Next Article in Journal
Systematic Trends in Hybrid-DFT Computations of BaTiO3/SrTiO3, PbTiO3/SrTiO3 and PbZrO3/SrZrO3 (001) Hetero Structures
Next Article in Special Issue
Study of Interacting Heisenberg Antiferromagnet Spin-1/2 and 1 Chains
Previous Article in Journal
Polarons in Rock-Forming Minerals: Physical Implications
Previous Article in Special Issue
On Three “Anomalous” Measurements of Nonlinear QPC Conductance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Physicochemical Properties of Ti3+ Self-Doped TiO2 Loaded on Recycled Fly-Ash Based Zeolites for Degradation of Methyl Orange

by
Iván Supelano García
1,
Carlos Andrés Palacio Gómez
2,*,
Marc H. Weber
3,
Indry Milena Saavedra Gaona
1,
Claudia Patricia Castañeda Martínez
4,
José Jobanny Martínez Zambrano
4,
Hugo Alfonso Rojas Sarmiento
4,
Julian Andrés Munevar Cagigas
5,
Marcos A. Avila
5,
Carlos Rettori
5,6,
Carlos Arturo Parra Vargas
1 and
Julieth Alexandra Mejía Gómez
2
1
Grupo Física de Materiales, Universidad Pedagógica y Tecnológica de Colombia (UPTC), Avenida Central del Norte 39-115, Tunja 150003, Boyacá, Colombia
2
Grupo GIFAM, Universidad Antonio Nariño, Carrera 7 #21-84, Tunja 150001, Boyacá, Colombia
3
Institute of Materials Research, Washington State University, Pullman, WA 99164-2711, USA
4
Grupo de Catálisis-UPTC, Universidad Pedagógica y Tecnológica de Colombia (UPTC), Avenida Central del Norte 39-115, Tunja 150003, Boyacá, Colombia
5
Centro de Ciências Naturais e Humanas, Universidade Federal do ABC (UFABC), Santo André 09210-580, SP, Brazil
6
Instituto de Física “Gleb Wataghin”, Universidade Estadual de Campinas (UNICAMP), Campinas 13083-859, SP, Brazil
*
Author to whom correspondence should be addressed.
Condens. Matter 2022, 7(4), 69; https://doi.org/10.3390/condmat7040069
Submission received: 29 September 2022 / Revised: 3 November 2022 / Accepted: 15 November 2022 / Published: 18 November 2022
(This article belongs to the Special Issue New Advances in Condensed Matter Physics)

Abstract

:
The extensive production of coal fly ash by coal combustion is an issue of concern due to its environmental impact. TiO2-zeolite composites were synthesized, at low cost, using recycled coal fly ash from a local thermoelectric power plant to produce the zeolite using the hydrothermal method. TiO2 was loaded by means of the impregnation method using ethanol and titanium isopropoxide between 8.7 and 49.45 wt% TiO2. The samples were characterized by X-ray diffraction, Raman, electron spin resonance, high-resolution transmission electron microscopy, N2 adsorption-desorption, doppler broadening of annihilation radiation, and diffuse reflectance techniques, and the photocatalytic activity of the composites was evaluated according to the degradation of methyl orange under UV light. The results show that TiO2 crystallizes in the anatase phase with a Ti3+ oxidation state, without post-treatment. TiO2 particles were located within the pores of the substrate and on its surface, increasing the surface area of the composites in comparison with that of the substrates. Samples with TiO2 at 8.7 and 25 wt% immobilized on hydroxysodalite show the highest degradation of methyl orange among all studied materials, including the commercial TiO2 Degussa P25 under UV light.

1. Introduction

Pollution by organic compounds originating from industrial waste waters results in a major threat to human health and aquatic life. To solve this issue, the metal oxide-based semiconductor photocatalysts are considered as a solution due to their capacity to promote oxidation–reduction processes under electromagnetic radiation [1,2]. Among the photocatalysts, the TiO2 is one of the most extensively studied materials due to its physicochemical properties, low cost, abundance, low toxicity, strong oxidation ability, ease of preparation, and stability [3]. To enhance the photocatalytic properties of TiO2 via light irradiation, procedures, including doping with metal and nonmetal ions and the insertion of intrinsic defects, such as Ti3+, to introduce states within the TiO2 band gap, have been developed to produce modified TiO2 nanoparticles [3]. However, TiO2 exhibits some undesired characteristics, such as a low adsorption capacity, a high aggregation of nanoparticles, and difficult reclamation, which reduce its photocatalytic activity [4]. Particle suspension and immobilized TiO2 systems have been designed to overcome the above limitations [2,5]. The immobilization of TiO2 particles on a larger particulate substrate is expected to be the most efficient photocatalysis system from an economical and performance viewpoint [5]. Substrates favoring the adsorption of pollutants, accelerating the photocatalytic rate, can avoid the agglomeration of TiO2 particles, facilitate its recovery from an aqueous medium, and can significantly reduce the recombination of electron-hole pairs in TiO2 [4]. Substrates with a porous structure, such as clays, activated carbon, zeolites, and silica, have been used, resulting in better photocatalytic performance [2]. Efficient dye degradation by TiO2 loaded on aluminosilicates depends on the adsorption processes, the intraparticle diffusion of dye molecules, light penetration, higher surface areas, adsorbed hydroxyl groups, dye and catalyst surface potential, dye concentration, the crystalline structure, the hydrophobicity of the substrate, and the type of Si/Al ratio, in the case of zeolites, among other factors [4,5,6,7,8]. In regards to the zeolites, commercially available [6,9,10,11,12,13] and natural [4,14,15] types are the most common employed for use as substrates for the loading of TiO2 particles.
On the other hand, the large-scale production of those systems increases the cost due to the fact that they may require the use of specific templates or surfactants using high temperature treatments [5]. A low cost material used as substrate to immobilize TiO2 particles is the zeolite synthesized from coal fly ash (CFA). CFA is a fine powder produced during coal combustion as a by-product; its chemical composition comprises amorphous Si and Al, with crystalline phases of quartz, mullite, and iron oxides, among others [16,17,18]. CFA is considered a pollutant, and to reduce its environmental impact, CFA has been used as a Portland cement replacement in cement-based products, as a precursor for geopolymers, and as a raw material for aluminosilicate synthesis [17,19,20,21,22,23,24]. Aluminosilicates, such as zeolite from CFA, can be obtained by alkaline activation using hydrothermal treatment [17,25,26,27]. Synthesis parameters of the hydrothermal treatment, including temperature, activation period, and CFA properties, including chemical composition and SiO2/Al2O3 ratio, are factors affecting the resulting products [17].
According to the revised literature, most of the research uses commercial and natural zeolites to load TiO2 particles. However, fewer papers reported the CFA-based zeolite to support TiO2 [5,28].
In previous works, coal fly ash from a local thermoelectric power plant was activated with NaOH by the hydrothermal method, and the effects of the concentration of NaOH, time of activation, and temperature of the synthesis on the formation of zeolite were evaluated [17,29]. In this work, two of those samples were selected to immobilize Ti3+ self-doped TiO2 on zeolite by the impregnation method, in which the amount of TiO2 was varied between 8.7 and 49.45 wt%. The photocatalytic activity of the composites was evaluated using methyl orange (MO) in an aqueous medium under UV radiation.

