Next Article in Journal
The Catalytic Wet Oxidation of Excess Activated Sludge from a Coal Chemical Wastewater Treatment Process
Previous Article in Journal
The Acquisition of Primary Amines from Alcohols through Reductive Amination over Heterogeneous Catalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

High Photocatalytic Efficiency of Al2O3-TiO2 Coatings on 304 Stainless Steel for Methylene Blue and Wastewater Degradation

by
Mónica Araceli Camacho-González
1,2,
Irina Victorovna Lijanova
1,
Joan Reyes-Miranda
3,
Estela Sarmiento-Bustos
4,
Maribel Quezada-Cruz
2,
Pedro Vera-Serna
5,
Miguel Ángel Barrón-Meza
3 and
Aristeo Garrido-Hernández
2,*
1
Centro de Investigación e Innovación Tecnológica, Instituto Politécnico Nacional, Ciudad de México 02250, Mexico
2
División Químico Biológicas, Universidad Tecnológica de Tecámac, Tecámac 55740, Mexico
3
Departamento de Ciencia de los Materiales, Universidad Autónoma Metropolitana-Azcapotzalco, Ciudad de México 02200, Mexico
4
División de Mecánica Industrial, Universidad Tecnológica Emiliano Zapata del Estado de Morelos, Emiliano Zapata 62765, Mexico
5
División de Ingeniería, Universidad Politécnica de Tecámac, Tecámac 55740, Mexico
*
Author to whom correspondence should be addressed.
Catalysts 2023, 13(10), 1351; https://doi.org/10.3390/catal13101351
Submission received: 11 September 2023 / Revised: 25 September 2023 / Accepted: 28 September 2023 / Published: 7 October 2023
(This article belongs to the Section Catalytic Materials)

Abstract

:
This work explores the novelty of achieving high photocatalytic efficiency and remarkable bactericidal activity with Al2O3-TiO2 coatings on perforated 304 stainless steel (SS) substrates, placed transversely along an airlift reactor of 0.980 L for wastewater treatment under visible light irradiation. The Al2O3-TiO2 coatings achieved methylene blue and total organic carbon (TOC) concentration reductions of 97.3 and 96.51%, respectively, in a wastewater sample with heterogeneous photocatalsis. The Al2O3-TiO2 coatings resulted in a 33.30% reduction in total and fecal coliforms and a remarkable 94.23% decrease in Salmonella spp. in the wastewater sample. XRD confirmed the TiO2 anatase–rutile phases and Al2O3 α-γ phases in the coating. The particle size distribution ranges from 100 to 500 nm, and the coating surface was homogeneous without cracks confirmed using SEM and AFM, respectively. The roughness and thickness of the coatings were 85 ± 5 nm and 250 ± 50 nm, respectively.

