to Ti–O–Si vibrations by various authors, indicating the existence of Ti–O–Si linkages [34]. This band can also be attributed to the Si–OH bond, which should be attributed to Si–OH bonding. However, this absorption band disappeared after the SiO2 particles were calcined at 450 °C. On the other hand, the 949 cm−1 absorption band was still observed for all TixSiy oxide samples that had been calcined at a temperature of 450 °C. Accordingly, the 949 cm−1 peak in the FTIR spectra for TixSiy oxide particles should not be attributed to Si–OH bonds but rather be associated with Si–O–Ti bonding [35,36]. This indicated titanium being combined into the framework of silica. Therefore, the influence of pH on the FTIR of the Ti50Si50 oxide was not observed. Based on the evidence from NMR and FTIR, the probable mechanism for the synthesized TixSiy oxides based on results of earlier studies [37] is suggested in Figure 7.
Figure 8 shows FE-SEM images of SiO2 and Ti50Si50 oxide powder nanoparticles synthesized at various pH values. SiO2 particles were mostly spherical in shape, with low agglomeration (Figure 8a). Meanwhile, large bulk particles with high agglomeration of Ti50Si50 oxides are shown in Figure 8b. This agglomeration with an irregular shape was a result of the Ti and Si precursors in the sol solution during the processing, synthesized due to the absence of appropriate hydroxide ions (OH ions) at pH 8.0 [38]. Figure 8c,d show that the nanoparticles were spherical in shape and the size distribution of particles became more uniform, with a respectable nanostructure, upon increasing the pH to alkaline conditions (pH 9.0 and 10.0). In this work, TixSiy oxide nanoparticles were essentially spherical in form, with low agglomeration. However, the size of the Ti50Si50 oxide particles at pH 10.0 was nano-sized, with the appearance of agglomeration (Figure 8d). Wu, Wu, and Lü, (2006) claimed that the colloidal sol–gel technique can generate large particles composed of agglomerated nanoparticles and either condensed or porous polycrystalline microparticles [39]. Therefore, pH 9.0 was selected to achieve the compositionally controlled synthesis of TixSiy oxide particles that are spherical in shape.
XRD patterns of Ti50Si50 oxide powders synthesized at various pH values (8.0, 9.0, and 10.0) are shown in Figure 9. Silica showed one amorphous peak located at 2ϴ angles at 23°. In most of the Ti50Si50 oxide powders prepared at pH 8.0 and 9.0 (Figure 9), XRD patterns exhibited a broad amorphous peak because of the existence of silica. Furthermore, the formation of the Ti50Si50 oxides suggests that titanium oxide was combined with the silica by atomic mixing, and consequently the distribution of Ti atoms occurred in the microstructure around the silicon atoms and the three-dimensional silica network. However, the Ti50Si50 oxide powders prepared at pH 10.0 exhibited an intense TiO2 anatase peak, suggesting that the formation of higher crystallites of anatase occurred in Ti50Si50 oxides when the high pH of the sol–gel solution was applied during hydrolysis [38]. It can be clearly seen that the crystallinity of anatase in Ti50Si50 oxide samples is enhanced by adjusting the pH value of the sol.
Furthermore, the XRD peaks for Ti50Si50 oxide powder with pH > 9 present a crystalline nature with 2ϴ angles at 25° (101) and 48° (200), which correspond to the crystalline phase (anatase) of TiO2 [3,34], due to the presence of a sufficient amount of OH to form TiO2. It is possible that the pH influences the particle size and crystallinity of Ti50Si50 oxides. The crystal structure of TiO2 was tetragonal; a = b = 3.782 Å, c = 9.502 Å [40]. However, no rutile phase was detected in any case, according to the absence of (110) diffraction peak at 27.4°.
Titanium K-edge X-ray-absorption near-edge structure (XANES) spectroscopy is widely used to derive information on the coordination environment of tetravalent Ti[Ti(IV)] in structurally complex oxide materials. Normalized Ti K-edge XANES spectra of anatase TiO2 standards, synthesized TiO2 materials, and Ti50Si50 oxides synthesized at various pH values were collected and are presented in Figure 10. The spectrum is mostly separated into two regions: (i) the pre-edge region (4940 to 4990 eV) and (ii) the post-edge region (4990 to 5010 eV) [41]. The anatase is composed of pre-edge features A1–A3 and a characteristic shoulder B [42,43]. The origin of these features was described by Farges et al. The first pre-edge (A1) is mainly attributed to quadrupolar transitions to t2g levels of the TiO6 octahedron. The second pre-edge (A2) and the third pre-edge (A3) are attributed to 1s to 3d dipolar transitions [43]. The B features may be attributed to the interactions of the central Ti 4p orbital hybridized with the near Ti or O atom (Wu et al., 1997). The intensive absorbance of the B feature may be due the local structures of the Ti orbital hybridized for the TiO2–SiO2 photocatalyst [41]. Both TiO2 anatase and synthesized TiO2 display three minor pre-edge peaks, which are attributed to transitions from the 1s level of Ti to 1t1g, 2t2g, and 3eg molecular orbitals (in sequence of increasing energy) [44]. The three minor pre-edge peaks for TiO2 anatase are outstanding fingerprints for the crystalline phase of titanium dioxide materials. Thus, the formation of anatase is specified by the near-edge region of TiO2.
As silicon merged into the Ti50Si50 oxide, a large single pre-edge feature was observed to dominate the other weaker feature. From Figure 10, the significant pre-edge feature in Ti50Si50 oxide samples indicates that a larger proportion of Ti atoms occupy tetrahedral symmetry positions, as expected for samples where Ti directly replaces Si in the oxide sample. Moreover, the pre-edge area of Ti50Si50 oxides was not affected by the various pH values applied during synthesis. Furthermore, the XANES spectra of the Ti50Si50 oxides have a higher intensity of pre-edge peaks than TiO2 anatase, indicating that Ti is no longer entirely confined to octahedral sites when combined with Si. Therefore, the results of the XANES spectra of mixed oxide samples are in reasonable agreement with the XRD results.
