Next Article in Journal / Special Issue
Epitaxial Growth of Ni-Mn-Ga on Al2O3(112¯0) Single-Crystal Substrates by Pulsed Laser Deposition
Previous Article in Journal
Functional Polyacrylate Textile Coatings with N,N-Diethyl-3-methylbenzamide (DEET) Immobilized on Zirconia, Alumina and Silica Sorbents
Previous Article in Special Issue
Thin Epitaxial Ionic Fluoride Films for Electronics Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

TPA and PET Photo-Degradation by Heterogeneous Catalysis Using a (Al2O3)0.75TiO2 Coating

by
Mónica A. Camacho-González
1,
Alberto Hernández-Reyes
1,
Aristeo Garrido-Hernández
2,
Octavio Olivares-Xometl
3,
Natalya V. Likhanova
4 and
Irina V. Lijanova
1,*
1
Instituto Politécnico Nacional, Centro de Innovación e Investigación Tecnológica, Cerrada Cecati S/N, Colonia Santa Catarina de Azcapotzalco, Ciudad de Mexico 02250, Mexico
2
Departamento de Materiales, Universidad Autónoma Metropolitana, Av. San Pablo 420, Col. Nueva el Rosario, Azcapotzalco, Ciudad de Mexico 02128, Mexico
3
Facultad de Ingeniería Química, Benemérita Universidad Autónoma de Puebla, Av. San Claudio y 18 Sur, Ciudad Universitaria, Col. San Manuel, Puebla 72570, Mexico
4
Instituto Mexicano del Petróleo, Eje Central Norte Lázaro Cárdenas No. 152, Col. San Bartolo Atepehuacan, Del. Gustavo A. Madero, Ciudad de Mexico 07730, Mexico
*
Author to whom correspondence should be addressed.
Surfaces 2025, 8(2), 34; https://doi.org/10.3390/surfaces8020034
Submission received: 25 March 2025 / Revised: 14 May 2025 / Accepted: 16 May 2025 / Published: 21 May 2025
(This article belongs to the Special Issue Surface Engineering of Thin Films)

Abstract

:
The combination of the catalytic properties of Al2O3/TiO2 formed an efficient system to degrade the ubiquitous pollutants TPA and PET. The coating (Al2O3)0.75TiO2 was characterized by X-ray diffraction. Stainless steel disks with photo-catalyst coating were placed transversely in a 3.0 L vertical glass reactor with ascending airflow for supplying oxygen to the reaction medium and visible light lamps for photo-activation. The analysis of the coating homogeneity, morphology and particle size distribution of the TiO2 coatings and (Al2O3)0.75TiO2 system were confirmed by SEM. Optical properties and band-gap energy were calculated by using the Tauc equation. UV–Vis spectrophotometry (UV–Vis) and chemical oxygen demand (COD) were the quantitative techniques to measure the reduction in the initial TPA and PET concentrations.

1. Introduction

Polyethylene terephthalate (PET) is the most widely used thermoplastic polymer in various areas due to its application diversity, as well as to its strength, lightness, elasticity, and transparency properties [1]. In contrast, its degradation rate is low by acid or basic hydrolysis, methanolysis, glycolysis, and acetolysis, which are chemical methods that can depolymerize PET into monomers such as terephthalic acid, ethylene glycol diacetate, colorants, and special additives [1,2]. Terephthalic acid (TPA) or 1,4-benzene dicarboxylic acid is the final product of successive oxidation reactions of p-xylene, serving as a raw material in the manufacture of PET and other plastics, dyes, perfumes, pesticides, medicines, etc. [3,4]. Wastewater discharged from the production of PET and TPA has high concentrations of different carboxylic acids like para-toluic, benzoic, acetic, phthalic, isophthalic, and an excess of terephthalic acid [5]. Due to the high stability of the aromatic structure, terephthalic acid is a ubiquitous pollutant in sediments, natural waters, soils, aquatic organisms, etc. [6]; TPA resists degradation by aerobic, anaerobic, and combined biological means and electrocoagulation [5,7,8,9]. In addition, TPA exerts an inhibitory effect on the growth of microorganisms, drastically decreasing the biomass during the degradation processes and extending the treatment period [10]. The refining process includes a neutralization reaction with alkalis, which produces an excessive amount of salt and substantially reduces the efficiency of or sometimes prevents anaerobic treatment [11,12,13].
Recently, advanced oxidation processes (AOPs) have become more common for the effective decomposition of organic matter in wastewater, including photocatalytic oxidation, ozonation, and Fenton oxidation (H2O2-Fe) [14,15,16]. AOPs proceed by generating super-reactive free radicals, which oxidize organic pollutants into less harmful or short-chain structures, which could be treated by biological means [17]. It should be emphasized that the symmetrical position of substituent groups in the aromatic ring of TPA improves its structural stability, which is widely used to detect •OH radicals generated in photocatalytic processes aiming to evaluate the effectiveness of various photocatalysts [18]. The degradation of TPA, desirably, must happen by means of non-toxic solid heterogeneous photocatalysts at ambient temperature and pressure, with the removal of multiple compounds from wastewater [19,20]. Therefore, the main challenge of AOPs is the design of an ideal photocatalyst with speedy-oxidizing power under radiation, high chemical and photo stability, low cost, and high availability [21]. Environmentally friendly TiO2 is the most promising UV-photocatalyst with wide bandgap (anatase, 3.2 eV) and acceptable recombination of photogenerated electrons. The activation of TiO2 by visible light could be promoted by doping it with different elements such as nickel, platinum, copper, etc., and metal oxides such as ZnO, WO3, and Al2O3 to generate heterostructures that function as traps for photogenerated electrons, thus reducing their recombination and shifting absorption to the visible spectrum [22,23,24,25,26].
The enhancement of the photocatalytic activity of mixed oxides under UV and visible radiation was shown by Al2O3–TiO2 in suspension or as thin films, in comparison with austere TiO2 without doping [27,28,29]. The combination of the catalytic properties of TiO2 and the relevant mechanical resistance, high chemical and thermal stability, and adsorption capacity of Al2O3 gives the system potential reusability without losing efficiency to degrade pollutants [30,31]. In specialized wastewater treatment, immobilized photocatalysts could help skip the filtration step at the end of the process; however, reducing the surface/volume ratio causes slow mass transfer and less availability of catalytically active sites [32,33]. Regardless of the nature of the photocatalyst, the design of reactors for the photocatalytic degradation of organic pollutants in aqueous solution is an area in constant development [34,35]. However, there are three fundamental limitations of a photocatalytic process: (a) mass transport, (b) efficient propagation of photons to the photocatalyst, and (c) the amount of molecular oxygen dissolved in a solution [36,37,38].
In this study, the purpose was to evaluate the efficiency of Al2O3–TiO2 photocatalyst coatings deposited on 304 stainless steel (304 SS) (Ciudad de Mexico, Mexico) perforated disks by the dip-coating technique to degrade TPA and PET by heterogeneous photocatalysis. The disks with photocatalyst coating were placed transversely in a 3.0 L vertical glass reactor with ascending air flow to supply oxygen to the reaction medium and visible light lamps for photo-activation. UV–Vis spectrophotometry and chemical oxygen demand (COD) were the quantitative techniques employed to measure the reduction in the initial TPA and PET concentrations.