1. Introduction

Every day, approximately 2 million tons of untreated wastewater and other effluents are discharged into rivers, lagoons, lakes, and seas, according to the United Nations [1,2]. Conventional wastewater treatment methods face challenges in eliminating emerging contaminants like dyes, pigments, antibiotics, and heavy metals [3,4,5]. Traditional techniques such as adsorbent resins, adsorption on activated carbon, ultrafiltration, reverse osmosis, ion exchange on synthetic materials, etc., remove dye pollutants from water, but these techniques only transfer organic compounds from water to another phase, promoting secondary pollution [6]. To address this issue, recent scientific and technological advancements have been focused on renewing and evolving water treatment methods [7,8]. Advanced oxidation processes (AOPs) have assisted or replaced some operations in wastewater treatment plants (WWTPs) [9]. AOPs can effectively and economically remove organic compounds and pathogenic microorganisms from wastewater [10]. One such AOP is heterogeneous photocatalysis, which employs a semiconductor material exhibiting photocatalytic properties in the ultraviolet or visible range [11,12]. When the energy is equal to or greater than its bandgap, highly oxidizing hydroxyl (OH) and superoxide (O2●−) radicals are generated. These species contribute to the total mineralization of the polluting organic compounds in water through successive redox reactions and to cellular lysis of pathogenic microorganisms due to oxidative stress [13,14].
There are significant differences between the use of photocatalysts in powder form versus immobilized form [15]. In the case of powders, dispersion of the photocatalyst in the solution requires stirring, which leads to increased processing time and cost due to the need for catalyst separation with filtration or centrifugation [16,17]. However, the use of photocatalyst coatings can eliminate this drawback by requiring high adherence between the substrate and coating [18,19]. Many methods for preparing ceramic coating have been employed, such as chemical vapor deposition, physical vapor deposition [20], sol–gel deposition [21], electroplating [22], spin-coating [23], dip-coating [24], chemical bath deposition [25], atomic layer deposition [26], and spray coating [27], and offer a diverse range of approaches. The dip-coating technique is exceptional because it provides outstanding coating uniformity and precise control over coating thickness. Furthermore, dip-coating is often cost-effective due to its simplicity and minimal equipment requirements. Another noteworthy aspect of dip-coating is its compatibility with various substrates, including plastics, metals, ceramics, and composites, further expanding its applicability across diverse materials and industries. One of the most important substrates of dip-coating is AISI 304 stainless steel (SS), a versatile and widely used material known for its affordability, excellent mechanical properties, ease of fabrication, and high resistance to corrosion and oxidation [28,29].
TiO2 is one of the most widely used photocatalysts [30,31], known for its mesoporous network, high surface area, and crystallinity, which favor the accessibility and diffusion of target molecules in photocatalytic applications using either UV or visible light [32]. However, a major drawback of TiO2 photocatalysts is that photo-generated charge carriers tend to recombine, thereby decreasing reaction efficiency [33]. Various methods, including heterojunction formation, ion doping, and nanosized crystals, have proven effective in improving the recombination process of photo-generated electrons in TiO2 [34].
The Al2O3-TiO2 material might reduce the recombination of photo-generated charge carriers and overcome the main shortcoming of TiO2 photocatalysts. Numerous research studies have reported better photocatalytic activity of Al2O3-TiO2 mixed oxides in suspension or as thin films under both UV and visible radiation when compared to pure TiO2 [35,36,37]. Magnone and coworkers [38] reported that TiO2-supported Al2O3 ceramic hollow fiber substrates exhibited a remarkable 91% removal efficiency of methylene blue MB (20 ppm) in one hour. These substrates displayed exceptional photocatalytic stability, retaining their high performance even after multiple cycles. For instance, 10 μm-thickness Al2O3-TiO2 membranes synthesized using the sol–gel spin-coating method demonstrated 99.25% degradation of methylene blue (MB) under 30 W UV light, and notably after five cycles, they still exhibited high MB degradation (above 96%) [39]. Moreover, the catalytic performance of Al2O3-TiO2 extends beyond MB degradation. In comparison to TiO2, Al2O3-TiO2 (75–90% of TiO2) powders displayed twice the photocatalytic degradation rate when applied to a salicylic acid (SA) solution (0.144 mM) in 135 mL and 1 g/L of catalyst [40]. Additionally, Martinez-Gomez [41] prepared a γ-Al2O3-TiO2 mixed oxide using hydrolysis, employing the boehmite method. The catalyst treated at 600 °C showed the highest activity since it degraded about 95% of phenol in 4 h in a glass reactor under UV light, using 200 mL of a phenol solution (400 ppm) and 100 mg of catalyst. These findings underscore the versatility and effectiveness of Al2O3-TiO2 coatings.
Prior research focused on Al2O3-TiO2 photocatalytic degradation primarily in low volumes with high-power UV lamps and lacked reactor design. Limited studies examined coatings using residential wastewater samples. To our knowledge, the photocatalytic and antibacterial activity of Al2O3-TiO2 (with Al/Ti atomic ratio = 1.5) coating on 304 stainless steel using airlift rector of 0.918 L volume employing low-power (4 Watt) visible light have not been explored.
The present study evaluated the photocatalytic activity of Al2O3-TiO2 coatings on perforated 304 SS substrates, which were placed transversely in a 0.980 L vertical glass reactor, employing visible light. The reduction in the concentration of MB as a model molecule was determined using UV-vis spectrophotometry. Furthermore, the concentration of total organic pollutants in a sample of residential wastewater was quantified as total organic carbon (TOC). Additionally, bactericidal activity was evaluated by quantifying the total and fecal coliforms.