The EXAFS analysis at the Ti K-edge was used to investigate the coordination number of Ti atoms and bond distances between Ti–O and Ti–Si atoms. The program package of ATHENA–ARTEMIS was used to perform the background noise correction and the normalization of the raw data [26]. The EXAFS data were FT to the R-space to define the local atom structure and relative bond length with respect to the absorbing atom. The ATOM and FEFF codes were used to simulate the FT data (in R-space) by creating a systematic theoretical structure of SiO2 [45]. The range of data reserved for the transformation was 3–9 Å−1 in k-space. Structural parameters were attained, without the phase corrections, by fitting the data in R-space, within the intermission of 1.1–3 Å (Figure 11). The shell parameters, such as Ti–O and Ti–Si bond distances (R), coordination number (CN), Debye–Waller factor (σ2), and R-factor, are presented in Table 3. Note that the amplitude reduction factor ( S 0 2 ) was preliminarily determined to be 0.52 by fitting the EXAFS spectrum of SiO2.
The first and second shells in the FTs arise from the single scattering paths of cation–oxygen and cation–cation, respectively [46]. In this study, we first focus on the Ti–O and Ti–Si shells of all TixSiy oxide samples. TiO2 anatase has an average Ti–O bond distance of 1.95 Å. The results from curve fitting of the EXAFS provide an interatomic distance of 1.99, 1.82, and 1.94 Å for Ti atom dilute silica of Ti50Si50 oxides synthesized at various pH levels of 8.0, 9.0, and 10.0, respectively. This result can be compared favorably with the bond distance of Ti–O in the titanium silicalite molecular sieve of 1.80 Å, corresponding to the cation existing in a tetrahedral environment [47,48]. Neurock and Manzer (1996) calculated the Ti–O bond distance using density functional methods, obtaining a value of of 1.81 Å, suggesting Ti in a tetrahedral cluster [49]. On the other hand, we found that there is evidence of a longer Ti–O distance at 1.99 at pH 8.0, indicating a six-fold Ti site (octahedral), which agrees with Loshmanov et al. [47], who investigated TiO2 in SiO2 with neutron diffraction and attempted to isolate the Ti–O bond by subtracting their radial distribution functions from that of pure SiO2. They assigned a value of 1.95 A. By analogy with the Ti–O distance in its octahedrally bonded oxides (1.91–2.01 Å), they inferred an octahedral coordination for Ti in SiO2. This may be due to the effect of pH on the reaction of the TixSiy oxides synthesized. Obviously, the value of 1.83 Å described for the Ti–O bond distance in TixSiy oxides synthesized at pH 9.0 is agreeable with the absolute value of the Ti atoms occupied in a tetrahedral environment, which supports the comparable assumption from XANES. Furthermore, all of the Ti50Si50 oxides presented a small value for the Debye–Waller factor, corresponding to the structural disturbance in the materials [45]. The Debye–Waller factor is used to describe the influence of static and thermal disorder on the EXAFS spectrum. A large value for the σ2 can be caused by a variance in the ligand distances (static disorder), whereas small Debye–Waller factors artificially increase the ligand contribution [50,51].
The chemical compositions of TiO2, SiO2, and Ti50Si50 oxide synthesized at various pH values by XRF are shown in Table 4. The XRF analysis showed that Ti50Si50 oxide synthesized at pH 8.0 showed a higher percentage of Si (54.39%) than those synthesized at pH 9.0 (47.87%) and pH 10.0 (9.82%) because TTIP can be hydrolyzed more quickly than TEOS in a solution at a high pH. In contrast, the atomic content of Ti in Ti50Si50 oxide synthesized at various pH values increased with the increasing pH value of the sol (43.72%).
The particle sizes of TiO2, SiO2, and Ti50Si50 oxides synthesized at various pH values as determined by the Zetasizer Nano ZS and SEM micrographs are shown in Table 5 and Figure 12. The mean particle size distributions of TiO2 and SiO2 were 40.3 and 147.8 nm, while Ti50Si50 oxide synthesized at pH 9.0 presented a mean particle size distribution of 136.2 nm. However, the mean particle size distribution of Ti50Si50 oxide synthesized at pH 8.0 and 10.0 covered a range of 350–650 nm due to the tendency towards agglomeration.
To determine the correct diameter of TiO2, SiO2, and TixSiy oxide, SEM micrographs (Figure 8) were used for measurement with ImageJ. The number average diameter, dn, was calculated from a minimum of 200 particles according to the following equations [52]:
d n = i n i d i i n i
S D = i n i d i d n 2 N
where di is the diameter of a particle and ni is the total number of particles having diameter, di.
It can be observed that the number average diameter of the sample increased with the increasing pH value of the synthesis.
The textural properties of the SiO2, TiO2, and Ti50Si50 oxides synthesized at various pH values are shown in Table 6. It can be observed that the surface areas (SBET) of the oxides were 177.02, 225.68, and 62.75 m2g−1 for Ti50Si50 synthesized at pH 8.0, 9.0, and 10.0, respectively. In addition, the mean pore diameter of the Ti50Si50 oxides was 2.56, 5.85, and 27.31 nm. Meanwhile, the SBET of SiO2 and TiO2 was 116.90 and 74.07 m2g−1, with an average pore size of 37.08 and 12.75 nm.
A typical nitrogen adsorption–desorption isotherm of the SiO2, TiO2, and Ti50Si50 oxides synthesized at various pH values is presented in Figure 13. The isotherm of SiO2 is a type II curve according to the IUPAC isotherm classification, indicating that the obtained SiO2 particles contain macropores. An inflection point occurs near the completion of the first adsorbed monolayer, which is typical for non-porous or macroporous materials with a pore size of >50 nm. In addition, all Ti50Si50 oxides exhibited type II isotherms with pore diameters of 2.56, 5.85, and 27.31 nm. However, the isotherm of TiO2 is a type IV curve according to the IUPAC isotherm classification. An abrupt hysteretic loop is detected from this curve, which is typical for mesoporous materials with pores in the range of 2 to 50 nm that show capillary condensation and evaporation [53,54].