2. Materials and Methods

2.1. Coating Preparation by the Sol-Gel Route

The (Al2O3)0.75TiO2 photocatalyst (molar ratio) was obtained by the sol-gel and dip-coating techniques from precursor solutions of aluminum sulfate [Al2(SO4)3]·18H2O (Meyer, Ciudad de Mexico, Mexico, 98%) and titanium tert-butoxide (TBT) (Sigma Aldrich, St. Louis, MI, USA, 97%). According to the methodology of [39], a 1 M NaOH (Meyer, Ciudad de Mexico, Mexico, ≥97%) solution was added dropwise to 95 mL of a 0.1 M [Al2(SO4)3] 12H2O solution until reaching pH = 10; the obtained colloidal suspension was kept under stirring at 70 °C for 5 h. Afterward, the suspension was washed three times with an ethanol/water (50:50 v/v) solution to remove SO42− ions. The wet solid was redispersed in 50 mL of ethanol (Sigma Aldrich, St. Louis, MI, USA, 99.9%); this suspension was aged for 24 h. Four hours before the end of the aging period of the aluminum suspension, 25 mL of a 0.3 M HNO3 (Meyer, Ciudad de Mexico, Mexico, 65%) solution were added dropwise to 30 mL of a TBT/ethanol (15% v/v) solution; the suspension was stirred for 2 h to achieve the completion of the hydrolysis–condensation reaction. Hereafter, the aluminum suspension was added to the forming [TiO(OH)2]n gel and kept under vigorous stirring until a white milky colloidal suspension appeared. Meanwhile, five circular AISI 304 SS disks with diameters of 104 mm and orifice diameters of 2 mm were washed with non-ionic detergent soap, followed by an acetone washing in a sonication bath at 400 kHz for 20 min to eliminate oil traces, and stored in isopropyl alcohol for 1 h. Subsequently, the disks were immersed in the milky colloidal suspension of aluminum–titanium hydroxides and withdrawn at 40 mm/min. After each immersion, the disks were dried at 80 °C for 1 h; in the last of the three depositions after the treatment at 80 °C, they were treated at 180 °C for 1 h and, finally, annealed at 700 °C for 5 h. Five 304 SS disks with TiO2 coatings were obtained by the same technique without mixing with the 0.1 M [Al2(SO4)3]·12H2O solution.

2.2. Structural Characterization Measurements

The structural study of the (Al2O3)0.75TiO2 photocatalyst was carried out first, with a Bruker D8 Advance diffractometer using Cu Kα radiation (1.54184 Å), room temperature, and 2θ angle ranging from 20° to 80° with 0.02 s−1 pulse to determine the crystallinity and crystal phase of the produced coating. The analysis of the morphology and size of the particles was performed by employing a field emission scanning electron microscope (FE-SEM) Hitachi SU5000; for the EDS analysis, a coupled Bruker Quantax XFlash 6/60 was used to investigate the compositional aspects of the coating. The optical characteristics of the (Al2O3)0.75TiO2 coating were examined by a Perkin Elmer model Lambda 35 UV–Vis spectrophotometer from 200 to 900 nm wavelength to register absorbance values of the coating scratch.

2.3. Photocatalytic Degradation

A 3 L cylindrical reactor was designed to measure all the degradation reactions. It was made of 4 mm thick low-Fe borosilicate glass, with internal diameter and height of 111.2 and 312 mm and a top lid of the same material and diameter with 4 inlets; ascending air flow and mechanical stirring were employed. Through the top central inlet, a CPVC pipe was passed, which in addition to protect the propeller rod, served as a guide for the disks with (a) TiO2 or (b) (Al2O3)0.75TiO2 coatings (~200 cm2 of coating area and 44 ± 5 mg of photocatalyst per substrate). Five 104 mm diameter AISI 304 SS circular disks (alternating 2-mm holes on the entire surface) with central hole diameter of 19 mm and with-coating thickness of 1.5 mm were separated from each other by 3/4 inch CPVC tubes 40 mm in length and with an alternating 15° cutting angle at their two ends with respect to the reactor transverse axis (see Figure 1a). Through one of the lid inlets, a Polypropylene (PP) pipe (Rotoplas, Ciudad de Mexico, Mexico) was introduced to take 25 mL of solution at each time interval to quantify the COD (NMX-AA-030/2-SCFI-2011) in a HACH spectrophotometer, DR2010 (HACH company, Loveland, CO, USA). Through another inlet, a probe was introduced to measure the average pH (9.5 ± 0.2) and temperature (30 ± 0.5 °C). Eight Phillips LED lamps (4 W, 400 lumens, MR16 with emission within a 410–760 nm interval with a maximum peak at around 600 nm) were placed 20 mm from the external wall reactor in pairs at 90 degrees to the center of the reactor for visible light irradiation (see Figure 1b,c). A black box covered the reactor and lamps to avoid external radiation.
In the reactor, 50 mg/L of TPA dissolved in 2.5 L of a 0.1% NaOH solution was homogenized by upward airflow at 5 L/min with mechanical stirring for 30 min in the absence of light to allow adsorption–desorption equilibrium between the surface of the (Al2O3)0.75TiO2 coatings and contaminant. The bubble displacement velocity by aeration-mechanical stirring was 7.26 mm/s from the time it left the diffuser to the time it left the liquid, and the flow regime was laminar (Re = 1035). The system homogeneity state was reached in 5.4 s. Subsequently, 25 mL of solution was extracted with a syringe to determine the initial TPA concentration as the COD (mg/L) immediately after the lamps were lit. Every 2 h, the same volume of solution was collected to determine the COD (mg/L). The same methodology was applied for TPA degradation with TiO2 coatings and PET degradation using (Al2O3)0.75TiO2. Each determination was performed in triplicate. Finally, the procedure was carried out under the same conditions to degrade PET using (Al2O3)0.75TiO2. The degradation experiments were carried out in triplicate.
The efficiency of the degradation of TPA and PET was calculated with Equation (1):
% d e g r a d a t i o n = C O D o C O D t C O D o × 100
where C O D o is the initial concentration as COD of each contaminant before turning on the lamps, and C O D t is the concentration of the contaminant after time t.