2. Results

2.1. Properties of CFA

Figure 1a depicts the SEM micrograph of CFA; it shows the morphology in which the major particles are spherical in shape, corresponding to cenospheres. Minor particles form agglomerates that correspond to carbonaceous particles. Figure 1b shows the X-ray diffraction pattern of CFA, which was indexed with quartz, mullite, maghemite, and hematite as major crystalline phases, with PDF cards: 01-078-2315, 01-083-1881, 00-039-1346, and 00-024-0072, respectively. The broad XRD signal at 2θ < 35° indicates the presence of amorphous content. The chemical composition of CFA, determined by X-ray fluorescence, shows that the CFA contains mainly SiO2 (62.57 (0.45)%), Al2O3 (24.62 (0.50)%), Fe2O3 (5.77 (0.82)%), K2O (1.48 (0.04)%), TiO2 (1.37 (0.05)%), CaO (1.27 (0.07)%), MgO (0.53 (0.04)%), and Na2O (0.46 (0.51)%), with an SiO2/Al2O3 ratio of 2.5.

2.2. Structural Properties of Synthesized TiO2 and Substrates

Figure 2 shows the XRD patterns for TiO2, and the H and C substrates. All diffraction peaks for TiO2 were indexed with a I 41/amd space group of anatase (ICSD 24276). The XRD data of the substrates was analyzed by Rietveld refinement, and the identified crystalline phases are listed in Table 1.
Hydroxysodalite and zeolite P1 correspond to zeolite materials and are the major phases identified in H substrate. Cancrinite is a zeolite material and is the major phase identified in C substrate. According to the DRX results, it can be concluded that the alkaline activation of CFA by NaOH and posterior heat treatment allows for the obtaining of zeolite materials.