2. Results and Discussion

2.1. Al2O3-TiO2 Coating Characterization

2.1.1. X-ray Diffraction

Figure 1 displays prominent peaks that are ascribed to the (111), (200), and (220) planes of Fe-γ (JCPDS Card No. 33-0397, austenite) related to 304 SS [42]. The diffraction planes of the Al2O3-TiO2 coating are shown at the bottom, corresponding to the magnified, red-framed section. The crystallization of the alumina α- and γ-phases (JCPDS Card Nos. 11-0425 and 10-0661) was induced by the synthesis route and thermal treatment, which promoted the transformation of AlOOH into α-Al2O3; this transformation normally takes place at low temperature (450 °C) [43], while the transformation into γ-Al2O3 occurs between 400–700 °C [44,45]. The presence of α-Al2O3 was evidenced by the diffraction peaks located at around 2θ = 25.3°, 33.2°, 35.6°, 37.9°, 43.3°, and 57.68°, which were assigned to the planes (102), (107), (104), (110), (113), and (116); the cubic phase was also detected, since the peaks at around 2θ = 45.6° and 66.51° corresponding to the (400) and (440) planes were found. The Al2O3-TiO2 coating on the SS substrate treated at 700 °C for 5 h reveals diffraction peaks corresponding to the (101), (103), (200), and (204) planes of the anatase phase (JCPDS Card No. 21-1272) located at around 2θ = 25.28°, 37.54°, 24.46°, and 62.41°, respectively. Crystallographic planes (110), (101), (200), (111), (211), and (215) of the rutile phase (JCPDS Card No. 21-1276) were found at 2θ = 27.35°, 34.88°, 39.25°, 41.75°, 54.09°, and 76.07°. Sebastian Miszczak and Bożena Pietrzyk [46] reported that at 400 °C, the TiO2 anatase phase was fully crystallized and low-intensity peaks of the rutile phase were observed at 700 °C, indicating the beginning of the transformation of anatase to rutile; this finding is in total agreement with the phases present in the Al2O3-TiO2 coating.
The peaks located at approximately 29.8° and 35.5° correspond to the (220) and (311) planes, respectively, confirming the presence of magnetite. Additionally, a peak at (110) corresponding to ferrite (α-iron) is observed, which is typically formed during the cooling stage of 304 stainless steel. The presence of chromium oxide on the surface of 304 stainless steel is described as follows: During the heat treatment of 304 stainless steel, a protective chromium oxide layer forms on the surface, acting as a barrier against oxidation and corrosion. As the heat treatment progresses, chromium atoms from the oxide layer may diffuse away, creating voids or vacancies. Subsequently, iron atoms from the stainless steel matrix can migrate through these voids and react with oxygen from the atmosphere or the residual oxygen in the system. This reaction leads to the formation of magnetite. Therefore, the presence of magnetite indicates that the heat treatment has caused the transformation of the original chromium oxide layer.
The Rietveld refinement method was used to determine the phase composition of the Al2O3-TiO2 coating on a 304 SS substrate [47]. The phases used in the refinement correspond to those indexed in Figure 1 and elements identified through EDS analysis. The Rietveld analysis was executed using Profex Software (version 4.3.6), yielding the following statistical values: Rwp = 40.47, Rexp = 3.77, X2 = 1.57, and a Goodness of Fit (GoF) score of 1.25. Table 1 listed the phase percentages and lattice constants of the Al2O3-TiO2 coating on a 304 SS substrate obtained through the Rietveld refinement. Focusing exclusively on the Al2O3-TiO2 coating, its composition comprises 49.5% alumina and 59.5% titania. In the alumina component, 35.2% exists in the alpha (α) phase, while 14.3% is in the gamma (γ) phase. Regarding titania, 32.9% is in the anatase phase, with 17.6% in the rutile phase.