3. Results and Discussion

3.1. (Al2O3)0.75TiO2 Coating Characterization

3.1.1. X-Ray Diffraction

The XRD patterns of the synthesized TiO2, Al2O3, and (Al2O3)0.75TiO2 oxide system powders in Figure 2 indicate that the synthesized TiO2 exhibits predominantly the rutile phase, determined by the presence of the (110), (011), (020), (111), (120), (220), (002), (130), and (031) planes at 27.45°, 36.1°, 39.2°, 41.3°, 44.0°, 56.6°, 63°, 64.0°, and 69.06°, and 76.1°, respectively [40]. According to the ICSD chart 98-008-2085, above 500 °C, the first peak of the (110) plane, characteristic of rutile, appears at 27.5°, and near 700 °C, the anatase phase is completely transformed into rutile, being the most predominant stable phase [41].
In the case of the synthesized Al2O3, the presence of the planes (012), (110), (024), (214), and (116) is associated with the α-Al2O3 phase, according to the ICDD chart 01-080-0956. The γ-Al2O3 phase is represented by the (121), (400), and (224) planes indicated in the ICSD chart 98-009-9836. The presence of the two phases is due to the transformation of boehmite AlO(OH) into α-Al2O3 and even γ-Al2O3 was promoted by the synthesis route and heat treatment at 400–700 °C [42,43,44]. The presence of Al2O3 alters the surface energy of TiO2, blocking the nucleation and growth of the rutile phase. This fact explains why in the (Al2O3)0.75TiO2 sample, the anatase phase has total dominance over the rutile phase according to the chart ICSD 98-015-4603, with diffraction peaks (101), (112), and (024) planes of 2θ = 25.28, 37.54, and 62.41, respectively.
The protective layers of chromium oxide and magnetite formed on the surface of 304 SS during the high-temperature treatment (near 800 °C) were detected through the structural analysis of the coating [45]. The Cr atoms could diffuse toward the grain boundaries, forming voids through which the Fe atoms in the 304 SS matrix could migrate and react with oxygen from the atmosphere or residual oxygen in the chromium-rich magnetite system: Fe2+(Fe3+,Cr3+)2O4 [46,47]. Figure 3 shows the diffraction planes of stainless-steel disks in the absence of coating, without heat treatment, and with heat treatment at 700 °C, observing this phenomenon in more detail for 5 h.
Camacho-González et al. analyzed by means of Rietveld refinement the phase composition of a 304 stainless steel substrate coated with Al2O3–TiO2, synthesized under the same conditions as in this work, finding that the presence of Fe and alloying elements in the substrate was 79% of the total present phases: 72% austenite, 4% ferrite, 1% magnetite, and 2% chromium oxide, and the remaining 21% corresponded to the Al2O3–TiO2 coating, where α-Al2O3 (7.4%) was more abundant than γ-Al2O3 (3%). In the case of TiO2, the anatase phase present at 6.9% almost doubled the rutile phase with only 3.7% [39].

3.1.2. Energy Dispersive Spectroscopy: Elemental Analysis

Figure 4 shows the micrographs and mappings of the elemental analysis of TiO2 (area: 0.25 µm2; magnification: 239) and (Al2O3)0.75TiO2 coating (area: 0.12 µm2; magnification: 478) by the energy dispersive X-ray spectroscopy (EDS) technique (resolution: 512 × 444 pixels; voltage: 20.0 kV). 304 SS is composed of approximately 65–75% of Fe and alloying elements in the substrate were confirmed too: 18–20% of chromium and 8–12% of nickel, with traces of manganese (<2%), silicon (<2%), phosphorus (<0.045%), and sulfur (<0.03%), maintaining a low carbon content (<0.03%) [48]. Figure 4a shows the result of the sintering process (700 °C), where Fe (39.89%), Cr (10.4%), and Ni (3.83%) presented migration to the surface, forming the respective oxides on the TiO2 coating surface.
Figure 4b shows the homogeneous distribution of the (Al2O3)0.75TiO2 system, confirming the presence of Al (3.83%) and Ti (2.62%) and the diffusion onto the surface of the formed Fe (32.23%) and Cr (25.15%) oxides. The constant and homogeneous distribution of oxygen in mass percentage from 22.96 to 25.46% was associated with the formation of M-O bonds between the oxygen atoms present in the coating and the surface metal atoms [49]. The molar ratio of 0.75 Al2O3 to 1 TiO2 is equivalent to an atomic ratio of Al/Ti = 1.5, which is verified by the present percentage mass of these two atoms in the coating (3.83%/2.62% = 1.46).

3.1.3. Scanning Electron Microscopy, Morphology and Particle Size

The analysis of the coating homogeneity, morphology, and particle size distribution in the TiO2 coatings and (Al2O3)0.75TiO2 system are shown in Figure 5; in both cases, the surface is homogeneously distributed, with no visible fractures or uncoated areas. For the TiO2 coatings (Figure 5a) on 304 stainless steel, deposits of filamentous microstructures like rice grains are observed whose mean length distribution is 240 ± 102 nm and 54% of the measured filaments are within the interval ranging from 200 to 300 nm (Figure 5a′). Due to the independent nature of the oxides forming the system and the synthesis conditions, two different and independent morphologies are observed for the (Al2O3)0.75TiO2 coating (Figure 5b). The presence of filamentous microstructures is related to TiO2 as previously analyzed, while the formation of spherical agglomerated particles of the (Al2O3)0.75TiO2 coating may be associated with the presence of Al2O3 and the sintering temperature (Figure 5b′). Liu and co-workers [50] synthesized TiO2–SiO2–Al2O3 composite coatings at different sintering temperatures on the surface of Q235 carbon steel by the sol-gel method to improve its corrosion resistance. As the authors increased the sintering temperature above 650 °C, the coating layer was composed of interlocked vine-shaped microstructures; in addition, the appearance of porosity was observed. In the analysis, the various coating morphologies were attributed to the increase in coating components on the surface, among which, besides Al2O3, Fe3O4 and Cr2O3 stood out.
The spherical particle size distribution of the (Al2O3)0.75TiO2 coating was 230 ± 120 nm. It is evident that the heterogeneity of the particles deposited on the surface allowed the formation of porosity and highly rough surfaces that could ensure a greater surface area available for the photocatalytic degradation process. In the sol-gel synthesis, during the formation of the sol, successive hydrolysis and condensation reactions gave rise to fractal aggregates, which successively joined, forming clusters whose expansion process brought the system to the solidification point or gel, changing the viscosity of the medium due to the rapid elimination of solvent. During aging, the phenomenon of syneresis (expulsion of liquid from the pores between clusters) occurred, and the successive thermal treatment caused the deformation of the polymeric network because of the loss of liquid between the pores. The increase in capillary pressure reflected the shrinkage of the network, which resulted in dense agglomerated ceramics with high porosity and high surface area, making these materials ideal catalysts and support matrices [51].

3.1.4. UV–Visible Spectroscopy: Optical Properties and Determination of the Forbidden Band Energy

The synthesis of the (Al2O3)0.75TiO2 system was conceived to have a photocatalyst whose energy absorption range in the electromagnetic spectrum shifted from the UV to visible range, because that Al2O3 effectively reduces the electron–hole recombination and improves the absorption characteristics in the visible spectrum.
In Figure 6, UV–Vis spectroscopy allowed to observe two maximum absorption bands in the UV range, due to the influence of TiO2 particles, between 220 and 230 nm and a weak absorption band between 300 nm and 350 nm. Martínez-Gómez and his team stated that these bands were the result of charge transfer from O2− to Ti4+, which corresponded to the excitation of electrons from the valence band of O2− (2p orbitals) to the conduction band of Ti4+ (3d orbitals), characteristic of the anatase phase [31]. The higher visible light collection capacity between 420–625 nm of the (Al2O3)0.75TiO2 system can be related to the coating surface defects or irregularities: (a) one has a higher surface area/volume ratio, which means more active sites for photocatalytic reactions, (b) the scattering of light into multiple directions enhances the photon absorption and generation of electron-hole pairs, which are essential for photocatalytic reactions, and (c) the effective charge separation enhances the use of the photogenerated charge to degrade pollutants [52,53].
According to Planck’s equation, E = h c λ , where E is the energy of incident radiation, h is Planck’s constant (4.136 × 10−15 eV·s), c is the speed of light (2.998 × 1017 nm/s), and λ is the wavelength (nm), the energy of 5.2 eV (220 nm) relates to the modified energy levels in Al2O3, showing weak energy absorption below 250 nm [54]. The lifetime of charge carriers in the (Al2O3)0.75(TiO2) system is observed by the absorption band between 300 and 380 nm [53].
The Tauc plot of the UV–Vis spectrum of the Al2O3–TiO2 coating reported band gap values of 3.4 and 2.9 eV, corresponding to the anatase and rutile phases of TiO2, respectively, and values of 4.6 and 5.2 eV for α-Al2O3 and γ-Al2O3, respectively [39]. These values are in good agreement with the data obtained from the photon energy calculated for the anatase and rutile phases of TiO2 and alpha and gamma phases of Al2O3 as the interface between these oxides reported in this work.