2.3. Properties of TiO2-Zeolite Composites

Figure 3a,b shows the XRD patterns for the HT and CT composites. In both HT and CT composites, the Bragg reflections of the anatase TiO2 phase and the Bragg reflections of the major phases of the substrates can be distinguished.
For HT composites, the intensity ratio I(101)TiO2/I(211)H of the major reflection of anatase (101), located at around of 25.4°, and hydroxysodalite (211), located at around of 24.8°, was calculated; correspondingly, for CT composites, the intensity ratio I(101)TiO2/I(101)C of the major reflection of anatase (101), located at around of 25.4°, and cancrinite (101), located at around of 19°, was calculated (Figure 4a). Those ratios tend to increase when the amount of TiO2 increases in both the CT and HT composites. These results suggest that the incorporation of TiO2 as anatase was successful on both the H and C substrates. For TiO2 anatase, in both CT and HT systems, the a and c lattice parameters were calculated from the (200) and (101) reflections, respectively, using the equations:
λ = 2dhklsin(θ)
1/d2hkl = (h2 + k2)/a2 + l2/c2
where λ is the wavelength of the incident X-rays, dhkl is the interplanar distance for a tetragonal system, h, k and l are the Miller indices, and a and c are the lattice parameters.
Figure 4b–d depicts the lattice parameters and cell volume of TiO2 for the HT and CT composites. The results show that lattice parameter a remains constant in both the CT and HT composites. The lattice parameter c and cell volume V exhibit a slight variation for 8.7, 25, and 49.45 wt% TiO2 in comparison with the other samples. The crystallite size for immobilized TiO2 in all the samples was calculated using the Williamson–Hall method, which considers the contribution of crystallite size (βD) and microstrain (βS) to the broadening of a peak, (βhkl) as [30]:
βhkl cosθ = 4εsinθ + Kλ/D
where βhkl is the full width at half maximum, K is a constant, λ is the wavelength of the X-ray, ε is the microstrain, and D is the crystallite size; the latter can be calculated from the analysis of the plot of βhklcosθ vs. 4sinθ. The crystallite size values are listed in Table 2; these values are larger in the HT composites than in the CT composites, indicating that the substrate H favors the growth of TiO2.
The structure of the anatase in all samples was verified by Raman measurements. Figure 5a,b shows the Raman spectra for the HT and CT composites, respectively; for comparison, the Raman spectra of synthesized TiO2 is shown. All samples depict the characteristic five peaks of the D4h19 I 41/amd space group of the anatase around 144, 197, 399, 519, and 638 cm−1, corresponding to Eg(1), Eg(2), B1g(1), an overlap of B1g(2) with the A1g and Eg(3) optical modes, respectively [31]. The majority of the samples exhibit a long-range crystalline order, except for CT-8.7. For both the HT and CT composites, a smooth variation of the Raman shift and the linewidth of the most intense Raman peak around 143 cm−1, in comparison with the values of synthesized TiO2, are depicted (Figure 5c,d). These values are similar in both CT and HT for each of the values of the immobilized TiO2, except for 25 wt%. According to the literature, the change in the peak position and linewidth are related to defects in the stoichiometry, the quantum size effects, the presence of secondary phases, and the influence of the substrate [32,33,34]. In this work, the results suggest that the change in the peak position and linewidth are mainly due to TiO2 quantity.
Figure 6a,b show the ESR spectra at room temperature between 500 G and 550 G for the substrates, composites, and synthesized TiO2. The spectra for H and C substrates show a characteristic resonance for a g value of 2.124 and 2.135, respectively; these are related to the presence of Cu2+ in the substrate matrices [35]. For synthesized TiO2, a resonance at g = 2.000 is observed; this is related to the Ti3+ electronic configuration due to oxygen vacancies [36,37,38,39,40]. For both the HT and CT composites, the contributions to the ESR spectra from the substrate and TiO2 related to the Ti3+ state are observed (Figure 6c,d). Table 2 lists the g value of the resonance, ΔH that corresponds to the distance between peaks (line width), and the relative number of spins calculated by integrating the ESR signal dI/dH twice and comparing this result with a standard sample of 0.11% of KCl with 366 × 1013 spin/cm.
According to the relative number of spins, the Ti3+ ESR signal is higher in HT composites than in CT composites; thus, the H substrate favor the formation of Ti3+ states in the TiO2 rather than C substrate. The ESR signal tends to increase when the TiO2 amount grows in both the HT and CT composites. The synthesis of TiO2 with oxygen vacancies is one of the most important research topics in the chemistry of metal oxides; therefore, a wide variety of methods have been developed to produce self-modified TiO2 materials [41]. Thus, it is worth noting that the samples produced in this work under the method described above allow the synthesis of TiO2 with the formation of oxygen vacancies, without post-treatment. Narrow ΔH values—around 5 G reported for TiO2 at low temperatures—are mainly attributed to Ti3+ ions in the bulk, rather than on the surface of the TiO2 particles [42]. In this work, the HT-8.7, HT-25, and HT-33.25 samples show a smaller ΔH value, which may indicate a larger contribution of the Ti3+ state in the bulk than in the other samples. The presence of Ti3+ suggests that it contributes to the observed variation in the Raman shift and the Raman linewidth, as a stoichiometry defect.
Figure 7 shows selected TEM images of the HT and CT composites with 8.7, 25, and 41.3 wt% of TiO2. The results show that the TiO2 particles are distributed in the zeolite surface, with sizes ranging from 3.85 nm, as in HT-25, Figure 7b, to sizes larger than 50 nm, as in HT-8.7, Figure 7a, possibly explained by agglomeration processes. The lattice spacing of around 0.256 nm assigned to (101) is found in HT-41.3, Figure 7c, and CT-41.3, Figure 7f, samples. In CT composites, the TiO2 particles are found in a wide range of sizes, ranging from 14 nm, as in CT-8.7, Figure 7d, to sizes larger than 33 nm. Lattice spacings of around 0.335 nm, 0.254 nm, and 0.208 nm, assigned to (101), (103), and (200), corresponding to anatase, were found, as shown in Figure 7d–f. From these images, it is observed that the interface between the TiO2 particles and the zeolite substrates varies gradually. It is apparent that the TiO2 particles were successfully incorporated onto the substrate, forming the TiO2-zeolite composites. The nonhomogeneous particle size of TiO2, especially those with a particle size smaller than 20 nm, makes an additional contribution to the observed variation in the Raman shift and Raman linewidth, as a quantum effect.
The N2 adsorption-desorption isotherms for synthesized TiO2 and the substrates, HT composites, and CT composites are shown in Figure 8a–c, respectively. The curves for all samples exhibit a similar shape, which is characteristic of type IV, according to the IUPAC classification, associated with mesoporous materials. The hysteresis loops are similar for C, H, and CT/HT composites characteristics of type H3, associated with lamellar pore structures with slit- and wedge-shaped pores. For synthesized TiO2, the hysteresis loop is characteristic of type H2, associated with cylindrical and spherical pores. The surface area, SBET, was calculated using a BET equation from isotherms curves, the average pore diameter was calculated by the BJH method using the desorption branch, and the pore volume was calculated at a single point of saturation. Table 2 summarizes the results. The surface area of all CT composites is higher than that of the C substrate. The surface area in the HT composites is higher than that of H substrate except for HT-25 and HT-33.25. In both composite systems the larger surface area was obtained for a loading of 41.3 wt% TiO2. The average pore diameter decreases in both the CT and HT composites in comparison with that of the C and H substrates, respectively. The total pore volume decreases in HT composites in comparison with that of the H substrate, whereas the total pore volume increases in the CT composites in comparison with that of C substrate, except for in the CT-49.45 composite. There is a reduction in the average pore diameter in both the CT and HT composites as compared to that of the C and T substrates, respectively, probably due to pore blocking caused by TiO2.