3.2. Effect of Ti/Si Ratio

FTIR spectra of TiO2, SiO2, and TixSiy oxide synthesized at pH 9.0 with different Ti/Si ratios are shown in Figure 14. All TixSiy oxides have three major peaks in their FTIR spectra, with wavenumbers of 801, 949, and 1052 cm−1. The peaks at 801 and 1052 cm−1 corresponded to symmetric and asymmetric stretching vibrations of Si-O-Si of silica, while the peak at 949 cm−1 was commonly attributed to Ti–O–Si vibrations, indicating the existence of Ti–O–Si linkages [34]. This suggested that titanium had been incorporated into the silica framework. Therefore, there was no effect of the Ti/Si atomic ratio on the FTIR of TixSiy oxide.
The FE-SEM image of TiO2 exhibits a large majority of particles with high agglomeration, as shown in Figure 15a. Ti70Si30 and Ti50Si50 oxides show particles that are homogeneous with a good nanostructure and mostly spherical in shape when Ti/Si is more than 1 (Figure 15b,c). However, the size of Ti40Si60 oxide particles becomes gradually aggregated with increasing moles of Si (Figure 15d).
From the XRD patterns, only broad-range signals due to the uniform dispersion of titanium atoms onto the SiO2 network can be seen in the XRD patterns of Ti70Si30, Ti50Si50, and Ti40Si60 oxides (Figure 16). As a result, the absence of XRD diffraction signals suggests that anatase crystallization is still hindered by the titania aggregate’s segregation from the silica network [55].
Ti K-edge XANES spectra of standard TiO2 anatase, synthesized TiO2 materials, and TixSiy oxides synthesized with different Ti/Si ratios were collected and are presented in Figure 17. Ti is expected to readily replace Si in the tetrahedral SiO2 framework at low Ti concentrations in Ti40Si60. The features and position in the rising edge of the B shoulders are clearer and the positions are lower. This suggests that the Ti–O–Ti linkages in TixSiy oxides increase and are incomplete in anatase [21].
The structural parameters of TixSiy oxide synthesized at pH 9.0 with varying Ti/Si ratios were attained without phase corrections by fitting the data in R-space within the intermission of 1.1–3 Å (Figure 18). The shell parameters, such as Ti–O and Ti–Si bond distances (R), coordination number (CN), Debye–Waller factor (σ2), and R-factor, are presented in Table 3. The amplitude reduction factor ( S 0 2 ) was estimated to be 0.52 based on the EXAFS spectra of SiO2. According to the EXAFS curve fitting findings, the interatomic distances for Ti atom dilute silica of Ti70Si30, Ti50Si50, and Ti40Si60 oxide are 1.90, 1.83, and 1.82, respectively. The average Ti–O bond distance dropped from 1.90 to 1.82 when Ti in TixSiy oxide increased. Furthermore, the 1.83 and 1.82 Ti–O bond lengths found for Ti50Si50 and Ti40Si60 oxide correspond with the absolute value of the Ti atoms occupied in a tetrahedral environment, validating the comparable assumption from XANES.
The proportion of Si atoms in Ti70Si30, Ti50Si50, and Ti40Si60 oxide was verified by XRF analysis to be 29.79, 47.87, and 60.19%, respectively. Ti content was 70.21, 52.13, and 39.81% in Ti70Si30, Ti50Si50, and Ti40Si60 oxide, respectively (Table 4).
Table 5 and Figure 19 display the particle sizes of TiO2, SiO2, and TixSiy oxides synthesized at various Ti/Si ratios, as measured by the Zetasizer Nano ZS and SEM micrographs. The mean particle size distribution of Ti70Si30, Ti50Si50, and Ti40Si60 oxide was 573.1, 136.1, and 825.5 nm, respectively. In addition, SEM micrographs (Figure 15) were used to measure the correct diameter of TiO2, SiO2, and TixSiy oxide by ImageJ. The number average diameter, dn, was estimated using Equations (1) and (2) from the minimum value given in [52]. Ti70Si30, Ti50Si50, and Ti40Si60 oxide had an average diameter of 149.3, 135.4, and 131.8 nm, respectively, as reported in Table 6. As the amount of Ti incorporated into the silica network increased, the number average diameter decreased.
The Ti70Si30, Ti50Si50, and Ti40Si60 oxides had surface areas (SBET) of 569.07, 225.68, and 68.34 m2g−1, respectively. The SBET displayed an upward trend as the Ti content increased. According to a prior study, the SBET of the TixSiy oxide was enhanced as the Ti concentration in the silica network increased [56].
A typical nitrogen adsorption–desorption isotherm of the SiO2, TiO2, and Ti70Si30, Ti50Si50, and Ti40Si60 oxides is presented in Figure 20. Type II isotherms with pore diameters of 10.96, 5.85, and 21.97 nm were found in Ti70Si30, Ti50Si50, and Ti40Si60 oxides, corresponding to mesoporous materials with pore sizes ranging from 2 to 50 nm. However, the TiO2 isotherm is classified as a type IV curve according to the IUPAC. This curve shows an abrupt hysteretic loop, which is common in mesoporous materials with pores between 2 and 50 nm that exhibit capillary condensation and evaporation [53,54].

3.3. Photocatalytic Degradation of Methylene Blue (MB)

The degradation of methylene blue (MB) was used to study the photocatalytic activity of PLA and PLA/TixSiy oxide composite films. In addition, UV irrigation may also lead to the decomposition of MB without the presence of any photocatalyst. The degradation of MB on PLA and PLA/TixSiy oxide composite films was caused by a change in MB concentration in aqueous solution under UV irrigation, as illustrated in Figure 21 and Figure 22. The addition of 3%wt TiO2, SiO2, and TixSiy oxide to the PLA film matrix increased the effectiveness of photocatalysis in the degradation of MB. TiO2 is more effective at degrading MB than Ti50Si50 pH 9.0, Ti50Si50 pH 10.0, Ti50Si50 pH 8.0, and SiO2. In the degradation of MB, the efficiency of the materials was of the following order: TiO2 > Ti70Si30 > Ti50Si50 > Ti40Si60 > SiO2.
Photocatalytic activity is generally accepted to occur at the surface of a photocatalyst. As a result, the surface area of PLA nanocomposite film, which is dependent on the size of nanoparticles, film morphology, and thickness, influences photocatalytic reactivity [7]. The PLA nanocomposite film including TiO2 degrades MB more efficiently than employing only photocatalysis because of the two mechanisms of degradation. The first is that in which UVC directly degrades MB, while the second occurs when TiO2 receives light energy greater than the bandgap energy, and electrons in the valence band (VB) are excited into the conduction band (CB), resulting in the generation of a hole (h+) (Equation (3)). This hole can oxidize MB (Equation (4)) or oxidize H2O to produce OH (Equation (5)). The e in CB can reduce O2 at the surface of TiO2 to generate O 2 (Equation (6)). Radicals (OH, O 2 ) and h+ interact with MB to create peroxide derivatives and hydroxylate or completely degrade to CO2 and H2O [7]. The photodegradation pathway can be summarized by Equations (3)–(6).
T i O 2 + U V C     e + h +
h + + M B     C O 2 + H 2 O
h + + H 2 O     H + + O H
e + O 2     O 2

4. Conclusions

Titania–silica oxides (TixSiy oxides) can be prepared by the sol–gel technique. The effect of pH and the Ti/Si atomic ratio of titanium–silicon binary oxide (TixSiy) on the structural characteristics of TixSiy oxide were investigated by using the sol–gel technique. 29Si solid-state NMR and FTIR measurements indicated that certain Si atoms in the SiO2 network had been replaced by Ti atoms, implying the formation of Si–O–Ti connections. As the pH was elevated to alkaline conditions (pH 9.0 and 10.0), the nanoparticles of Ti50Si50 oxide acquired a more spherical shape and their size distribution became more uniform, resulting in an acceptable nanostructure. Agglomeration was reduced in Ti50Si50 oxide nanoparticles, which were mostly spherical in form. However, the Ti50Si50 oxide particles at pH 10.0 become nano-sized and agglomerated. In Ti50Si50 oxide samples, the existence of a large pre-edge feature indicated that a higher percentage of Ti atoms occupied tetrahedral symmetry positions, as expected in samples where Ti directly substitutes Si. The result of 1.83 Å for the Ti–O bond distance in TixSiy oxides generated at pH 9.0 accords with the fraction of Ti atoms occupied in a tetrahedral environment. The average diameter of the sample increased with the increasing pH of the sol. In addition, Ti70Si30, Ti50Si50, and Ti40Si60 oxide had an average diameter of 149.3, 135.4, and 131.8 nm, respectively. As the amount of Ti incorporated into the silica network increased, the average diameter decreased. The SBET displayed an upward trend as the Ti content increased. Type II isotherms with pore diameters of 10.96, 5.85, and 21.97 nm were found in Ti70Si30, Ti50Si50, and Ti40Si60 oxides, corresponding to mesoporous materials with pore sizes ranging from 2 to 50 nm, which is common in mesoporous materials. However, the TiO2 isotherm was classified as a type IV curve. The addition of 3%wt TiO2, SiO2, and TixSiy oxide to the PLA film matrix increased the effectiveness of photocatalysis in the degradation of MB. TiO2 was more effective at degrading MB than Ti50Si50 pH 9.0, Ti50Si50 pH 10.0, Ti50Si50 pH 8.0, and SiO2. In the degradation of MB, the order of efficiency is TiO2 > Ti70Si30 > Ti50Si50 > Ti40Si60 > SiO2. Moreover, PLA/Ti70Si30 also increased the effectiveness of the photocatalytic activity of PLA. Finally, PLA/Ti70Si30 improved the efficiency of the photocatalytic activity of PLA.