3.2. Evaluation of the Photocatalytic Efficiency in the Degradation of Terephthalic Acid by the Immobilized (Al2O3)0.75TiO2 System

TPA exhibits characteristic absorption maxima at 190 nm, 241 nm, and 285 nm when dissolved in an acidic medium [55]. The appearance of a weak absorption band between 280 and 290 nm, which shifts to shorter wavelengths, as the polarity of the solvent increases, indicates the presence of a carbonyl group. This type of shift, called “blue shift” or “hypsochromic”, is related to the behavior of the chromophore. Likewise, an absorption band near 260 nm with fine vibrational structure is indicative of an aromatic ring [56]. Spectra in the UV region of aromatic hydrocarbons are characterized by three sets of bands originating from π → π* transitions. Figure 7 shows the TPA spectra from 200 to 300 nm at different photocatalytic degradation breakthrough times. For TPA dissolved in 0.1% NaOH, the absorption maxima occurred at 247 nm, where the band was blue-shifted, due to the basic nature of the solution. The bands at 229 and 242 nm were assigned to the π → π* transitions of the sp2 C=C conjugated bonds of the aromatic ring. Likewise, at 201–202 nm, the characteristic band of the carboxyl group with n → π* transition type was observed.
As the reactions elapsed at time t, the broad band near 247 nm split into two from the first hour of reaction, which could be associated with the binding of the photogenerated OH· radical to the aromatic ring, generating 2-hydroxy-terephthalic acid (Figure 7) [57]. This reaction can be related to the attenuation of the band near 242 nm and the alternating appearance of a band around 235 nm. The band of the carboxyl group did not undergo apparent change and its conversion to CO2 was minimal according to the 28.12% reduction in the absorbance value for this band [58]. The efficiency of photocatalytic oxidation (ZnO 2.5 g/L) of TPA reached 95%, where the operative parameters were: volume = 1.5 L, pH = 9.0, at 30 °C, adding amounts of H2O2 as a scavenger of photogenerated electrons to produce OH- radicals [59]. The same studies reported that the degradation of the subsequent hydroxylation of TPA intermediates led to the cleavage of the aromatic ring, forming short-length carboxylic acids such as oxalic, formic, maleic, fumaric, and acetic that are capable of reacting directly with hydroxyl radicals, finally being mineralized to CO2 [60].
The degradation of TPA by the action of incident light (photolysis) in the absence of disks was 4.04% and, in the presence of disks without coating, treated at 700 °C for 5 h, it was 6.22%; so, the formation of Cr2O3 and Fe3O4 by the sensitization process could promote additional TPA degradation of 1.54%. The degradation efficiency of TPA employing TiO2 coatings on 304 SS steel disks was 28.39%, measured from the remaining concentration by absorbance at 247 nm, and this behavior was corroborated with the residual concentration measurement as COD analysis with 30.63% of degradation, observing a difference between both techniques of 1.24%, but with similar performance (Figure 8a).
The degradation efficiency of TPA with (Al2O3)0.75TiO2 over 8 h, which was the duration of the reaction study, was 42.58%, measured by absorbance at 247 nm, and 46.60%, measured as COD, with a difference of 9.4% (Figure 8b).
Considering the COD measurements on TPA degradation, pseudo-first-order kinetics was observed using the Langmuir–Hinshelwood formalism as shown in the plots of ln C/Co vs. time. The apparent kinetic constants of kapp = 0.047 h−1 for TiO2, and kapp = 0.0697 h−1 for (Al2O3)0.75TiO2 showed that the heterostructure presented higher efficiency than TiO2 under the same operating conditions for 8 h of study (Figure 9).
A band diagram for the (Al2O3)0.75TiO2 catalyst is presented in Figure 10 to understand the catalytic degradation mechanism. The synthesis methodology allowed the formation of a step-gap type heterojunction between TiO2 and Al2O3, where the latter acted as an electron sink (charge carrier) that facilitated electronic transfers from anatase/rutile TiO2 to alumina by delaying the h+-e− recombination time; the photogenerated holes on the rutile/anatase TiO2 surface favored the formation of hydroxyl radicals by dissociation of water molecules and consequently, successive oxidation-reduction reactions could be enhanced to degrade TPA and PET into less complex organic compounds [61,62].
Since TPA is a molecule that helps identify the generated OH radicals, future works will be able to verify not only the transformation of TPA into hydroxy-TPA by measuring fluorescence values during TPA degradation [18,63], but also the formation of less complex organic products using HPLC or GC-MS to validate the proposed degradation mechanism.

3.3. Evaluation of the Photocatalytic Efficiency of PET Degradation by the Immobilized (Al2O3)0.75TiO2 System

The UV–Vis spectroscopy analysis of PET shows only two broad bands at t0 = 0 within the 203–207 nm interval, where the n → π* transitions of carboxyl (-COO-) and -OH- groups present in polyethylene terephthalate are commonly found (Figure 11). The other broad band between 208–213 nm is characterized by the π → π* transitions of the sp2 C=C conjugated bonds of the aromatic ring and the bond occurring with an sp2 carbon.
In PET, the carboxyl groups of terephthalate and OH groups of ethylene glycol are linked together by an ester bond (203–207 nm), while the repeating structural unit has, in addition to ester bonds, Csp3-Csp2 covalent bonds (208–213 nm). The attack of hydroxyl radicals on the ester bonds progresses the depolymerization reaction, causing the linking of the n repeating units of PET and the formation of n units of mono 2-hydroxyethyl terephthalate (MHET), and consequently PET precursors TPA and ethylene glycol. The decrease in the intensity of the absorption band of the ester bond between 203–207 nm and the appearance of the band at 201 nm, characteristic of free carboxyl groups, confirm the results [64]. The ethylene glycol molecule could be easily mineralized to CO2 and H2O by biological means. In addition, several studies that have addressed the microbial bioremediation of PET by various microbial consortia refer to its transformation into mono (2-hydroxyethyl) terephthalic acid (MHET) by the extracellular enzyme PETase. MHET passes to the periplasmic region of the microorganisms, where it is transformed by the enzyme MHETase into terephthalic acid and ethylene glycol [65]. Such a mechanism agrees with the analysis of the UV–Vis spectra of heterogeneous photocatalysis by the (Al2O3)0.75TiO2 system with the advantage that no strict control favoring enzymatic activity is necessary for PET biodegradation.
Figure 12 shows that the degradation efficiency of PET was 37.03%, measured by absorbance at 210 nm, and 41.18%, measured as COD; in both cases, a continuous degradation trend is observed, which could continue for more than 8 h until obtaining degradation equal to or greater than 50% of PET.
According to the test, pseudo-first-order kinetics was observed with a photocatalysis rate constant of kapp = 0.0534 h−1, considering the COD measurements of PET degradation. Following pseudo-first-order kinetics, 50% degradation of PET would occur in a period of approximately 13 h, which is equivalent to the half-life of the reaction (Figure 12b).