Figure 9a,b shows the positron annihilation line shape S as a function of incident positron energy in the HT and CT composites, respectively; for comparison purposes, the response in the synthesized TiO2 and the substrates is shown. The vertical line separates the surface and bulk region; the annihilation of the positrons in the bulk and/or the annihilation of the p-Ps (para-positronium) in open volumes or defects take place between 1 and 5 keV. The average of the S parameter between 1 and 5 keV is listed in Table 3. The S parameter is higher in the substrates than in the composites and synthesized TiO2, and is higher in the H than in the C substrate. In these samples, the S parameter is related to the interaction of positrons with pores. According to synthesized TiO2, the value of S can be attributed to the oxygen vacancies detected by ESR spectroscopy. For the HT composites, the S parameters in synthesized TiO2 is lower than in other HT composites; for these samples, the S value lies between that of zeolite and synthesized TiO2. This result suggests that any quantity of TiO2 deposits into the pores of the substrate and does not fill its cavities. Regarding the CT composites, the value of S for the composites is close to that of synthesized TiO2. These results suggest that a larger number of pores are filled by TiO2 in the H substrate in comparison with the C substrate, and a larger amount of TiO2 is immobilized on the surface rather than in the pores in the C substrate in comparison with that in the H substrate. This explains the behavior observed concerning the total pore volume in the HT and CT composites.
Figure 10 and Figure 11 show the S-W plot for synthesized TiO2, substrates and both composite systems, respectively. The slope of the S-W plots is useful to observe the change in the nature of the positron trapping defect. Theoretically, S depends linearly on W for each kind of defect. The lowest W value indicates the response from the surface; the higher value of W indicates the response of the bulk. For the HT and CT composites, it can be considered that all points are represented for the same slope that will indicate one kind of defect (mainly pores), since there is no variation in the positron annihilation fraction in the trapping sites in each of the samples; as a function of incident positron energy, the dispersion is attributed to a variation of the pore sizes, which is not homogeneous. Slope values are summarized in Table 3. It is noted that all samples exhibit a small change in the slope, and there is a trend to decreases the slope when TiO2 increases in the HT composites, whereas there is not a clear trend in the CT composites; this indicates that a variation exists in the positron annihilation fraction as a function of the loading of TiO2 particles, which depends on the substrate; thus, it shows the changing size of the trapping sites, and thus, a change in the pores of the samples.
Figure 12 displays the UV-Vis DRS spectra for the HT and CT composites; for comparison, the spectra for synthesized TiO2 and the substrates are shown. For both HT and CT composites, the range of the reflection edge increases in comparison with that of synthesized TiO2. The results suggest that the reflection edge observed in the composites is due to TiO2 and is strongly affected by the interaction between TiO2 and the substrate. The band gap energy (Eg) for the synthesized TiO2 and the composites was calculated by a linear fit of the modified Kubelka–Munk equation as a function of the energy ((F(R)*)1/2 vs. ) and the allowed indirect transition, and the Eg values are listed in Table 2. As can be seen from the table, the calculated Eg for synthesized TiO2 is 2.99 eV, which is lower than 3.2 eV for bulk TiO2 anatase, which may be caused by the Ti3+ electronic configuration and the crystallite size.
Figure 13 shows UV-Vis absorption spectra of the degradation of MO using the samples HT-8.7, HT-25, synthesized TiO2, and commercial Degussa P25 TiO2. The spectra display two absorption bands at 270 nm and 463 nm related to the transition of the aromatic rings and the azo bonds, respectively [43], whose intensity decreases with irradiation time, indicating the degradation of MO.
Figure 14a,b shows the variation in the relative concentration (C/C0) of MO as a function of the reaction time for the HT and CT composites, respectively. The composites labeled as HT-8.7 and HT-25 presented better photocatalytic behavior than other composites and the commercial Degussa P25 TiO2. Thus, after 210 min of photocatalytic reaction, the HT-8.7 and HT-25 composites favored a dye degradation of 84% and 87%, respectively, compared to the 93% and 67% degradation evidenced using synthesized TiO2 and Degussa P25 TiO2, respectively.
On the other hand, nearly 17% of the dye was degraded using the HT-16.85 and HT-33.25 composites; 19% and 30% using HT-41.3 and HT-49.45, respectively; 3% using CT-8.7 and CT-25; 20%, 6%, 8%, and 0% using CT-16.85, CT-41.3, CT-49.45, and CT-6-425-33.25, respectively. Desorption processes shown in any samples, such as HT-33.25 and CT-33.25, can be attributed to a low generation of ˙OH and O 2 · radicals and/or a high electron-hole recombination rate [44].
The total organic carbon analysis was carried out for the reactions involving the materials that evidenced the best photocatalytic behavior. The contents of TOC after 210 min of reaction were 46, 36, 15, and 64% for HT-8.7, HT-25, Degussa P25 TiO2, and synthesized TiO2, respectively. These results indicate that the mineralization of MO using these materials occurs in the following order: synthesized TiO2 > HT-8.7 > HT-25 > commercial TiO2. The higher degradation of MO using synthesized TiO2, followed by the HT composites, can be explained in terms of the presence of Ti states that promote the charge separation process, acting as a holes trap [45]. However, the larger surface area and larger crystallinity of TiO2 aids in the separation of the photogenerated electron and holes [4]. The degradation efficiency exhibited by the HT-8.7 and HT-25 composites can be explained in terms of a larger generation of ˙OH and O 2 · radicals due to a larger irradiated surface area and a low electron-hole recombination rate [44]. Usually, the photodegradation of organic dyes by semiconducting oxides is described by the pseudo first order equation [43,46]:
ln(C0/C) = kt
where C0 is the concentration of the dye at adsorption equilibrium, C is the concentration of the dye at time t, and k is the apparent rate constant. The ln(C0/C) vs. t plot is displayed in Figure 14c for the HT-8.7, HT-25, and synthesized TiO2 samples, in which the values of the rate constant k are 7.60 × 10−3, 7.50 × 10−3, and 7.93 × 10−3, respectively.
According to the results obtained in this work and in the literature [43,44,45,46,47,48,49], the photocatalytic degradation of MO can be described as follows (Figure 15): when TiO2 is loaded onto the substrate surface, it is illuminated by UV light, and the photogenerated electron-hole pair may then migrate to the surface. The electrons on the surface reduce the adsorbed oxygen molecules to O 2 · radicals, and the photogenerated holes migrate toward the surface to oxidize the adsorbed water to form ˙OH species. The O 2 · radicals, in turn, may generate HO2˙ species. Due to the Ti3+ states, this avoids the electron-hole pair recombination and increases the number of O 2 · species generated in the system. The O 2 · , ˙OH, and HO2˙ radicals are very reactive and decompose the MO molecules adsorbed on the surface of the composite; this process is faster than the velocity of the desorption of the MO molecules, allow for the degradation of the dye. The larger photocatalytic activity in HT-8.7 and HT-25 can be explained in terms of a larger crystallinity degree evidenced in the Raman results, a synergic effect between the substrate and the amount of TiO2 that lies on a relatively high surface area, with a partial fill up of substrate pores evidenced by N2 physisorption and DBAR measurements, in comparison with other composites.