2.1.2. Scanning Electronic Microscopy

The field emission scanning electron microscopy (FE-SEM) images in Figure 2a reveal that the Al2O3-TiO2 coating is composed of particles with heterogeneous morphology due to their oxide nature. The tetrahedron form corresponds to Al2O3, while the rice-like morphology corresponds to TiO2. The particle size distribution of the Al2O3-TiO2 coating of 210 ± 90 nm considerably influences the catalysis due to its surface area [48]; it is worth noting that the particle size distribution agrees with the coating thickness estimated at 200 ± 50 nm with profilometry. The interfacial interaction between these two morphologies and their high adherence to the substrate form a heterojunction with the chromium oxide due to their differing electronic properties. This results in the Al2O3-TiO2 coating exhibiting distinctive electronic properties [18].
Furthermore, the elemental analysis of the Al2O3-TiO2 coating using the energy-dispersive X-ray spectroscopy (EDS) technique in Figure 2b confirms the presence of Fe and alloying elements in the substrate. 304 stainless steel is primarily composed of approximately 65–75% iron, 18–20% chromium, and 8–12% nickel, with trace amounts of manganese (<2%), silicon (<2%), phosphorus (<0.045%), and sulfur (<0.03%), while maintaining a low carbon content (<0.03%). Additionally, it may contain traces of phosphorus and sulfur [49]. The oxygen content of 21.46 wt.% is associated with the formation of Ti-O and Al-O bonds [50,51]. The EDS mapping reveals Ti, Al, and O, which is related to the Al2O3-TiO2 coating.

2.1.3. Atomic Force Microscopy

Images obtained with atomic force microscopy (AFM), which allowed the study of the effect of immersion and thermal treatment parameters on the morphological characteristics of the Al2O3-TiO2 coating with a scan size area of 2 μm, are shown in Figure 3. The surface shown in Figure 3a (height sensor) presents granular structures due to metal oxides on the 304 SS substrate, indicating efficient deposition conditions [52]. Figure 3b displays the topographic image of the Al2O3-TiO2 coating, revealing a homogeneous surface with high uniformity; no cracks or other delamination phenomena are observed [53]. The topography image shows that the root mean square roughness (RMSR) was around 85± 5 nm, calculated from the valleys and peaks profile, Figure 3b. Since the particle size distribution (210 ± 90 nm) of the coating is very similar to its thickness (250 ± 50 nm), particles prefer surface interaction; as a result, a monolayer of particles is formed. Therefore, the particles tend to self-assemble into a compact monolayer on the surface of the substrate rather than randomly distributed throughout the bulk of the coating. This preference for surface interaction can be due to the balance between the van der Waals and electrostatic forces between the particles and the surface. A monolayer coating has significant implications for the properties and performance of the film, including its optical and mechanical characteristics. The SEM images also elucidate the monolayer since the SEM reveals that mostly all particles are in contact with the substrate (dark region); therefore, the interfacial interaction with chromium and nickel of Al2O3 and TiO2 is high.