Author Contributions

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

Funding

This research was funded by Thailand Science Research and Innovation (TSRI), grant number 42853.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors are grateful to Suranaree University of Technology (SUT); to the Center of Excellence on Petrochemical and Materials Technology (PETROMAT); to the Science, Research and Innovation Promotion Fund from Thailand Science Research and Innovation (TSRI); and to the Research Center for Biocomposite Materials for Medical Industry and Agricultural and Food Industry for the financial support. In addition, we would like to express our gratitude to the Synchrotron Light Research Institute (Public Organization) for the XAS experiment.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Rasalingam, S.; Peng, R.; Koodali, R.T. Review Article: Removal of Hazardous Pollutants from Wastewaters: Applications of TiO2-SiO2 Mixed Oxide Materials. J. Nanomater. 2014, 2014, 1–42. [Google Scholar] [CrossRef] [Green Version]
  2. Wang, X.; Xue, J.; Wang, X.; Liu, X. Heterogeneous Ag-TiO2-SiO2 composite materials as novel catalytic systems for selective epoxidation of cyclohexene byH2O2. PLoS ONE 2017, 12, e0176332. [Google Scholar] [CrossRef]
  3. Sahu, D.R.; Hong, L.Y.; Wang, S.-C.; Huang, J.-L. Synthesis, analysis and characterization of ordered mesoporous TiO2/SBA-15 matrix: Effect of calcination temperature. Microporous Mesoporous Mater. 2009, 117, 640–649. [Google Scholar] [CrossRef]
  4. Petronella, F.; Pagliarulo, A.; Truppi, A.; Lettieri, M.; Masieri, M.; Calia, A.; Curri, M.L.; Comparelli, R. TiO2 Nanocrystal Based Coatings for the Protection of Architectural Stone: The Effect of Solvents in the Spray-Coating Application for a Self-Cleaning Surfaces. Coatings 2018, 8, 356. [Google Scholar] [CrossRef] [Green Version]
  5. Zhu, Y.; Buonocore, G.G.; Lavorgna, M. Photocatalytic activity of PLA/TIO2 nanocomposites and TIO2-active multilayered hybrid coatings. Ital. J. Food Sci. 2012, 24, 102–106. [Google Scholar]
  6. Tiainen, H.D.; Wiedmer, H.; Haugen, J. Processing of highly porous TiO2 bone scaffolds with improved compressive strength. J. Eur. Ceram. Soc. 2013, 33, 15–24. [Google Scholar] [CrossRef]
  7. Chen, L.; Zheng, K.; Liu, Y. Geopolymer-supported photocatalytic TiO2 film: Preparation and characterization. Constr. Build. Mater. 2017, 151, 63–70. [Google Scholar] [CrossRef]
  8. Allodi, V.; Brutti, S.; Giarola, M.; Sgambetterra, M.; Navarra, M.A.; Panero, S.; Mariotto, G. Structural and Spectroscopic Characterization of a Nanosized Sulfated TiO2 Filler and of Nanocomposite Nafion Membranes. Polymers 2016, 8, 68. [Google Scholar] [CrossRef] [Green Version]
  9. Bertuna, A.; Comini, E.; Poli, N.; Zappa, D.; Sberveglieri, G. Titanium Dioxide Nanostructures Chemical Sensor. Procedia Eng. 2016, 168, 313–316. [Google Scholar] [CrossRef] [Green Version]
  10. Zapata, P.A.; Palza, H.; Cruz, L.S.; Lieberwirth, I.; Catalina, F.; Corrales, T.; Rabagliati, F.M. Polyethylene and poly(ethylene-co-1-octadecene) composites with TiO2 based nanoparticles by metallocenic “in situ” polymerization. Polymer 2013, 54, 2690–2698. [Google Scholar] [CrossRef]
  11. Fonseca, C.; Ochoa, A.; Ulloa, M.T.; Alvarez, E.; Canales, D.; Zapata, P.A. Poly(lactic acid)/TiO2 nanocomposites as alternative biocidal and antifungal materials. Mater. Sci. Eng. C 2015, 57, 314–320. [Google Scholar] [CrossRef]
  12. Wu, F.; Lan, X.; Ji, D.; Liu, Z.; Yang, W.; Yang, M. Grafting polymerization of polylactic acid on the surface of nano-SiO2 and properties of PLA/PLA-grafted-SiO2 nanocomposites. J. Appl. Polym. Sci. 2013, 129, 3019–3027. [Google Scholar] [CrossRef]
  13. Serenko, O.A.; Muzafarov, A.M. Polymer composites with surface modified SiO2 nanoparticles: Structures, properties, and promising applications. Polym. Sci. Ser. C 2016, 58, 93–101. [Google Scholar] [CrossRef]
  14. Ha, S.-W.; Weitzmann, M.N.; Beck, G.R., Jr. Chapter 4—Dental and Skeletal Applications of Silica-Based Nanomaterials A2—Subramani, Karthikeyan. In Nanobiomaterials in Clinical Dentistry; Ahmed, W., Hartsfield, J.K., Eds.; William Andrew Publishing: Norwich, NY, USA, 2013; pp. 69–91. [Google Scholar]
  15. Gao, X.; Wachs, I.E. Titania–silica as catalysts: Molecular structural characteristics and physico-chemical properties. Catal. Today 1999, 51, 233–254. [Google Scholar] [CrossRef]
  16. Pabón, E.; Retuert, J.; Quijada, R.; Zarate, A. TiO2–SiO2 mixed oxides prepared by a combined sol–gel and polymer inclusion method. Microporous Mesoporous Mater. 2004, 67, 195–203. [Google Scholar] [CrossRef]
  17. Pabón, E.; Retuert, J.; Quijada, R. Synthesis of mixed silica–titania by the sol–gel method using polyethylenimine: Porosity and catalytic properties. J. Porous Mater. 2007, 14, 151–158. [Google Scholar]
  18. Milne, C.J.; Penfold, T.J.; Chergui, M. Recent experimental and theoretical developments in time-resolved X-ray spectroscopies. Coord. Chem. Rev. 2014, 277–278, 44–68. [Google Scholar]