3. Materials and Methods

3.1. Materials

CFA with a size smaller than 38 μm were selected by sieving, and were collected from the Sochagota TermoPaipa IV thermoelectric power plant (Boyacá, Colombia). NaOH (PanReac AppliChem 98%) and deionized water were employed to perform the alkaline activation of CFA. Titanium isopropoxide (Alfa Aesar 97+%) and ethanol (Alfa Aesar 96%) were used to synthetize TiO2.

3.2. Properties of CFA

The chemical composition of CFA was determined by using a MagixPro PW-2440 Philips X-ray fluorescence spectrometer. The morphology of CFA was observed by using a Carl Zeiss EVO-MA10 electron microscope at the accelerating voltage of 10 kV. The crystalline phases were identified by using a PANalytical X’Pert’s X-ray diffractometer with λKα-Co = 1.789007 Å.

3.3. Alkaline Activation of CFA by the Hydrothermal Method

Two different samples, with larger conversions of CFA into zeolite and under conditions of a low synthesis temperature and concentration of NaOH, as described in previous works, were selected for use as substrates for TiO2 immobilization [22,35]. Briefly, CFA was activated by the conventional hydrothermal method using a Teflon® vessel placed within a steel reactor and covered by a thermal insulation in order to keep the temperature at a constant value, which was monitored with a K thermocouple with the following synthesis parameters:
(a)
4.2 g of CFA was added in NaOH solution (3 M) at 99.5 °C for 24 h.
(b)
10 g of CFA was added in NaOH solution (4 M) at 95 °C for 24 h.
After the alkaline activation, the solution was filtered, washed with ethanol, filtered again, and dried. In addition, in this work, the as prepared samples were calcined at 500 °C. From a previous procedure, two samples were obtained with zeolite as the major crystalline phase, corresponding to hydroxysodalite, and cancrinite labeled as H and C, respectively; these samples were used as substrates to immobilize the TiO2 particles.

3.4. Synthesis of TiO2-Zeolite Composites

Wetness impregnation of TiO2 over H and C samples was carried out as follows: H and C substrates were placed in ethanol under magnetic stirring at 200 rpm, titanium isopropoxide was added drop by drop with different nominal TiO2 loading at 60 °C, and the solution was kept under magnetic stirring for 6 h. The as-prepared composites were dried at 90 °C overnight and calcined at 425 °C for 2 h. A sample of TiO2 was synthesized following this procedure at 425 °C for 6 h. The obtained composites were labelled according to the scheme of Table 4, in which the HT nomenclature refers to TiO2 (T) impregnated over the substrate with main phase of hydroxysodalite (H) to obtain the composite HT and, the CT label refers to TiO2 (T) impregnated over the substrate with the main phase of cancrinite (C) to obtain the composite CT.

3.5. Characterization of TiO2-Zeolite Composites

The structural characterization was performed by X-ray diffraction (XRD) using a Bruker AXS D8 focus spectrometer with λCu-Kα = 0.15405 nm. The Raman spectra were obtained in a Horiba Yobin-Yvon T64000 dispersive Raman spectrometer using a laser source of 532 nm and a power of 0.001 W over the sample. Electron spin resonance (ESR) spectra were obtained at room temperature in a Bruker EMX 10-2.7 Plus spectrometer. The morphological properties were observed by high resolution transmission electron microscopy (HRTEM) in a Tecnai F20 Super Twin TMP microscope of FEI; low-temperature N2 physisorption measurements were performed using Micromeritics ASAP 2020 equipment, with Doppler broadening annihilation radiation of the positrons, and energy measurements (DBAR) were performed in the slow positron beam facility at Washington State University, USA; the 511 keV annihilation peak was recorded using an ORTEC HPGe detector. The optical properties were studied by ultraviolet-visible diffuse reflectance spectroscopy (DRS) using an UV-VIS Cary 50 VARIAN spectrophotometer equipped with an integrating sphere; BaSO4 was used as the reference material.
The photocatalytic behavior of the materials was studied in the degradation of MO dye (C14H14N3NaO3S). For this purpose, 200 mL of a MO aqueous solution at 25 ppm and 50 mg of photocatalytic material were added in a Pyrex reactor. The suspension was maintained under magnetic stirring at 700 rpm, and it was bubbled with constant air flow. The system was irradiated by employing a low-pressure mercury lamp (254 nm, 4.4 mW) protected with a quartz tube, which was immersed in the suspension. To favor the adsorption-desorption equilibrium, the reaction system was maintained in the dark for 30 min. Then, the irradiation was initiated, and aliquots, which were filtered by a membrane filter, were extracted every 30 min. The samples were analyzed by using an Evolution 300 UV-Vis spectrophotometer. Finally, the residual total organic carbon (TOC) was estimated using a TOC/TN analyzer multi N/C 2100.