3.4. ANOVA Analysis

Analysis of variance (ANOVA) is a statistical method that was used to analyze the differences between absorbance and COD measurements performed in triplicate for each contaminant, assessing whether the group means were significantly different between them (Table 1) or not. ANOVA calculates the F value (Fischer test value) and the sum of squares to evaluate the significance of the parameters. A p-value less than 0.05, corresponding to a significance level of 5% or a confidence interval of 95%, indicates statistical significance.
In all the experiments, the critical F value was 3.48. Since the F values are much higher and the p values are less than 0.05, all the measurements made for each contaminant by both COD and absorbance are statistically significant, i.e., the three data sets from each experiment are different but can be considered as equivalent. These results suggest that each test, performed in triplicate, is solely dependent on the photocatalyst concentration with no other variables influencing the degradation of TPA and PET.
The highest F values were observed for the experiments that quantified the concentration by COD, indicating that this analysis is more meaningful and reproducible than the analysis by UV–Vis spectroscopy.

4. Conclusions

The selection of TPA and PET contaminants was made because TPA is the hydrolysis product of PET, PET can be degraded to TPA, and the latter can also be degraded to less toxic contaminants, which represents a challenge for water treatment technology. The reduction in the COD measurement for an initial TPA concentration of 50 mg/L in the reactor with the modified arrangement of stainless-steel disks with the (Al2O3)0.75TiO2 coating and lamps reached 46.60%. This kind of reduction could be associated with the low-scale formation of CO2 from the TPA carboxyl groups. The UV–Vis analysis for 8 h of determination showed the splitting of initial bands, which resulted in the formation of intermediate products. Commercial PET sampling in the presence of the immobilized (Al2O3)0.75TiO2 photocatalyst reduced the COD value to 41% and the UV–Vis analysis revealed degradation efficiency of 37%.

Author Contributions

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

Funding

This research received no external funding.