  19. Niltharach, A.; Kityakarn, S.; Worayingyong, A.; Thienprasert, J.T.; Klysubun, W.; Songsiriritthigul, P.; Limpijumnong, S. Structural characterizations of sol–gel synthesized TiO2 and Ce/TiO2 nanostructures. Phys. B Condens. Matter 2012, 407, 2915–2918. [Google Scholar] [CrossRef]
  20. Kim, W.B.; Choi, S.H.; Lee, J.S. Quantitative Analysis of Ti−O−Si and Ti−O−Ti Bonds in Ti−Si Binary Oxides by the Linear Combination of XANES. J. Phys. Chem. B 2000, 104, 8670–8678. [Google Scholar] [CrossRef]
  21. Matsuo, S.; Sakaguchi, N.; Wakita, H. Pre-edge features of Ti K-edge X-ray absorption near-edge structure for the local structure of sol-gel titanium oxides. Anal. Sci. 2005, 21, 805–809. [Google Scholar] [CrossRef] [Green Version]
  22. Alias, S.S.; Mohamad, A.A. ZnO: Effect of pH on the Sol–Gel Process. In Synthesis of Zinc Oxide by Sol–Gel Method for Photoelectrochemical Cells; SpringerBriefs in Materials; Springer: Singapore, 2014; pp. 9–25. [Google Scholar]
  23. Kim, T.G.; An, G.S.; Han, J.S.; Hur, J.U.; Park, B.G.; Choi, S.-C. Synthesis of Size Controlled Spherical Silica Nanoparticles via Sol-Gel Process within Hydrophilic Solvent. J. Korean Ceram. Soc. 2017, 54, 49–54. [Google Scholar] [CrossRef] [Green Version]
  24. Karkare, M.M. Choice of precursor not affecting the size of anatase TiO2 nanoparticles but affecting morphology under broader view. Int. Nano Lett. 2014, 4, 111. [Google Scholar] [CrossRef] [Green Version]
  25. Klysubun, W.; Tarawarakarn, P.; Thamsanong, N.; Amonpattaratkit, P.; Cholsuk, C.; Lapboonrueng, S.; Chaichuay, S.; Wongtepa, W. Upgrade of SLRI BL8 beamline for XAFS spectroscopy in a photon energy range of 1–13 keV. Radiat. Phys. Chem. 2020, 175, 108145. [Google Scholar] [CrossRef]
  26. Ravel, B.; Newville, M. THENA, ARTEMIS, HEPHAESTUS: Data analysis for X-ray absorption spectroscopy using IFEFFIT. J. Synchrotron Radiat. 2005, 12, 537–541. [Google Scholar] [CrossRef] [Green Version]
  27. Schraml-Marth, M.; Walther, K.L.; Wokaun, A.; Handy, B.E.; Baiker, A. Porous silica gels and TiO2/SiO2 mixed oxides prepared via the sol-gel process: Characterization by spectroscopic techniques. J. Non-Cryst. Solids 1992, 143, 93–111. [Google Scholar] [CrossRef]
  28. Degirmenci, V.; Erdem, Ö.F.; Ergun, O.; Yilmaz, A.; Michel, D.; Uner, D. Synthesis and NMR Characterization of Titanium and Zirconium Oxides Incorporated in SBA-15. Top. Catal. 2008, 49, 204–208. [Google Scholar] [CrossRef]
  29. Dirken, P.J.; Smith, M.E.; Whitfield, H.J. 17O and 29Si Solid State NMR Study of Atomic Scale Structure in Sol-Gel-Prepared TiO2-SiO2 Materials. J. Phys. Chem. 1995, 99, 395–401. [Google Scholar] [CrossRef]
  30. Zeitler, V.A.; Brown, C.A. The Infrared Spectra of Some Ti–O–Si, Ti–O–Ti and Si–O–Si Compounds. J. Phys. Chem. 1957, 61, 1174–1177. [Google Scholar] [CrossRef]
  31. Praveen, P. Structural, Functional and Optical Characters of TiO2 Nanocrystallites: Anatase and Rutile Phases. St.Joseph’s J. Humanit. Sci. 2019, 6, 43–53. [Google Scholar]
  32. Chen, Y.-F.; Lee, C.-Y.; Yeng, M.-Y.; Chiu, H.-T. The effect of calcination temperature on the crystallinity of TiO2 nanopowders. J. Cryst. Growth 2003, 247, 363–370. [Google Scholar] [CrossRef]
  33. Vasquez, J.; Lozano, H.; Lavayen, V.; Lira-Cantú, M.; Gómez-Romero, P.; Santa Ana, M.A.; Benavente, E.; Gonzalez, G. High-yield preparation of titanium dioxide nanostructures by hydrothermal conditions. J. Nanosci. Nanotechnol. 2009, 9, 1103–1107. [Google Scholar] [CrossRef] [Green Version]
  34. Yamashita, H.; Kawasaki, S.; Ichihashi, Y.; Harada, M.; Takeuchi, M.; Anpo, M.; Stewart, G.; Fox, M.A.; Louis, C.; Che, M. Characterization of Titanium−Silicon Binary Oxide Catalysts Prepared by the Sol−Gel Method and Their Photocatalytic Reactivity for the Liquid-Phase Oxidation of 1-Octanol. J. Phys. Chem. B 1998, 102, 5870–5875. [Google Scholar] [CrossRef]
  35. Vives, S.; Meunier, C. Influence of the synthesis route on sol–gel SiO2–TiO2 (1:1) xerogels and powders. Ceram. Int. 2008, 34, 37–44. [Google Scholar] [CrossRef]
  36. Yu, H.-F.; Wang, S.-M. Effects of water content and pH on gel-derived TiO2–SiO2. J. Non-Cryst. Solids 2000, 261, 260–267. [Google Scholar] [CrossRef]
  37. Yoldas, B.E. Formation of titania-silica glasses by low temperature chemical polymerization. J. Non-Cryst. Solids 1980, 38, 81–86. [Google Scholar] [CrossRef]
  38. Alias, S.S.; Ismail, A.B.; Mohamad, A.A. Effect of pH on ZnO nanoparticle properties synthesized by sol–gel centrifugation. J. Alloys Compd. 2010, 499, 231–237. [Google Scholar] [CrossRef]
  39. Wu, L.; Wu, Y.; Lü, Y. Self-assembly of small ZnO nanoparticles toward flake-like single crystals. Mater. Res. Bull. 2006, 41, 128–133. [Google Scholar] [CrossRef]
  40. Etacheri, V.; Di Valentin, C.; Schneider, J.; Bahnemann, D.; Pillai, S.C. Visible-light activation of TiO2 photocatalysts: Advances in theory and experiments. J. Photochem. Photobiol. C Photochem. Rev. 2015, 25, 1–29. [Google Scholar] [CrossRef] [Green Version]