4. Conclusions

Alkaline activation of CFA by the hydrothermal method under hydrothermal conditions allows for the obtaining of aluminosilicates with zeolite materials as majority phases, which were then used as substrates to immobilize TiO2, whereby it was possible to produce TiO2-zeolite composites at low cost. The synthesis method permits the obtaining of TiO2 in an anatase phase, with a similar lattice parameter and a long-range of crystalline order for most of the samples. The procedure performed to synthesize the TiO2-composites allows for the obtaining of anatase TiO2 with a Ti3+ state, indicating the presence of oxygen vacancies without subsequent treatment. The textural properties of the composites were improved in comparison with those of the substrates, since, due to TiO2 incorporation, the surface of the substrate was changed. The results verify that TiO2 particles were located on the surface and within the pores of the substrate. On the surface, the particles of TiO2 crystallize in a wide range of particle sizes due to the agglomeration effect. The HT-8.7, HT-25, and synthesized TiO2 showed a degradation of MO in an aqueous medium under UV light, which was higher than that of commercial Degussa P25 TiO2, due to a synergic effect between the substrate and the amount of TiO2.

Author Contributions

Conceptualization, writing, formal analysis, investigation, I.S.G.; writing, investigation, formal analysis, resources, C.A.P.G.; writing, investigation, M.H.W.; investigation, I.M.S.G.; writing, investigation, formal analysis, C.P.C.M.; writing, formal analysis, J.J.M.Z.; resources, H.A.R.S.; writing, conceptualization, formal analysis, J.A.M.C.; resources, supervision, M.A.A.; writing, formal analysis, C.R.; conceptualization, resources, supervision, C.A.P.V.; conceptualization, resources, supervision, J.A.M.G. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financed by the Gobernación de Boyacá (Grant No. 733 Colciencias), Universidad Antonio Nariño, under the project No PI/UAN-2021-687GFM and FAPESP (Grant No. 2017/10581-1).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We acknowledge CEM—Centrais Experimentais Multiusuário—UFABC for its support of this project.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Rafiq, A.; Ikram, M.; Ali, S.; Niaz, F.; Khan, M.; Khan, Q.; Maqbool, M. Photocatalytic degradation of dyes using semiconductor photocatalysts to clean industrial water pollution. J. Ind. Eng. Chem. 2021, 97, 111–128. [Google Scholar] [CrossRef]
  2. Mehta, A.; Mishra, A.; Sharma, M.; Singh, S.; Basu, S. Effect of silica/titania ratio on enhanced photooxidation of industrial hazardous materials by microwave treated mesoporous SBA-15/TiO2 nanocomposites. J. Nanopart. Res. 2016, 18, 209. [Google Scholar] [CrossRef]
  3. Wajid Shah, M.; Zhu, Y.; Fan, X.; Zhao, J.; Li, Y.; Asim, S.; Chuanyi, W. Facile Synthesis of Defective TiO2−x Nanocrystals with High Surface Area and Tailoring Bandgap for Visible-light Photocatalysis. Sci. Rep. 2015, 5, 15804. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Zhang, G.; Song, A.; Duan, Y.; Zheng, S. Enhanced photocatalytic activity of TiO2/zeolite composite for abatement of pollutants. Microporous. Mesoporous. Mater. 2018, 255, 61–68. [Google Scholar] [CrossRef]
  5. Yang, L.; Wang, F.; Hakki, A.; Macphee, D.E.; Liu, P.; Hu, S. The influence of zeolites fly ash bead/TiO2 composite material surface morphologies on their adsorption and photocatalytic performance. Appl. Surf. Sci. 2017, 392, 687–696. [Google Scholar] [CrossRef]
  6. Chong, M.N.; Tneu, Z.Y.; Poh, P.E.; Jin, B.; Aryal, R. Synthesis, characterization and application of TiO2–zeolite nanocomposites for the advanced treatment of industrial dye wastewater. J. Taiwan Inst. Chem. Eng. 2015, 50, 288–296. [Google Scholar] [CrossRef]
  7. Setthaya, N.; Chindaprasirt, P.; Yin, S.; Pimraksa, K. TiO2-zeolite photocatalysts made of metakaolin and rice husk ash for removal of methylene blue dye. Powder Technol. 2017, 313, 417–426. [Google Scholar] [CrossRef]
  8. Alberti, S.; Caratto, V.; Peddis, D.; Belviso, C.; Ferretti, M. Synthesis and characterization of a new photocatalyst based on TiO2 nanoparticles supported on a magnetic zeolite obtained from iron and steel industrial waste. J. Alloy Compd. 2019, 797, 820–825. [Google Scholar] [CrossRef]
  9. Jansson, I.; Suárez, S.; Garcia-Garcia, F.J.; Sánchez, B. Zeolite–TiO2 hybrid composites for pollutant degradation in gas phase. Appl. Catal B 2015, 178, 100–107. [Google Scholar] [CrossRef]
  10. Liu, C.; Zhang, R.; Wei, S.; Wang, J.; Liu, Y.; Li, M.; Liu, R. Selective removal of H2S from biogas using a regenerable hybrid TiO2/zeolite composite. Fuel 2015, 157, 183–190. [Google Scholar] [CrossRef]
  11. Guesh, K.; Mayoral, A.; Marquez-Alvarez, C.; Chebude, Y.; Díaz, I. Enhanced photocatalytic activity of TiO2 supported on zeolites tested in real wastewaters from the textile industry of Ethiopia. Microporous. Mesoporous. Mater. 2016, 225, 88–97. [Google Scholar] [CrossRef]
  12. Yamaguchi, S.; Fukura, T.; Imai, Y.; Yamauara, H.; Yahiro, H. Photocatalytic activities for partial oxidation of a-methylstyrene over zeolite-supported titanium dioxide and the influence of water addition to reaction solvent. Electrochim. Acta 2010, 55, 7745–7750. [Google Scholar] [CrossRef]
  13. Xu, Y.; Langford, C.H. Photoactivity of Titanium Dioxide Supported on MCM41, Zeolite X, and Zeolite Y. J. Phys. Chem. B 1997, 101, 3115–3121. [Google Scholar] [CrossRef]
  14. Mendoza, J.A.; Lee, D.H.; Kang, J.H. Photocatalytic removal of NOx using TiO2-coated zeolite. Environ. Eng. Res. 2016, 3, 291–296. [Google Scholar] [CrossRef]
  15. Sun, Q.; Hu, X.; Zheng, S.; Sun, Z.; Liu, S.; Li, H. Influence of calcination temperature on the structural, adsorption and photocatalytic properties of TiO2 nanoparticles supported on natural zeolite. Powder Technol. 2015, 274, 88–97. [Google Scholar] [CrossRef]
  16. Hefni, Y.; El Zaher, Y.A.; Wahab, M.A. Influence of activation of fly ash on the mechanical properties of Concrete. Constr. Build. Mater. 2018, 172, 728–734. [Google Scholar] [CrossRef]
  17. Supelano, G.I.; Gómez Cuaspud, J.A.; Moreno-Aldana, L.C.; Ortiz, C.; Trujillo, C.A.; Palacio, C.A.; Parra Vargas, C.A.; Mejía Gómez, J.A. Synthesis of magnetic zeolites from recycled fly ash for adsorption of methylene blue. Fuel 2020, 263, 116800. [Google Scholar] [CrossRef]
  18. Sivalingam, S.; Sen, S. Optimization of synthesis parameters and characterization of coal fly ash derived microporous zeolite X. Appl. Surf. Sci. 2018, 455, 903–910. [Google Scholar] [CrossRef]
  19. Hafez, H.; Kurda, R.; Cheung, W.M.; Nagaratnam, B. Comparative life cycle assessment between imported and recovered fly ash for blended cement concrete in the UK. J. Clean. Prod. 2020, 244, 118722. [Google Scholar] [CrossRef]
  20. Gomaa, E.; Gheni, A.A.; Kashosi, C.; El Gawady, A.A. Bond strength of eco-friendly class C fly ash-based thermally cured alkali-activated concrete to portland cement concrete. J. Clean. Prod. 2019, 235, 404–416. [Google Scholar] [CrossRef]
  21. Lan, T.; Guo, S.; Li, X.; Guo, J.; Bai, T.; Zhao, Q.; Yang, W.; Li, P. Mixed precursor geopolymer synthesis for removal of Pb(II) and Cd(II). Mater. Lett. 2020, 274, 127977. [Google Scholar] [CrossRef]
  22. Zhao, X.; Liu, C.; Zuo, L.; Wang, L.; Zhu, Q.; Liu, Y.; Zhou, B. Synthesis and characterization of fly ash geopolymer paste for goal backfill: Reuse of soda residue. J. Clean. Prod. 2020, 260, 121045. [Google Scholar] [CrossRef]