Data Availability Statement

Data is contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Peng, Y.; Yang, J.; Deng, C.; Deng, J.; Shen, L.; Fu, Y. Acetolysis of waste polyethylene terephthalate for upcycling and life-cycle assessment study. Nat. Commun. 2023, 14, 3249. [Google Scholar] [CrossRef] [PubMed]
  2. Fang, P.; Liu, B.; Xu, J.; Zhou, Q.; Zhang, S.; Ma, J.; Lu, X. High-efficiency glycolysis of poly(ethylene terephthalate) by sandwich-structure polyoxometalate catalyst with two active sites. Polym. Degrad. Stab. 2018, 156, 22–31. [Google Scholar] [CrossRef]
  3. Xiao, Y.; Luo, W.P.; Zhang, X.Y.; Guo, C.C.; Liu, Q.; Jiang, G.F.; Li, Q.H. Aerobic Oxidation of p-Toluic Acid to Terephthalic Acid over T(p-Cl)PPMnCl/Co(OAc)2 Under Moderate Conditions. Catal. Lett. 2010, 134, 155–161. [Google Scholar] [CrossRef]
  4. Zhang, H.; Wang, Y.; Li, X.; Deng, K.; Yu, H.; Xu, Y.; Wang, H.; Wang, Z.; Wang, L. Electrocatalytic upcycling of polyethylene terephthalate plastic to formic acid coupled with energy-saving hydrogen production over hierarchical Pd-doped NiTe nanoarrays. Appl. Catal. B Environ. 2024, 340, 123236. [Google Scholar] [CrossRef]
  5. Ma, K.; Li, X.; Bao, L.; Li, X.; Cui, Y. The performance and bacterial community shifts in an anaerobic-aerobic process treating purified terephthalic acid wastewater under influent composition variations and ambient temperatures. J. Clean. Prod. 2020, 276, 124190. [Google Scholar] [CrossRef]
  6. Billings, A.; Jones, K.C.; Pereira, M.G.; Spurgeon, D.J. Plasticisers in the terrestrial environment: Sources, occurrence and fate. Environ. Chem. 2021, 18, 111–130. [Google Scholar] [CrossRef]
  7. Chang, B.V.; Yang, C.M.; Cheng, C.H.; Yuan, S.Y. Biodegradation of phthalate esters by two bacteria strains. Chemosphere 2004, 55, 533–538. [Google Scholar] [CrossRef]
  8. Kleerebezem, R.; Beckers, J.; Hulshoff Pol, L.W.; Lettinga, G. High rate treatment of terephthalic acid production wastewater in a two-stage anaerobic bioreactor. Biotechnol. Bioeng. 2005, 91, 169–179. [Google Scholar] [CrossRef]
  9. Sannino, D.; Vaiano, V.; Sacco, O.; Morante, N.; De Guglielmo, L.; Di Capua, G.; Femia, N. Visible Light Driven Degradation of Terephthalic Acid: Optimization of Energy Demand by Light Modulation Techniques. J. Photocatal. 2021, 2, 49–61. [Google Scholar] [CrossRef]
  10. Li, H.; Zhang, W.; Xia, D.; Ye, L.; Ma, W.; Li, H.; Li, Q.; Wang, Y. Improved anaerobic degradation of purified terephthalic acid wastewater by adding nanoparticles or co-substrates to facilitate the electron transfer process. Environ. Sci. Nano 2022, 9, 1011–1024. [Google Scholar] [CrossRef]
  11. Muñoz Sierra, J.D.; Lafita, C.; Gabaldón, C.; Spanjers, H.; van Lier, J.B. Trace metals supplementation in anaerobic membrane bioreactors treating highly saline phenolic wastewater. Bioresour. Technol. 2017, 234, 106–114. [Google Scholar] [CrossRef] [PubMed]
  12. Zhang, H.; Wu, J.; Li, R.; Kim, D.-H.; Bi, X.; Zhang, G.; Jiang, B.; Yong Ng, H.; Shi, X. Novel intertidal wetland sediment-inoculated moving bed biofilm reactor treating high-salinity wastewater: Metagenomic sequencing revealing key functional microorganisms. Bioresour. Technol. 2022, 348, 126817. [Google Scholar] [CrossRef]
  13. Ma, X.; Li, X.; Wang, X.; Liu, G.; Zuo, J.; Wang, S.; Wang, K. Impact of salinity on anaerobic microbial community structure in high organic loading purified terephthalic acid wastewater treatment system. J. Hazard. Mater. 2020, 383, 121132. [Google Scholar] [CrossRef]
  14. Zhang, H.; Tian, L.; Zhang, Z.; Han, J.; Wu, Z.; Wei, Z.; Wang, S.; Cao, Y.; Zhang, S.; Zhang, Y. Effective degradation of phenol by in-situ photocatalytic-Fenton-like technology with BiVO4/Bi2WO6/Ti3C2 QDs. Surf. Interfaces 2024, 55, 105315. [Google Scholar] [CrossRef]
  15. Fuentes, I.; Julia, L.R.; Hugo, T.; José, M.R.-H.; Isaac, C.; Poznyak, T. Terephthalic acid decomposition by photocatalytic ozonation with VxOy/ZnO under different UV-A LEDs distributions. Chem. Eng. Commun. 2020, 207, 263–277. [Google Scholar] [CrossRef]
  16. Zhu, Y.; Luo, H.; Wang, T.; Zhou, J.; Zhu, Y.; Hou, J. Advanced treatment of landfill leachate by catalytic ozonation with MnCeOx/γ-Al2O3 catalyst. Surf. Interfaces 2024, 46, 104113. [Google Scholar] [CrossRef]
  17. Jin, Y.; Gao, Y.; Pang, D.; Li, J.; Wang, X.; Chen, G.; Wang, W.; Mao, Y. Fe3C-modified carbon composites for boosted peroxymonosulfate activation and organic pollutant decontamination: Contributions of dual radical and nonradical pathways. Surf. Interfaces 2024, 46, 104085. [Google Scholar] [CrossRef]
  18. Žerjav, G.; Albreht, A.; Vovk, I.; Pintar, A. Revisiting terephthalic acid and coumarin as probes for photoluminescent determination of hydroxyl radical formation rate in heterogeneous photocatalysis. Appl. Catal. A Gen. 2020, 598, 117566. [Google Scholar] [CrossRef]
  19. Sacco, O.; Vaiano, V.; Rizzo, L.; Sannino, D. Photocatalytic activity of a visible light active structured photocatalyst developed for municipal wastewater treatment. J. Clean. Prod. 2018, 175, 38–49. [Google Scholar] [CrossRef]
  20. Jiang, Q.; Liu, J.; Qi, T.; Liu, Y. Enhanced visible-light photocatalytic activity and antibacterial behaviour on fluorine and graphene synergistically modified TiO2 nanocomposite for wastewater treatment. Environ. Technol. 2022, 43, 3821–3834. [Google Scholar] [CrossRef]
  21. Merino, N.; Qu, Y.; Deeb, R.A.; Hawley, E.L.; Hoffmann, M.R.; Mahendra, S. Degradation and Removal Methods for Perfluoroalkyl and Polyfluoroalkyl Substances in Water. Environ. Eng. Sci. 2016, 33, 615–649. [Google Scholar] [CrossRef]
  22. Byrne, C.; Subramanian, G.; Pillai, S.C. Recent advances in photocatalysis for environmental applications. J. Environ. Chem. Eng. 2018, 6, 3531–3555. [Google Scholar] [CrossRef]
  23. Al Miad, A.; Saikat, S.P.; Alam, M.K.; Sahadat Hossain, M.; Bahadur, N.M.; Ahmed, S. Metal oxide-based photocatalysts for the efficient degradation of organic pollutants for a sustainable environment: A review. Nanoscale Adv. 2024, 6, 4781–4803. [Google Scholar] [CrossRef]
  24. Krishnan, A.; Swarnalal, A.; Das, D.; Krishnan, M.; Saji, V.S.; Shibli, S.M.A. A review on transition metal oxides based photocatalysts for degradation of synthetic organic pollutants. J. Environ. Sci. 2024, 139, 389–417. [Google Scholar] [CrossRef]
  25. Wang, Y.; Xu, M.; Li, J.; Zhang, T. Photocatalytic Degradation of Organic Pollutants Using Al/TiO2 Composites Under Visible Light. Environ. Eng. Sci. 2024, 41, 204–215. [Google Scholar] [CrossRef]
  26. Ran, Y.; Zhong, J.; Li, J.; Li, M.; Tian, C. Substantially boosted photocatalytic detoxification activity of TiO2 benefited from Eu doping. Environ. Technol. 2023, 44, 1313–1321. [Google Scholar] [CrossRef]
  27. Zhang, Q.; Li, R.; Li, Z.; Li, A.; Wang, S.; Liang, Z.; Liao, S.; Li, C. The dependence of photocatalytic activity on the selective and nonselective deposition of noble metal cocatalysts on the facets of rutile TiO2. J. Catal. 2016, 337, 36–44. [Google Scholar] [CrossRef]
  28. Yakdoumi, F.Z.; Hadj-Hamou, A.S. Effectiveness assessment of TiO2-Al2O3 nano-mixture as a filler material for improvement of packaging performance of PLA nanocomposite films. J. Polym. Eng. 2020, 40, 848–858. [Google Scholar] [CrossRef]