  41. Li Hsiung, T.; Paul Wang, H.; Wang, H.C. XANES studies of photocatalytic active species in nano TiO2–SiO2. Radiat. Phys. Chem. 2006, 75, 2042–2045. [Google Scholar] [CrossRef]
  42. Rossi, T.C.; Grolimund, D.; Nachtegaal, M.; Cannelli, O.; Mancini, G.F.; Bacellar, C.; Kinschel, D.; Rouxel, J.R.; Ohannessian, N.; Pergolesi, D.; et al. X-ray absorption linear dichroism at the Ti K-edge of anatase TiO2 single crystals. Phys. Rev. B 2019, 100, 245207. [Google Scholar] [CrossRef] [Green Version]
  43. Farges, F.; Brown, G.E.; Rehr, J.J. Ti K-edge XANES studies of Ti coordination and disorder in oxide compounds: Comparison between theory and experiment. Phys. Rev. B 1997, 56, 1809–1819. [Google Scholar] [CrossRef]
  44. Davis, R.J.; Liu, Z. Titania−Silica: A Model Binary Oxide Catalyst System. Chem. Mater. 1987, 9, 2311–2324. [Google Scholar] [CrossRef]
  45. Sharma, A.; Varshney, M.; Park, J.; Ha, T.-K.; Chae, K.-H.; Shin, H.-J. XANES, EXAFS and photocatalytic investigations on copper oxide nanoparticles and nanocomposites. RSC Adv. 2015, 5, 21762–21771. [Google Scholar] [CrossRef]
  46. Khemthong, P.; Photai, P.; Grisdanurak, N. Structural properties of CuO/TiO2 nanorod in relation to their catalytic activity for simultaneous hydrogen production under solar light. Int. J. Hydrogen Energy 2013, 38, 15992–16001. [Google Scholar] [CrossRef]
  47. Loshmanov, A.A.; Sigaev, V.N.; Khodakovskaya, R.Y.; Pavlushkin, N.M.; Yamzin, I.I. Small-angle neutron scattering on silica glasses containing titania. J. Appl. Crystallogr. 1974, 7, 207–210. [Google Scholar] [CrossRef]
  48. Sandstrom, D.R.; Lytle, F.W.; Wei, P.S.P.; Greegor, R.B.; Wong, J.; Schultz, P. Coordination of Ti in TiO2–SiO2 glass by X-ray absorption spectroscopy. J. Non-Cryst. Solids 1980, 41, 201–207. [Google Scholar] [CrossRef]
  49. Neurock, M.; Manzer, L.E. Theoretical insights on the mechanism of alkene epoxidation by H2O2 with titanium silicalite. Chem. Commun. 1996, 10, 1133–1134. [Google Scholar] [CrossRef]
  50. Scherk, C.G.; Ostermann, A.; Achterhold, K.; Iakovleva, O.; Nazikkol, C.; Krebs, B.; Knapp, E.W.; Meyer-Klaucke, W.; Parak, F.G. The X-ray absorption spectroscopy Debye-Waller factors of an iron compound and of met-myoglobin as a function of temperature. Eur. Biophys. J. 2001, 30, 393–403. [Google Scholar]
  51. Wellenreuther, G.; Parthasarathy, V.; Meyer-Klaucke, W. Towards a black-box for biological EXAFS data analysis. II. Automatic BioXAS Refinement and Analysis (ABRA). J. Synchrotron Radiat. 2010, 17, 25–35. [Google Scholar] [CrossRef] [Green Version]
  52. Wu, D.; Zhang, Y.; Yuan, L.; Zhang, M.; Zhou, W. Viscoelastic interfacial properties of compatibilized poly(ε-caprolactone)/polylactide blend. J. Polym. Sci. Part B Polym. Phys. 2010, 48, 756–765. [Google Scholar] [CrossRef]
  53. Gregg, S.J.; Sing, K.S.W. Adsorption, Surface Area and Porosity Auflage, 2nd ed.; John Wiley & Sons, Ltd: Hoboken, NJ, USA; Academic Press: London, UK, 1982. [Google Scholar]
  54. Zhang, X.; Zhang, F.; Chan, K.-Y. Synthesis of titania–silica mixed oxide mesoporous materials, characterization and photocatalytic properties. Appl. Catal. A Gen. 2005, 284, 193–198. [Google Scholar] [CrossRef]
  55. Van Grieken, R.; Aguado, J.; López-Muñoz, M.J.; Marugán, J. Sol-Gel Titania and Titania-Silica Mixed Oxides Photocatalysts. Solid State Phenom. 2010, 162, 221–238. [Google Scholar] [CrossRef]
  56. Kosuge, K. Titanium-Containing Porous Silica Prepared by a Modified Sol−Gel Method. J. Phys. Chem. B 1999, 103, 3563–3569. [Google Scholar] [CrossRef]
Figure 1. Experimental procedures for preparing TixSiy oxide particles.
Figure 1. Experimental procedures for preparing TixSiy oxide particles.
Polymers 14 02729 g001
Figure 2. Schematic illustration of the experimental setup for photocatalytic process.
Figure 2. Schematic illustration of the experimental setup for photocatalytic process.
Polymers 14 02729 g002
Figure 3. Wavelength–absorbance curves of methylene blue (MB) aqueous solutions.
Figure 3. Wavelength–absorbance curves of methylene blue (MB) aqueous solutions.
Polymers 14 02729 g003
Figure 4. Calibration curve of methylene blue (MB) aqueous solutions.
Figure 4. Calibration curve of methylene blue (MB) aqueous solutions.
Polymers 14 02729 g004
Figure 5. 29Si-NMR patterns of (a) SiO2 and (b) Ti50Si50 oxide.
Figure 5. 29Si-NMR patterns of (a) SiO2 and (b) Ti50Si50 oxide.
Polymers 14 02729 g005
Figure 6. FTIR spectra of TiO2, SiO2, and TixSiy oxides synthesized at pH 8.0, 9.0, and 10.0.
Figure 6. FTIR spectra of TiO2, SiO2, and TixSiy oxides synthesized at pH 8.0, 9.0, and 10.0.
Polymers 14 02729 g006
Figure 7. Proposed probable mechanism for synthesized TixSiy oxides.
Figure 7. Proposed probable mechanism for synthesized TixSiy oxides.
Polymers 14 02729 g007