  23. Verrecchia, G.; Cafiero, L.; de Caprariis, B.; Dell’Era, A.; Pettiti, I.; Tuffi, R.; Scarsella, M. Study of the parameters of zeolites synthesis from coal fly ash in order to optimize their CO2 adsorption. Fuel 2020, 276, 118041. [Google Scholar] [CrossRef]
  24. de Castro Amoni, B.; Lima de Freitas, A.D.; Rodrigues Loiola, A.; Barbosa Soares, J.; de Aguiar Soares, S. A method for NaA zeolite synthesis from coal fly ash and its application in warm mix asphalt. Road Mater. Pavement Des. 2019, 20, S558–S567. [Google Scholar] [CrossRef]
  25. Derkowski, A.; Franus, W.; Beran, E.; Czímerová, A. Properties and potential applications of zeolitic materials produced from fly ash using simple method of synthesis. Powder Technol. 2006, 166, 47–54. [Google Scholar] [CrossRef]
  26. Kotova, O.B.; Shabalin, I.N.; Shushkov, D.A.; Kocheva, L.S. Hydrothermal synthesis of zeolites from coal fly ash. Adv. Appl. Ceram. 2016, 115, 152–157. [Google Scholar] [CrossRef] [Green Version]
  27. Gamero-Vega, K.Y.; Medina-Ramírez, A.; Khamkure, S.; Orozco-Núñez, S.I.; Aguilera-González, E.N.; Gamero-Melo, P. Upscaling of W zeolite direct synthesis from coal fly ash and its water adsorption capacity. J. Chem. Technol. Biotechnol. 2019, 94, 3479–3487. [Google Scholar] [CrossRef]
  28. Liu, Z.; Liu, Z.; Cui, T.; Dong, L.; Zhang, J.; Han, L.; Li, J.; Liu, C. Photocatalyst from one-dimensional TiO2 nanowires/synthetic zeolite composites. Mater. Express 2014, 4, 465–474. [Google Scholar] [CrossRef]
  29. Supelano García, I.; Ortíz Otálora, C.A.; Parra Vargas, C.A.; Mejía Gómez, J.A. Assessment of hydrothermal parameters on alkaline activation of fly ashes using a central composite design. J. Met. Mater. Min. 2021, 31, 54–61. [Google Scholar] [CrossRef]
  30. Irfa, H.; Mohamed, R.K.; Anand, S. Microstructural evaluation of CoAl2O4 nanoparticles by Williamson–Hall andsize–strain plot methods. J. Asian Ceram. Soc. 2018, 6, 54–62. [Google Scholar] [CrossRef]
  31. Ohsaka, T.; Izumi, F.; Fujiki, Y. Raman Spectrum of Anatase, TiO2. J. Raman. Spectrosc. 1978, 7, 321–324. [Google Scholar] [CrossRef]
  32. Mazzolini, P.; Russo, V.; Casari, C.S.; Hitosugi, T.; Nakao, S.; Hasegawa, T.; Bassi, A.L. Vibrational–Electrical Properties Relationship in Donor-Doped TiO2 by Raman Spectroscopy. J. Phys. Chem. C 2016, 120, 18878–18886. [Google Scholar] [CrossRef] [Green Version]
  33. Choi, H.C.; Jung, Y.M.; Kim, S.B. Size effects in the Raman spectra of TiO2 nanoparticles. Vib. Spectrosc. 2005, 37, 33–38. [Google Scholar] [CrossRef]
  34. Liu, G.; Han, C.; Pelaez, M.; Zhu, D.; Liao, S.; Likodimos, V.; Ioannidis, N.; Kontos, A.G.; Falaras, P.; Dunlop, P.S.M.; et al. Synthesis, characterization and photocatalytic evaluation of visible light activated C-doped TiO2 nanoparticles. Nanotechnology 2012, 23, 294003. [Google Scholar] [CrossRef] [PubMed]
  35. Leith, I.R.; Leach, H.F. Adsorbate interactions on copper-exchanged X-type zeolite catalysts studied by E.P.R. spectroscopy. Proc. R. Soc. A 1972, 330, 247–263. [Google Scholar]
  36. Zhao, Y.; Zhao, Y.; Shi, R.; Wang, B.; Waterhouse, G.I.N.; Wu, L.Z.; Tung, C.H.; Zhang, T. Tuning Oxygen Vacancies in Ultrathin TiO2 Nanosheets to Boost Photocatalytic Nitrogen Fixation up to 700 nm. Adv. Mater. 2019, 31, 1806482. [Google Scholar] [CrossRef] [PubMed]
  37. Xing, M.; Fang, W.; Nasir, M.; Ma, Y.; Zhang, J.; Anpo, M. Self-doped Ti3+-enhanced TiO2 nanoparticles with a high-performance photocatalysis. J. Catal. 2013, 297, 236–243. [Google Scholar] [CrossRef]
  38. Aronne, A.; Fantauzzi, M.; Imparato, C.; Atzei, D.; De Stefano, L.; D’Errico, G.; Sannino, F.; Rea, I.; Pirozzi, D.; Elsener, B.; et al. Electronic properties of TiO2-based materials characterized by high Ti3+ self-doping and low recombination rate of electron–hole pairs. RSC Adv. 2017, 7, 2373. [Google Scholar] [CrossRef] [Green Version]
  39. Dong, J.; Han, J.; Liu, Y.; Nakajima, A.; Matsushita, S.; Wei, S.; Gao, W. Defective Black TiO2 Synthesized via Anodization for Visible-Light Photocatalysis. ACS. Appl. Mater. Interfaces 2014, 6, 1385–1388. [Google Scholar] [CrossRef]
  40. Fang, W.; Zhou, Y.; Dong, C.; Xing, M.; Zhang, J. Enhanced photocatalytic activities of vacuum activated TiO2 catalysts with Ti3+ and N co-doped. Catal. Today 2016, 266, 188–196. [Google Scholar] [CrossRef]
  41. Su, J.; Zou, X.; Chen, J.S. Self-modification of titanium dioxide materials by Ti3+ and/or oxygen vacancies: New insights into defect chemistry of metal oxides. RSC Adv. 2014, 4, 13979. [Google Scholar] [CrossRef]
  42. Xiong, L.B.; Li, J.L.; Yang, B.; Yu, Y. Ti3+ in the Surface of Titanium Dioxide: Generation, Properties and Photocatalytic Application. J. Nanomater. 2012, 2012, 831524. [Google Scholar] [CrossRef] [Green Version]
  43. Saputera, W.H.; Mul, G.; Hamdy, M.S. Ti3+-containing titania: Synthesis tactics and photocatalytic performance. Catal. Today 2015, 246, 60–66. [Google Scholar] [CrossRef]
  44. Li, W.; Li, D.; Lin, Y.; Wang, P.; Chen, W.; Fu, X.; Shao, Y. Evidence for the Active Species Involved in the Photodegradation Process of Methyl Orange on TiO2. J. Phys. Chem. 2012, 116, 3552–3560. [Google Scholar] [CrossRef]
  45. Ghafoor, S.; Ata, S.; Mahmood, N.; Arshad, S.N. Photosensitization of TiO2 nanofibers by Ag2S with the synergistic effect of excess surface Ti3+ states for enhanced photocatalytic activity under simulated sunlight. Sci. Rep. 2017, 7, 255. [Google Scholar] [CrossRef] [Green Version]
  46. Wang, J.; Yang, P.; Huang, B. Self-doped TiO2−x nanowires with enhanced photocatalytic activity: Facile synthesis and effects of the Ti3+. Appl. Surf. Sci. 2015, 356, 391–398. [Google Scholar] [CrossRef]
  47. Ruirui, L.; Zhijiang, J.; Jing, W.; Jinjun, Z. Mesocrystalline TiO2/sepiolite composites for the effective degradation of methyl orange and methylene blue. Front. Mater. Sci. 2018, 12, 292–303. [Google Scholar]
  48. Qiu, M.; Tian, Y.; Chen, Z.; Yang, Z.; Li, W.; Wang, K.; Wang, L.; Wang, K.; Zhang, W. Synthesis of Ti3+ self-doped TiO2 nanocrystals based on Le Chatelier’s principle and their application in solar light photocatalysis. RSC Adv. 2016, 6, 74376. [Google Scholar] [CrossRef]
  49. Khan, H.; Swati, I.K. Fe3+-doped Anatase TiO2 with d−d Transition, Oxygen Vacancies and Ti3+ Centers: Synthesis, Characterization, UV−vis Photocatalytic and Mechanistic Studies. Ind. Eng. Chem. Res. 2016, 55, 6619–6633. [Google Scholar] [CrossRef]
Figure 1. (a) SEM image of CFA, (b) XRD pattern for CFA indexed with quartz, mullite, maghemite, and hematite.
Figure 1. (a) SEM image of CFA, (b) XRD pattern for CFA indexed with quartz, mullite, maghemite, and hematite.
Condensedmatter 07 00069 g001
Figure 2. XRD pattern for (a) synthesized TiO2 indexed with anatase phase, (b) H substrate indexed with hydroxysodalite, zeolite P1, mullite, and sodium aluminosilicate, (c) C substrate indexed with cancrinite and an aluminosilicate.
Figure 2. XRD pattern for (a) synthesized TiO2 indexed with anatase phase, (b) H substrate indexed with hydroxysodalite, zeolite P1, mullite, and sodium aluminosilicate, (c) C substrate indexed with cancrinite and an aluminosilicate.
Condensedmatter 07 00069 g002
Figure 3. XRD pattern for (a) HT composites (b) CT composites. For comparative purposes, the XRD of C, H, and synthesized TiO2 are depicted.
Figure 3. XRD pattern for (a) HT composites (b) CT composites. For comparative purposes, the XRD of C, H, and synthesized TiO2 are depicted.
Condensedmatter 07 00069 g003