2.1.4. UV-Vis Spectroscopy

Figure 4a displays the UV–visible spectra of the Al2O3-TiO2 coatings. Two distinct bands of maximum absorption in the UV range, approximately around 250 and 350 nm, attributed to TiO2 particles, are observed. Martínez-Gómez and his research team [12] identified these bands as charge transfer processes from O2− to Ti4+. These bands corresponds to the excitation of valence band electrons (O, 2p) to the conduction band (Ti, 3d), a characteristic feature of the anatase phase. The broadband from 400 to 650 nm reveals visible light absorption, which is shown in Figure 4a. TiO2 nanoparticles doped with nonmetals, metallic cations, and metal oxides can alter the bandgap structure of TiO2 by introducing energy levels within the bandgap and enhancing visible light absorption. This, in turn, extends the lifetime of charge carriers and reduces recombination phenomena in the Al2O3-TiO2 coating [54,55]. The Tauc plot from the UV-vis spectrum of the Al2O3-TiO2 coating is shown in Figure 4b. The Tauc calculation revealed bandgap values of 3.4 and 2.9 eV corresponding to the rutile and anatase phases of TiO2, respectively. Values of 4.6 and 5.2 eV were observed for α-Al2O3 and γ-Al2O3, respectively. Those values are in the bandgap range for the anatase and rutile phases of TiO2 and the alpha and gamma phases of Al2O3 [56,57].
The interfacial iteration between TiO2 and Al2O3 and the combined factors, such as synthesis condition, substrate, thickness, and particle size distribution, promotes special spatial confinement. One of the main remarks in this work is the employment of visible light to activate photodegradation. It is worth mentioning that UV light usually activates TiO2. The thermal treatment at 700 °C promotes the sensitization of the 304 SS substrate, enabling Cr-rich carbide in the grain boundaries, and the chromium in the carbide reacts with oxygen from the surrounding environment forming chromium oxide. Additionally, during sensitization, chromium from the grain boundaries can migrate to the surface due to the high 18.24% chromium content in the metallic material, when chromium reaches the surface, it can interact with oxygen from the atmosphere, forming a passive oxide layer, primarily chromium oxide (Cr2O3). This migration may result in interfacial interactions with the Al2O3 and TiO2 phases in the coating, as both the coating structuration and crystallization originate from the same solution. It has been reported in the literature that heterojunctions with carbonaceous materials, metal nanoparticles, metal oxides, and phosphates enhance the photocatalytic activity, leading to better response in the visible spectrum and improvement of the TiO2 antibacterial activity [58,59]. In this context, both the interfacial interaction and heterojunction (considering chromium oxide’s nature to be a semiconductor) may explain the high efficiency of the reactor.
It is well-known that coupling alumina with titanium dioxide can enhance the absorption of visible light and broaden its photocatalytic activity. The sol–gel method employed for synthesizing and depositing alumina and titania onto the 304 SS substrate ensures a controlled and uniform distribution, leading to improved interfacial alumina–titania phases and their interaction with chromium oxide (as a result of the sensitization from the heat treatment). Interfaces often introduce defect states or energy level mismatches, altering the electronic band structure and affecting the bandgap.