Figure 8. Low (×10k WD = 7.0–7.5 mm EHT = 3.00 kV) and high (×100k WD = 6.9–7.3 mm EHT = 3.00 kV) magnification SEM images of Ti50Si50 oxides synthesized at various pH values: (a) SiO2, (b) Ti50Si50 pH 8.0, (c) Ti50Si50 pH 9.0, (d) Ti50Si50 pH 10.0.
Figure 8. Low (×10k WD = 7.0–7.5 mm EHT = 3.00 kV) and high (×100k WD = 6.9–7.3 mm EHT = 3.00 kV) magnification SEM images of Ti50Si50 oxides synthesized at various pH values: (a) SiO2, (b) Ti50Si50 pH 8.0, (c) Ti50Si50 pH 9.0, (d) Ti50Si50 pH 10.0.
Polymers 14 02729 g008
Figure 9. XRD patterns of Ti50Si50 oxide synthesized with pH 8.0, 9.0, and 10.0.
Figure 9. XRD patterns of Ti50Si50 oxide synthesized with pH 8.0, 9.0, and 10.0.
Polymers 14 02729 g009
Figure 10. Ti K-edge XANES spectra of Ti50Si50 oxide synthesized with pH 8.0, 9.0, and 10.0.
Figure 10. Ti K-edge XANES spectra of Ti50Si50 oxide synthesized with pH 8.0, 9.0, and 10.0.
Polymers 14 02729 g010
Figure 11. Ti K-edge XANES spectra of Ti50Si50 oxide synthesized at pH 8.0, 9.0, and 10.0.
Figure 11. Ti K-edge XANES spectra of Ti50Si50 oxide synthesized at pH 8.0, 9.0, and 10.0.
Polymers 14 02729 g011
Figure 12. Particle size distributions measured on a Zetasizer Nano ZS of SiO2 and Ti50Si50 oxide synthesized at various pH values: (a) SiO2, (b) Ti50Si50 pH 8.0, (c) Ti50Si50 pH 9.0, (d) Ti50Si50 pH 10.0.
Figure 12. Particle size distributions measured on a Zetasizer Nano ZS of SiO2 and Ti50Si50 oxide synthesized at various pH values: (a) SiO2, (b) Ti50Si50 pH 8.0, (c) Ti50Si50 pH 9.0, (d) Ti50Si50 pH 10.0.
Polymers 14 02729 g012
Figure 13. Nitrogen adsorption–desorption isotherms of SiO2, TiO2, and Ti50Si50 oxide synthesized at various pH values.
Figure 13. Nitrogen adsorption–desorption isotherms of SiO2, TiO2, and Ti50Si50 oxide synthesized at various pH values.
Polymers 14 02729 g013
Figure 14. FTIR spectra of TiO2, SiO2, and TixSiy oxide synthesized at pH 9.0 with different Ti/Si ratios.
Figure 14. FTIR spectra of TiO2, SiO2, and TixSiy oxide synthesized at pH 9.0 with different Ti/Si ratios.
Polymers 14 02729 g014
Figure 15. Low (×10k WD = 7.3 mm EHT = 3.00 kV) and high (×100k WD = 6.9–7.3 mm EHT = 3.00 kV) magnification SEM images of TiO2 and TixSiy oxide synthesized at pH 9.0 with different Ti/Si ratios: (a) TiO2, (b) Ti70Si30, (c) Ti50Si50, (d) Ti40Si60.
Figure 15. Low (×10k WD = 7.3 mm EHT = 3.00 kV) and high (×100k WD = 6.9–7.3 mm EHT = 3.00 kV) magnification SEM images of TiO2 and TixSiy oxide synthesized at pH 9.0 with different Ti/Si ratios: (a) TiO2, (b) Ti70Si30, (c) Ti50Si50, (d) Ti40Si60.
Polymers 14 02729 g015
Figure 16. XRD patterns of SiO2, TiO2, and TixSiy oxide synthesized at pH 9.0 with different Ti/Si ratios.
Figure 16. XRD patterns of SiO2, TiO2, and TixSiy oxide synthesized at pH 9.0 with different Ti/Si ratios.
Polymers 14 02729 g016
Figure 17. Ti K-edge XANES spectra of TiO2 and TixSiy oxide synthesized at pH 9.0 with different Ti/Si ratios.
Figure 17. Ti K-edge XANES spectra of TiO2 and TixSiy oxide synthesized at pH 9.0 with different Ti/Si ratios.
Polymers 14 02729 g017
Figure 18. The experimental EXAFS data of TixSiy oxide synthesized at pH 9.0 with different Ti/Si ratios in R-space within the intermission of 1.1–3 Å.
Figure 18. The experimental EXAFS data of TixSiy oxide synthesized at pH 9.0 with different Ti/Si ratios in R-space within the intermission of 1.1–3 Å.
Polymers 14 02729 g018
Figure 19. Particle size distributions measured on a Zetasizer Nano ZS of TiO2 and TixSiy oxide synthesized at pH 9.0 with different Ti/Si ratios: (a) TiO2, (b) Ti70Si30, (c) Ti50Si50, (d) Ti40Si60.
Figure 19. Particle size distributions measured on a Zetasizer Nano ZS of TiO2 and TixSiy oxide synthesized at pH 9.0 with different Ti/Si ratios: (a) TiO2, (b) Ti70Si30, (c) Ti50Si50, (d) Ti40Si60.
Polymers 14 02729 g019
Figure 20. Nitrogen adsorption–desorption isotherms of SiO2, TiO2, and TixSiy oxide synthesized at pH 9.0 with different Ti/Si ratios.
Figure 20. Nitrogen adsorption–desorption isotherms of SiO2, TiO2, and TixSiy oxide synthesized at pH 9.0 with different Ti/Si ratios.
Polymers 14 02729 g020
Figure 21. Concentration of methylene blue (MB) due to absorption of PLA film and PLA/Ti50Si50 oxide (synthesized at pH 8.0, 9.0, and 10.0) films under UV irrigation.
Figure 21. Concentration of methylene blue (MB) due to absorption of PLA film and PLA/Ti50Si50 oxide (synthesized at pH 8.0, 9.0, and 10.0) films under UV irrigation.
Polymers 14 02729 g021
Figure 22. Concentration of methylene blue (MB) due to absorption of PLA film and PLA/TixSiy oxide (synthesized at pH 9.0 with different Ti/Si ratios) films under UV irrigation.
Figure 22. Concentration of methylene blue (MB) due to absorption of PLA film and PLA/TixSiy oxide (synthesized at pH 9.0 with different Ti/Si ratios) films under UV irrigation.
Polymers 14 02729 g022
Table 1. Moles of components in the preparation of Ti50Si50 oxide synthesized at various pH values.
Table 1. Moles of components in the preparation of Ti50Si50 oxide synthesized at various pH values.