Figure 4. (a) I(101)TiO2/I(211)H and I(101)TiO2/I(101)C intensity ratios. (b) Lattice parameters a and (c) c; and (d) cell volume. The continuous line represents the value for synthesized TiO2.
Figure 4. (a) I(101)TiO2/I(211)H and I(101)TiO2/I(101)C intensity ratios. (b) Lattice parameters a and (c) c; and (d) cell volume. The continuous line represents the value for synthesized TiO2.
Condensedmatter 07 00069 g004aCondensedmatter 07 00069 g004b
Figure 5. Raman shift spectra for (a) HT composites, (b) CT composites, (c) variation of Raman shift, (d) linewidth for the 144 cm−1 peak as a function of TiO2 weight percentage. the continuous line represents the parameters for synthesized TiO2.
Figure 5. Raman shift spectra for (a) HT composites, (b) CT composites, (c) variation of Raman shift, (d) linewidth for the 144 cm−1 peak as a function of TiO2 weight percentage. the continuous line represents the parameters for synthesized TiO2.
Condensedmatter 07 00069 g005
Figure 6. ESR spectra between 500 and 5500 G for (a) HT composites; (b) CT composites. ESR spectra around 3375 G for (c) HT composites; (d) CT composites.
Figure 6. ESR spectra between 500 and 5500 G for (a) HT composites; (b) CT composites. ESR spectra around 3375 G for (c) HT composites; (d) CT composites.
Condensedmatter 07 00069 g006
Figure 7. HRTEM images for (a) HT-8.7, (b) HT-25, (c) HT-41.3, (d) CT-8.7, (e) CT-25, and (f) CT-41.3. Red line represents the interface between zeolite and TiO2 in Figure 7a and the lattice spacings for Figure 7b–f.
Figure 7. HRTEM images for (a) HT-8.7, (b) HT-25, (c) HT-41.3, (d) CT-8.7, (e) CT-25, and (f) CT-41.3. Red line represents the interface between zeolite and TiO2 in Figure 7a and the lattice spacings for Figure 7b–f.
Condensedmatter 07 00069 g007
Figure 8. N2 adsorption isotherm of (a) synthesized TiO2 substrates; (b) HT composites; (c) CT composites.
Figure 8. N2 adsorption isotherm of (a) synthesized TiO2 substrates; (b) HT composites; (c) CT composites.
Condensedmatter 07 00069 g008aCondensedmatter 07 00069 g008b
Figure 9. Variation with positron beam energy for the S parameter for (a) HT composites; (b) CT composites. Dashed line separates the surface and bulk region.
Figure 9. Variation with positron beam energy for the S parameter for (a) HT composites; (b) CT composites. Dashed line separates the surface and bulk region.
Condensedmatter 07 00069 g009
Figure 10. S-W plot for synthesized TiO2, H, and C substrates. Continuous line represents the average slope.
Figure 10. S-W plot for synthesized TiO2, H, and C substrates. Continuous line represents the average slope.
Condensedmatter 07 00069 g010
Figure 11. S-W plot for HT and CT composites. Continuous line represents the average slope.
Figure 11. S-W plot for HT and CT composites. Continuous line represents the average slope.
Condensedmatter 07 00069 g011
Figure 12. UV-Vis diffuse reflectance spectra for (a) HT composites; (b) CT composites.
Figure 12. UV-Vis diffuse reflectance spectra for (a) HT composites; (b) CT composites.
Condensedmatter 07 00069 g012
Figure 13. UV-Vis spectra of MO solution photoreduction by (a) HT-8.7, (b) HT-25, (c) synthesized, and (d) commercial TiO2.
Figure 13. UV-Vis spectra of MO solution photoreduction by (a) HT-8.7, (b) HT-25, (c) synthesized, and (d) commercial TiO2.
Condensedmatter 07 00069 g013aCondensedmatter 07 00069 g013b
Figure 14. Time-dependent degradation of MO by photocatalytic activity of (a) HT and (b) CT composites. (c) The ln(C0/C) vs. t plot.
Figure 14. Time-dependent degradation of MO by photocatalytic activity of (a) HT and (b) CT composites. (c) The ln(C0/C) vs. t plot.
Condensedmatter 07 00069 g014aCondensedmatter 07 00069 g014b
Figure 15. Scheme of photocatalytic mechanism for MO degradation.
Figure 15. Scheme of photocatalytic mechanism for MO degradation.
Condensedmatter 07 00069 g015
Table 1. Crystalline phases identified in H and C substrates.
Table 1. Crystalline phases identified in H and C substrates.
SubstratePhaseChemical FormulaSpace
Group
ICSD Card No.Percentage
of Phase (%)
HHydroxysodaliteNa8(AlSiO4)6(OH)2P -4 3 n6084059 (1.30)
Zeolite P1H24Al6Na6O44Si10I -4955021 (1.87)
MulliteAl2.272O4.864Si0.728Pbam10080518 (0.80)
AluminosilicateH3.92Al1.92Na2O12Si3.08I -42d0672102 (0.12)
CCancriniteNa8(AlSiO4)6(CO3)(H2O)2P 632033468 (0.92)
AluminosilicateNaAlSiO4P 3243318132 (0.72)
Table 2. Morphological and optical parameters for HT and CT composites.
Table 2. Morphological and optical parameters for HT and CT composites.
SampleTiO2
Crystallite
Size (nm) *
Eg
(eV) **
G ***ΔH
(G) ***
Relative
Number
of Spins
(×1015) ***
SBET
(m2/g) ****
Desorption
Average
Pore
Diameter
(nm) ****
Total
Pore
Volume
(cm3/g)
(×10−2) ****
HT-8.773.83.31 (2)1.999 (4)3.20.2215154
HT-2564.53.31 (4)2.000 (4)3.81.3310122
HT-33.2576.63.84 (2)1.998 (7)3.91.0610143
HT-41.388.63.84 (2)1.999 (8)8.70.401683
HT-49.4542.32.85 (1)1.998 (5)7.81.551383
H----------11244
TiO246.82.99 (2)2.000 (4)8.32.432843
CT-8.748.13.50 (3)------6242
CT-2553.53.64 (2)1.999 (3)8.30.278132
CT-33.2574.13.92 (7)2.000 (6)8.30.268122
CT-41.331.53.44 (7)1.999 (4)8.30.069152
CT-49.4533.63.68 (3)1.999 (4)7.80.047122
C----------6252
* Calculated from XRD patterns using the Williamson–Hall method. ** Calculated from UV-Vis DRS spectra using the modified Kubelka–Munk equation. *** Calculated from ESR data. **** Calculated from N2 adsorption-desorption isotherms.
Table 3. Representative slope of the S-W plot and <S> parameter between 1–5 keV.
Table 3. Representative slope of the S-W plot and <S> parameter between 1–5 keV.
SampleSlope<S>SampleSlope<S>
HT-8.7−1.32 (24)0.46219 (140)CT-8.7−1.81 (33)0.45419 (137)
HT-25−1.79 (30)0.46434 (137)CT-25−1.62 (18)0.45558 (137)
HT-33.25−1.76 (18)0.45775 (138)CT-33.25−1.95 (23)0.45445 (136)
HT-41.3−1.68 (14)0.45756 (137)CT-41.3−1.90 (12)0.45321 (136)
HT-49.45−2.27 (14)0.45917 (137)CT-49.45−1.86 (16)0.45324 (136)
H−1.88 (14)0.46975 (138)C−2.24 (13)0.46679 (139)
Table 4. Labeling of samples showing the amount of immobilized TiO2.
Table 4. Labeling of samples showing the amount of immobilized TiO2.
TiO2 Loaded over H SubstrateTiO2 Loaded over C SubstrateTiO2 (wt%)
HT-8.7CT-8.78.70
HT-25CT-2525.00
HT-33.25CT-33.2533.25
HT-41.3CT-41.341.30
HT-49.45CT-49.4549.45
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Supelano García, I.; Palacio Gómez, C.A.; Weber, M.H.; Saavedra Gaona, I.M.; Castañeda Martínez, C.P.; Martínez Zambrano, J.J.; Rojas Sarmiento, H.A.; Munevar Cagigas, J.A.; Avila, M.A.; Rettori, C.; et al. Physicochemical Properties of Ti3+ Self-Doped TiO2 Loaded on Recycled Fly-Ash Based Zeolites for Degradation of Methyl Orange. Condens. Matter 2022, 7, 69. https://doi.org/10.3390/condmat7040069

AMA Style

Supelano García I, Palacio Gómez CA, Weber MH, Saavedra Gaona IM, Castañeda Martínez CP, Martínez Zambrano JJ, Rojas Sarmiento HA, Munevar Cagigas JA, Avila MA, Rettori C, et al. Physicochemical Properties of Ti3+ Self-Doped TiO2 Loaded on Recycled Fly-Ash Based Zeolites for Degradation of Methyl Orange. Condensed Matter. 2022; 7(4):69. https://doi.org/10.3390/condmat7040069

Chicago/Turabian Style

Supelano García, Iván, Carlos Andrés Palacio Gómez, Marc H. Weber, Indry Milena Saavedra Gaona, Claudia Patricia Castañeda Martínez, José Jobanny Martínez Zambrano, Hugo Alfonso Rojas Sarmiento, Julian Andrés Munevar Cagigas, Marcos A. Avila, Carlos Rettori, and et al. 2022. "Physicochemical Properties of Ti3+ Self-Doped TiO2 Loaded on Recycled Fly-Ash Based Zeolites for Degradation of Methyl Orange" Condensed Matter 7, no. 4: 69. https://doi.org/10.3390/condmat7040069

Article Metrics

Back to TopTop