2.2. Photocatalytic Degradation

2.2.1. MB Degradation

Photocatalytic Al2O3-TiO2 coating deposited on the six substrates inside the reactor reduced 97.3% of the MB concentration (Figure 5a). The high efficiency may be related to the high roughness that increases the retention of MB molecules; the coating thickness promotes a heterojunction with the chromium oxide and TiO2. The airflow increases the dissolved oxygen in the aqueous solution, thus enhancing the oxygen mass transfer efficiency. As a result, a high number of oxygen species, singlet oxygen (1O2), hydrogen peroxide (H2O2), hydroxyl radical (OH), and superoxide radical (O2●−) is formed by the Al2O3-TiO2 coating interaction with the visible light. Some authors have shown that the horizontal alignment of the lamps improves the photocatalytic degradation, and the light source with higher Watts increases the number of irradiated photons, improving the photocatalytic efficiency of the photocatalyst [59]. Komaraiah et al. [60] reported 92.03% MB for a TiO2 film after 4 h using a 200 W lamp and a volume of 70 mL, while Yu et al. [61] obtained up to 99.9% using a Cu2O/TiO2 thin film after 3 h, with a 500 W lamp and a total volume of 30 mL; in comparison, in this work, a total volume of 0.98 L was used, and four lamps of 4 W were employed.
For comparison, MB degradation was evaluated under photolysis (without substrates) and using non-heat-treated uncoated 304 SS substrates under the same experimental conditions (Figure 5a). Photolysis resulted in 2.6% of MB degradation, while the uncoated 304 SS substrate showed 4.9% of MB degradation. In contrast, the blue line in the same figure represents a significant increase in MB degradation, reaching 12.5%, when using the heat-treated 304 SS substrate (subjected to the same heat treatment for obtaining the Al2O3-TiO2 coating). This remarkable increase is attributed to chromium oxide and magnetite formation promoted by the sensitization of stainless steel in the heat treatment, as evidenced in the XRD section.
Table 2 presents a comprehensive review of various studies investigating the photocatalytic efficiency of TiO2, both in its undoped form and when doped with metals and nonmetals, as well as its heterojunctions with other metal oxides. Notably, the doping of TiO2 with metals [60,62,63,64,65,66] has shown significantly high photocatalytic efficiencies, comparable to the heterojunctions of TiO2 with metal oxides [48,66,67,68] suspended in the medium.
The immobilization of the catalyst can reduce photocatalytic activity due to the loss of surface area. Despite this drawback, utilizing coatings on 304 SS substrates offers practicality since no recovery step for the catalyst is required. Moreover, the interfacial interaction between the coating and the substrate yields beneficial effects in photocatalysis. Al2O3-TiO2 coatings on stainless steel substrates have demonstrated superior photocatalytic activity compared to the same powder system [48]. This enhanced photocatalytic activity of Al2O3-TiO2 may be attributed to the presence of oxides of Cr and Fe, resulting from sensitization during heat treatment, which synergistically act with the photocatalyst. One of the main findings in this work is the high efficiency (97%) achieved by only using visible light provided by four lamps of 4 W, and an airlift reactor configuration of 0.98 L.

2.2.2. TOC Quantification

The Al2O3-TiO2 coating with enhanced photocatalytic activity resulted in higher total organic degradation due to its ability to effectively produce reactive oxygen species, which can then degrade organic compounds. Since TOC monitoring is one of the most important parameters to be considered for drinking water and wastewater facilities, it is critical in helping to optimize treatment processes. TOC is a helpful technique for detecting organic contaminants such as petroleum products, organic acids, pesticides, pathogens, etc. TOC levels can be used to measure treatment efficacy and as an indicator of contamination [69]. For this reason, the Al2O3-TiO2 coating in the airlift reactor was tested to degrade organic carbon. The degradation of organic contaminants in a sample from the local WWTP was 96.51%, measured as TOC, Figure 5b. The residual concentration of 3.49% represents 35.6 mg/L of TOC; the result is within the interval of the maximum permissible limits of the standards for the discharge of treated wastewater in national assets or can be discharged into the sewage system or for reuse in public service (SEMARNAT, 1996; 1997) according to ECOL-001, Standard 002 and 003 (SEMARNAT, 1996; 1997). In comparison, the WWTPs with a treatment cycle of 12 of 24 h obtain final concentrations from 64.9 to 134.5 mg/L of TOC [70,71]. The degradation of organic compounds followed a pseudo-first-order rate kTOC = 0.0118 min−1. The rate constant for MB degradation is comparable to Cu2O/TiO2 thin film [72], which diminished from 0.013 to 0.010 min−1 with eight reuses, and the c rate constant of 0.0108 min−1 for TiO2 nanorod array thin films on fluorine-doped tin oxide (FTO) glass under 254 nm irradiation [73].