  29. Karunakaran, C.; Magesan, P.; Gomathisankar, P.; Vinayagamoorthy, P. Absorption, emission, charge transfer resistance and photocatalytic activity of Al2O3/TiO2 core/shell nanoparticles. Superlattices Microstruct. 2015, 83, 659–667. [Google Scholar] [CrossRef]
  30. Magnone, E.; Kim, M.K.; Lee, H.J.; Park, J.H. Facile synthesis of TiO2-supported Al2O3 ceramic hollow fiber substrates with extremely high photocatalytic activity and reusability. Ceram. Int. 2021, 47, 7764–7775. [Google Scholar] [CrossRef]
  31. Martinez-Gómez, C.; Rangel-Vazquez, I.; Zarraga, R.; del Ángel, G.; Ruíz-Camacho, B.; Tzompantzi, F.; Vidal-Robles, E.; Perez-Larios, A. Photodegradation and Mineralization of Phenol Using TiO2Coated γ-Al2O3: Effect of Thermic Treatment. Processes 2022, 10, 1186. [Google Scholar] [CrossRef]
  32. Zhu, Y.; Ning, Y.; Li, L.; Chen, Z.; Li, H.; Zhang, Y. Effective removal of hexavalent chromium from aqueous system by biochar-supported titanium dioxide (TiO2). Environ. Chem. 2022, 19, 432–445. [Google Scholar] [CrossRef]
  33. Kanakaraju, D.; Motti, C.A.; Glass, B.D.; Oelgemöller, M. Photolysis and TiO2-catalysed degradation of diclofenac in surface and drinking water using circulating batch photoreactors. Environ. Chem. 2014, 11, 51–62. [Google Scholar] [CrossRef]
  34. Zhuang, Q.; Shi, K.; Wang, J.; Zhou, H.; Zhao, P.; Lou, Y. Revolutionizing pollution control with innovative CuO@TiO2 nanocomposite for enhanced photocatalytic degradation and antimicrobial efficacy. Surf. Interfaces 2024, 55, 105410. [Google Scholar] [CrossRef]
  35. Yu, X.; Wang, S.; Zhang, Y.; Yu, X.; Zhao, X.; Gao, H.; Yang, H.; Fang, L.; Syed, A. Biochar coupling induced charge transfer switching of a NiAl2O4/NiCr2O4 photocatalyst to enhance photocatalytic degradation of tetracycline hydrochloride. Surf. Interfaces 2024, 54, 105222. [Google Scholar] [CrossRef]
  36. Duan, L.; Wang, B.; Heck, K.; Guo, S.; Clark, C.A.; Arredondo, J.; Wang, M.; Senftle, T.P.; Westerhoff, P.; Wen, X.; et al. Efficient Photocatalytic PFOA Degradation over Boron Nitride. Environ. Sci. Technol. Lett. 2020, 7, 613–619. [Google Scholar] [CrossRef]
  37. Samy, M.; Ibrahim, M.G.; Fujii, M.; Diab, K.E.; ElKady, M.; Gar Alalm, M. CNTs/MOF-808 painted plates for extended treatment of pharmaceutical and agrochemical wastewaters in a novel photocatalytic reactor. Chem. Eng. J. 2021, 406, 127152. [Google Scholar] [CrossRef]
  38. Ibarra, I.C.R.; Dubé, I.Z.; Sánchez, M.F.G. 1. MX2019002787—Reactor Fotocatalítico para la Degradación de Contaminantes Orgánicos a Presiones por Encima de los Valores Ambientales. 2019002787. 2020. Available online: https://www.wipo.int/ipcpub/?symbol=B01J0003000000&menulang=es&lang=es (accessed on 1 August 2024).
  39. Camacho-González, M.A.; Lijanova, I.V.; Reyes-Miranda, J.; Sarmiento-Bustos, E.; Quezada-Cruz, M.; Vera-Serna, P.; Barrón-Meza, M.Á.; Garrido-Hernández, A. High Photocatalytic Efficiency of Al2O3-TiO2 Coatings on 304 Stainless Steel for Methylene Blue and Wastewater Degradation. Catalysts 2023, 13, 1351. [Google Scholar] [CrossRef]
  40. Habibi, S.; Jamshidi, M. Sol–gel synthesis of carbon-doped TiO2 nanoparticles based on microcrystalline cellulose for efficient photocatalytic degradation of methylene blue under visible light. Environ. Technol. 2020, 41, 3233–3247. [Google Scholar] [CrossRef]
  41. Farhadian Azizi, K.; Bagheri-Mohagheghi, M.-M. Transition from anatase to rutile phase in titanium dioxide (TiO2) nanoparticles synthesized by complexing sol–gel process: Effect of kind of complexing agent and calcinating temperature. J. Sol-Gel Sci. Technol. 2013, 65, 329–335. [Google Scholar] [CrossRef]
  42. Zhu, L.; Lu, Q.; Lv, L.; Wang, Y.; Hu, Y.; Deng, Z.; Lou, Z.; Hou, Y.; Teng, F. Ligand-free rutile and anatase TiO2 nanocrystals as electron extraction layers for high performance inverted polymer solar cells. RSC Adv. 2017, 7, 20084–20092. [Google Scholar] [CrossRef]
  43. Sadeq, Z.S.; Mahdi, Z.F.; Hamza, A.M. Low cost, fast and powerful performance interfacial charge transfer nanostructured Al2O3 capturing of light photocatalyst eco-friendly system using hydrothermal method. Mater. Lett. 2019, 254, 120–124. [Google Scholar] [CrossRef]
  44. Xu, L.; Song, H.; Chou, L. Facile synthesis of nano-crystalline alpha-alumina at low temperature via an absolute ethanol sol–gel strategy. Mater. Chem. Phys. 2012, 132, 1071–1076. [Google Scholar] [CrossRef]
  45. Askeland, D. Ciencia e Ingeniería de Materiales, 3rd ed.; Thomson, Ed.; Cengage Learning, Inc.: Toronto, ON, Canada, 2012. [Google Scholar]
  46. Miglierini, M.B.; Pašteka, L.; Cesnek, M.; Kmječ, T.; Bujdoš, M.; Kohout, J. Influence of surface treatment on microstructure of stainless steels studied by Mössbauer spectrometry. J. Radioanal. Nucl. Chem. 2019, 322, 1495–1503. [Google Scholar] [CrossRef]
  47. Zhang, X.; Yang, S.; Li, J.; Wu, J. Temperature-dependent evolution of oxide inclusions during heat treatment of stainless steel with yttrium addition. Int. J. Miner. Metall. Mater. 2020, 27, 754–763. [Google Scholar] [CrossRef]
  48. da Trindade, C.d.M.; da Silva, S.W.; Bortolozzi, J.P.; Banús, E.D.; Bernardes, A.M.; Ulla, M.A. Synthesis and characterization of TiO2 films onto AISI 304 metallic meshes and their application in the decomposition of the endocrine-disrupting alkylphenolic chemicals. Appl. Surf. Sci. 2018, 457, 644–654. [Google Scholar] [CrossRef]
  49. Oladijo, O.P.; Popoola, A.P.I.; Booi, M.; Fayomi, J.; Collieus, L.L. Corrosion and mechanical behaviour of Al2O3.TiO2 composites produced by spark plasma sintering. South Afr. J. Chem. Eng. 2020, 33, 58–66. [Google Scholar] [CrossRef]
  50. Liu, X.; Wan, Y.; Zhang, X. Preparation and Corrosion Properties of TiO2-SiO2-Al2O3 Composite Coating on Q235 Carbon Steel. Coatings 2023, 13, 1994. [Google Scholar] [CrossRef]
  51. Brinker, C.J.; Scherer, G.W. Sol-Gel Science: The Physics and Chemistry of Sol-Gel Processing; Academic Press: Cambridge, MA, USA, 2013; ISBN 0080571034. [Google Scholar]
  52. Tiss, B.; Zayoud, W.; Sekrafi, H.E.; Bouguila, N.; Cristea, D.; Croitoru, C.; Velicu, L.; Tiron, V.; Prepelita, P.; Craciun, V.; et al. Enhancing photocatalysis through annealing: Unveiling the role of physical properties in photocatalytic behavior. Mater. Chem. Phys. 2024, 313, 128665. [Google Scholar] [CrossRef]
  53. Ahmad, I.; Zou, Y.; Yan, J.; Liu, Y.; Shukrullah, S.; Naz, M.Y.; Hussain, H.; Khan, W.Q.; Khalid, N.R. Semiconductor photocatalysts: A critical review highlighting the various strategies to boost the photocatalytic performances for diverse applications. Adv. Colloid Interface Sci. 2023, 311, 102830. [Google Scholar] [CrossRef]
  54. Akkaya Arıer, U.O.; Tepehan, F.Z. Influence of Al2O3:TiO2 ratio on the structural and optical properties of TiO2–Al2O3 nano-composite films produced by sol gel method. Compos. Part B Eng. 2014, 58, 147–151. [Google Scholar] [CrossRef]
  55. Kolesnik, O. UV-Vis Spectrum of Terephthalic Acid. SIELC Technol. Available online: https://sielc.com/uv-vis-spectrum-of-terephthalic-acid (accessed on 1 August 2024).
  56. Skoog, D.A.; Holler, F.J.; Crouch, S.R. Principles of Instrumental Analysis. Cengage Learning. 2017. Available online: https://books.google.com.mx/books?id=D13EDQAAQBAJ (accessed on 1 August 2024).