SamplesTTIP (mol)TEOS (mol)C2H5OH (mol)HCl/NH4OH (mol)H2O (mol)
SiO2-0.1201.8890.060 b0.440
Ti50Si50 pH 8.00.0110.1091.8890.025 b0.440
Ti50Si50pH 9.00.0110.1091.8890.060 b0.440
Ti50Si50 pH 10.00.0110.1091.8890.075 b0.440
TiO20.120-4.2930.081 a0.702
a moles of HCl, b moles of NH4OH.
Table 2. Moles of components in TixSiy oxide synthesized at pH 9.0 with different Ti/Si ratios.
Table 2. Moles of components in TixSiy oxide synthesized at pH 9.0 with different Ti/Si ratios.
SamplesTTIP (mol)TEOS (mol)C2H5OH (mol)NH4OH/HCl (mol)H2O (mol)
SiO2-0.1201.8890.060 b0.440
Ti70Si300.0200.1001.8890.060 b0.440
Ti50Si500.0110.1091.8890.060 b0.440
Ti40Si600.0080.1121.8890.060 b0.440
TiO20.120-4.2930.081 a0.702
a moles of HCl, b moles of NH4OH.
Table 3. Structural parameters (obtained from fittings of Ti K-edge EXAFS data) of first two coordination shells around Ti atom of Ti and Si in Ti50Si50 oxide synthesized at various pH values and TixSiy oxide synthesized at pH 9.0 with different Ti/Si ratios.
Table 3. Structural parameters (obtained from fittings of Ti K-edge EXAFS data) of first two coordination shells around Ti atom of Ti and Si in Ti50Si50 oxide synthesized at various pH values and TixSiy oxide synthesized at pH 9.0 with different Ti/Si ratios.
SampleShellBondΔE (eV)CNR(Å)σ2R-Factor
Ti50Si50
pH 8.0
1Ti-O4.11 ± 3.344.14 ± 0.181.99 ± 0.020.0053 ± 0.00130.0176
2Ti-Si1.27 ± 0.442.79 ± 0.030.0018 ± 0.0019
Ti50Si50
pH 9.0
1Ti-O2.72 ± 2.034.07 ± 0.121.83 ± 0.010.0042 ± 0.00080.0082
2Ti-Si1.00 ± 0.132.77 ± 0.020.0020 ± 0.0020
Ti50Si50
pH 10.0
1Ti-O6.70 ± 3.294.08 ± 0.191.94 ± 0.020.0049 ± 0.00240.0158
2Ti-Si1.52 ± 0.222.81 ± 0.030.0013 ± 0.0031
Ti70Si301Ti-O6.68 ± 3.234.15 ± 0.191.90 ± 0.020.0051 ± 0.00140.0198
2Ti-Si0.97 ± 0.212.800 ± 0.040.0006 ± 0.0050
Ti50Si501Ti-O2.72 ± 2.034.07 ± 0.121.83 ± 0.010.0042 ± 0.00080.0082
2Ti-Si1.00 ± 0.132.77 ± 0.020.0020 ± 0.0020
Ti40Si601Ti-O6.54 ± 2.274.09 ± 0.131.82 ± 0.010.0048 ± 0.00100.0103
2Ti-Si1.19 ± 0.032.79 ± 0.030.0023 ± 0.0017
Table 4. Percentage of atomic and atomic ratio of Ti and Si in TiO2, SiO2, and Ti50Si50 oxide synthesized at various pH values and TixSiy oxide synthesized at pH 9.0 with different Ti/Si ratios.
Table 4. Percentage of atomic and atomic ratio of Ti and Si in TiO2, SiO2, and Ti50Si50 oxide synthesized at various pH values and TixSiy oxide synthesized at pH 9.0 with different Ti/Si ratios.
SampleAtomic (%)Atomic Ratio of Ti/Si
SiTi
SiO21000-
Ti50Si50 pH 8.054.3945.610.84
Ti50Si50 pH 9.047.8752.131.09
Ti50Si50 pH 10.09.8293.189.49
Ti70Si3029.7970.212.36
Ti50Si5047.8752.131.09
Ti40Si6060.1939.810.66
TiO20100-
Table 5. Particle sizes of SiO2, TiO2, and Ti50Si50 oxide synthesized at various pH values, and TixSiy oxide synthesized at pH 9.0 with different Ti/Si ratios.
Table 5. Particle sizes of SiO2, TiO2, and Ti50Si50 oxide synthesized at various pH values, and TixSiy oxide synthesized at pH 9.0 with different Ti/Si ratios.
SampleZetasizer Nano ZSSEM
Particle Size, d
(nm)
Polydispesity
Index, PDI
Value of Average Diameter, dn
(nm)
SiO2147.80.248144.2 ± 11.3
Ti50Si50 pH 8.0639.30.62741.5 ± 16.7
Ti50Si50 pH 9.0136.20.239135.4 ± 12.3
Ti50Si50 pH 10.0397.00.322114.2 ± 24.2
Ti70Si30573.10.630149.3 ± 15.5
Ti50Si50136.20.239135.4 ± 12.3
Ti40Si60825.50.907131.8 ± 13.3
TiO240.30.49027.8 ± 6.3
Table 6. Specific surface area, pore volume, and mean pore size of TiO2, SiO2, and Ti50Si50 oxide synthesized at various pH values, and TixSiy oxide synthesized at pH 9.0 with different Ti/Si ratios.
Table 6. Specific surface area, pore volume, and mean pore size of TiO2, SiO2, and Ti50Si50 oxide synthesized at various pH values, and TixSiy oxide synthesized at pH 9.0 with different Ti/Si ratios.
SampleSpecific Surface Area, SBET (m2g−1)Pore Volume, Vp
(cm3g−1)
Mean Pore Diameter,
dp (nm)
SiO2116.901.0651.08
TixSiy pH 8.0177.020.112.56
TixSiy pH 9.0225.680.335.85
TixSiy pH 10.062.750.4227.31
Ti70Si30569.071.4210.96
Ti50Si50225.680.335.85
Ti40Si6068.340.3621.97
TiO274.070.2312.75
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Teamsinsungvon, A.; Ruksakulpiwat, C.; Amonpattaratkit, P.; Ruksakulpiwat, Y. Structural Characterization of Titanium–Silica Oxide Using Synchrotron Radiation X-ray Absorption Spectroscopy. Polymers 2022, 14, 2729. https://doi.org/10.3390/polym14132729

AMA Style

Teamsinsungvon A, Ruksakulpiwat C, Amonpattaratkit P, Ruksakulpiwat Y. Structural Characterization of Titanium–Silica Oxide Using Synchrotron Radiation X-ray Absorption Spectroscopy. Polymers. 2022; 14(13):2729. https://doi.org/10.3390/polym14132729

Chicago/Turabian Style

Teamsinsungvon, Arpaporn, Chaiwat Ruksakulpiwat, Penphitcha Amonpattaratkit, and Yupaporn Ruksakulpiwat. 2022. "Structural Characterization of Titanium–Silica Oxide Using Synchrotron Radiation X-ray Absorption Spectroscopy" Polymers 14, no. 13: 2729. https://doi.org/10.3390/polym14132729

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