  57. Zielińska-Jurek, A.; Wysocka, I.; Janczarek, M.; Stampor, W.; Hupka, J. Preparation and characterization of Pt–N/TiO2 photocatalysts and their efficiency in degradation of recalcitrant chemicals. Sep. Purif. Technol. 2015, 156, 369–378. [Google Scholar] [CrossRef]
  58. Agostini, I.; Ciuffi, B.; Gallorini, R.; Rizzo, A.M.; Chiaramonti, D.; Rosi, L. Recovery of Terephthalic Acid from Densified Post-consumer Plastic Mix by HTL Process. Molecules 2022, 27, 7112. [Google Scholar] [CrossRef]
  59. Shafaei, A.; Nikazar, M.; Arami, M. Photocatalytic degradation of terephthalic acid using titania and zinc oxide photocatalysts: Comparative study. Desalination 2010, 252, 8–16. [Google Scholar] [CrossRef]
  60. Mahmoodi, N.M.; Arami, M. Bulk phase degradation of Acid Red 14 by nanophotocatalysis using immobilized titanium(IV) oxide nanoparticles. J. Photochem. Photobiol. A Chem. 2006, 182, 60–66. [Google Scholar] [CrossRef]
  61. Miwa, T.; Kaneco, S.; Katsumata, H.; Suzuki, T.; Ohta, K.; Chand Verma, S.; Sugihara, K. Photocatalytic hydrogen production from aqueous methanol solution with CuO/Al2O3/TiO2 nanocomposite. Int. J. Hydrog. Energy 2010, 35, 6554–6560. [Google Scholar] [CrossRef]
  62. Ali, I.; Suhail, M.; Alothman, Z.A.; Alwarthan, A. Recent advances in syntheses, properties and applications of TiO2 nanostructures. RSC Adv. 2018, 8, 30125–30147. [Google Scholar] [CrossRef]
  63. Sotiles, A.R.; Barbosa, L.V.; Vedovato, Y.P.; Wypych, F.; de Faria, E.H. Effect of Zn2Cr layered double hydroxide on photocatalytic hydroxylation of terephthalic acid and photodegradation of sodium diclofenac. Catal. Today 2023, 423, 114287. [Google Scholar] [CrossRef]
  64. Tawfik, M.; Eskander, S. Recycling of Polyethylene Terephthalate Plastic Wastes based on unsaturated Diol. KGK Rubberpoint 2015, 68, 21–27. [Google Scholar]
  65. Kulkarni, R.M.; Patwardhan, N.S.; Iyer, P.B.; Bharadwaj, T.D. A review on microbial bioremediation of polyethylene terephthalate microplastics. Environ. Qual. Manag. 2024, 34, e22264. [Google Scholar] [CrossRef]
Figure 1. Airlift reactor: (a) arrangement of 304 SS disks with (Al2O3)0.75TiO2 coatings; (b) top view of the arrangement of lamps and 304 SS disks; (c) placement of lamps. The numbers correspond to lengths in millimeters. (d) Spectrum of the Philips MR16 LED lamps with color temperature of 4000 K used for the white-light optical transmission measurements.
Figure 1. Airlift reactor: (a) arrangement of 304 SS disks with (Al2O3)0.75TiO2 coatings; (b) top view of the arrangement of lamps and 304 SS disks; (c) placement of lamps. The numbers correspond to lengths in millimeters. (d) Spectrum of the Philips MR16 LED lamps with color temperature of 4000 K used for the white-light optical transmission measurements.
Surfaces 08 00034 g001
Figure 2. XRD patterns of the (Al2O3)0.75TiO2 coating powders.
Figure 2. XRD patterns of the (Al2O3)0.75TiO2 coating powders.
Surfaces 08 00034 g002
Figure 3. XRD patterns of 304 SS disks with and without thermal treatment.
Figure 3. XRD patterns of 304 SS disks with and without thermal treatment.
Surfaces 08 00034 g003
Figure 4. Micrographs and mapping of the elemental analysis of the coatings of (a) TiO2 and (b) (Al2O3)0.75TiO2 deposited on 304 stainless steel.
Figure 4. Micrographs and mapping of the elemental analysis of the coatings of (a) TiO2 and (b) (Al2O3)0.75TiO2 deposited on 304 stainless steel.
Surfaces 08 00034 g004
Figure 5. Micrographs of (a) TiO2 coating on 304 stainless steel disks, (a′) magnification of the white zone of the TiO2 coating and particle size distribution plots; (b) (Al2O3)0.75TiO2 coating, (b′) magnification of the white zone and particle size distribution plots.
Figure 5. Micrographs of (a) TiO2 coating on 304 stainless steel disks, (a′) magnification of the white zone of the TiO2 coating and particle size distribution plots; (b) (Al2O3)0.75TiO2 coating, (b′) magnification of the white zone and particle size distribution plots.
Surfaces 08 00034 g005
Figure 6. UV–Vis absorption spectrograms of the (Al2O3)0.75TiO2 powders.
Figure 6. UV–Vis absorption spectrograms of the (Al2O3)0.75TiO2 powders.
Surfaces 08 00034 g006
Figure 7. Monitoring of the UV–Vis absorption during the photocatalytic degradation of TPA.
Figure 7. Monitoring of the UV–Vis absorption during the photocatalytic degradation of TPA.
Surfaces 08 00034 g007
Figure 8. Photocatalytic degradation of TPA with (a) TiO2 and (b) (Al2O3)0.75TiO2 coatings.
Figure 8. Photocatalytic degradation of TPA with (a) TiO2 and (b) (Al2O3)0.75TiO2 coatings.
Surfaces 08 00034 g008
Figure 9. Photocatalytic degradation kinetics of TPA using (a) TiO2, and (b) (Al2O3)0.75TiO2 coatings.
Figure 9. Photocatalytic degradation kinetics of TPA using (a) TiO2, and (b) (Al2O3)0.75TiO2 coatings.
Surfaces 08 00034 g009
Figure 10. Band diagram for the (Al2O3)0.75TiO2 catalyst: formation of a step-gap type heterojunction between TiO2 and Al2O3.
Figure 10. Band diagram for the (Al2O3)0.75TiO2 catalyst: formation of a step-gap type heterojunction between TiO2 and Al2O3.
Surfaces 08 00034 g010
Figure 11. UV–Vis analysis of PET structural changes by heterogeneous photocatalysis by (Al2O3)0.75TiO2.
Figure 11. UV–Vis analysis of PET structural changes by heterogeneous photocatalysis by (Al2O3)0.75TiO2.
Surfaces 08 00034 g011
Figure 12. (a) Photocatalytic degradation of PET: (Al2O3)0.75TiO2 coating; (b) degradation kinetics.
Figure 12. (a) Photocatalytic degradation of PET: (Al2O3)0.75TiO2 coating; (b) degradation kinetics.
Surfaces 08 00034 g012
Table 1. ANOVA analysis of triplicate experiments.
Table 1. ANOVA analysis of triplicate experiments.
ExperimentSum of SquaresDegree of FreedomMean SquareFvaluepvalueSignificance
TPA_TiO2_DQO12,102.8343025.713272.21<0.05Optimal
TPA_TiO2_Abs2470.1240.03116.60<0.05Optimal
TPA_(Al2O3)0.75TiO2_DQO25,757.9646439.494164.36<0.05Optimal
TPA_(Al2O3)0.75TiO2_Abs2470.2140.051847.31<0.05Optimal
PET_(Al2O3)0.75TiO2_DQO308,610.69477,152.6722,932.53<0.05Optimal
PET_(Al2O3)0.75TiO2_Abs2100.3540.09126.28<0.05Optimal
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Camacho-González, M.A.; Hernández-Reyes, A.; Garrido-Hernández, A.; Olivares-Xometl, O.; Likhanova, N.V.; Lijanova, I.V. TPA and PET Photo-Degradation by Heterogeneous Catalysis Using a (Al2O3)0.75TiO2 Coating. Surfaces 2025, 8, 34. https://doi.org/10.3390/surfaces8020034

AMA Style

Camacho-González MA, Hernández-Reyes A, Garrido-Hernández A, Olivares-Xometl O, Likhanova NV, Lijanova IV. TPA and PET Photo-Degradation by Heterogeneous Catalysis Using a (Al2O3)0.75TiO2 Coating. Surfaces. 2025; 8(2):34. https://doi.org/10.3390/surfaces8020034

Chicago/Turabian Style

Camacho-González, Mónica A., Alberto Hernández-Reyes, Aristeo Garrido-Hernández, Octavio Olivares-Xometl, Natalya V. Likhanova, and Irina V. Lijanova. 2025. "TPA and PET Photo-Degradation by Heterogeneous Catalysis Using a (Al2O3)0.75TiO2 Coating" Surfaces 8, no. 2: 34. https://doi.org/10.3390/surfaces8020034

APA Style

Camacho-González, M. A., Hernández-Reyes, A., Garrido-Hernández, A., Olivares-Xometl, O., Likhanova, N. V., & Lijanova, I. V. (2025). TPA and PET Photo-Degradation by Heterogeneous Catalysis Using a (Al2O3)0.75TiO2 Coating. Surfaces, 8(2), 34. https://doi.org/10.3390/surfaces8020034

Article Metrics

Back to TopTop