2.2.3. Bactericidal Activity

Several mechanisms have been proposed to explain the bactericidal activity of TiO2 nanocomposites. Oxygen species formed by TiO2 nanoparticles can enhance the peroxidation of the polyunsaturated phospholipids in the bacterial membrane and promote its disruption [74].
Furthermore, the disruption of both the nucleus and mitochondrial membranes can lead to DNA degradation and protein denaturation. Simultaneously, the holes formed in the valence band of TiO2 nanoparticles have the capacity to capture electrons from coenzyme A (CoA), a vital molecule involved in various cellular processes. This electron capture results in CoA dimerization, where two CoA molecules combine [75]. When CoA loses an electron and forms a dimer, it disrupts the essential cellular respiration process, which is critical for bacterial survival. This disruption of cell respiration represents one of the mechanisms by which TiO2 nanoparticles impact bacterial cells.
Before visible light exposure, 2400 MPN of total and fecal coliform bacteria and 1.56 × 103 CFU of Salmonella spp. were in the wastewater sample. The bacterial disinfection by photocatalysis is different from organic pollutant degradation, Figure 5b; the bacterial population was drastically unchanged during the first hour, which could be due to the fact that bacteria have enzymatic protection and recovery mechanisms to overcome oxidative stress, catalyzing reactions that prevent reactive oxygen species’ accumulation. In the case of organic pollutants, the photocatalytic efficiency corresponds to their disappearance and sometimes to their mineralization [76]. In contrast, the total inactivation of a microorganism depends on the modification of the entire cellular structure due to partial decomposition of the outer membrane followed by the disordering of the cytoplasmic membrane, resulting in cell death promoted by the accumulation of reactive oxygen species such as O2●−, OH, and H2O2, Figure 6 [77]. After this period, the bactericidal effect can be ascribed to the inhibition caused by blocking nutrient uptake because the organic contaminants decreased by almost 40% measured as TOC, and this combined with the oxidative stress caused by OH and O2●− species, generated by photoactivation of the Al2O3-TiO2 coating, promoted disruption of the cell membrane, DNA degradation, and protein denaturation [78].
The low efficiency of the Al2O3-TiO2 coating against fecal and total coliforms is attributed to the presence of various bacteria with distinct characteristics. As a result, the Al2O3-TiO2 coating only affects certain bacteria similar to Salmonella spp., while other bacteria remain unaffected. Since the bubbling and irradiation were maintained during the experiment, the mechanism and bacterial reduction rate should have remained constant. To bolster this claim, mathematical equations were defined and used.
As fecal and total coliforms have bacteria that can be affected or unaffected by the Al2O3-TiO2 coating, XDB is the bacteria fraction that the Al2O3-TiO2 coating can reduce in fecal and total coliforms. The fraction reduction of Salmonella spp. at a certain time ( y s t ) is used in the computation as shown in Equation (1):
y s t = C F U 0 C F U t C F U 0
where C F U 0 stands for colony forming units at the initial time, and C F U t stands for colony forming units at time “t”. To study the behavior of the fecal and total coliform reduction, Equation (2) was employed and is as follows:
M P N t = M P N 0 ( M P N 0 × y s t × X D B )
where M P N t is the most probable number at time “t”, and M P N 0 is the most probable number at the initial time. If all fecal and total coliforms were composed of Salmonella spp. or bacteria with similar features, then XDB = 1; therefore, a similar equation of y s t could be obtained. The reduction percentage of coliforms can be determined with Equation (3):
% D c = M N P t M P N 0 × 100
Equation (4) is obtained by combining Equations (2) and (3):
% D c t = 1 1 × y s t × X D B × 100
By equating the calculated percentage reduction of total coliforms ( % D c t ) with the experimental value at the end of the process, the equation enabled the calculation of X D B . Surprisingly, the plotting of these data revealed a square correlation factor of 0.9765, which provides strong evidence for the high efficiency of the Al2O3-TiO2 coating. The obtained XDB value of 0.355 indicates that only 35.5% of the coliforms are suitable for reduction, further confirming the effectiveness of the coating.

3. Materials and Methods

3.1. Coating Preparation Using the Sol–Gel Route

First, 1 M NaOH (Meyer ≥ 97%) solution was added dropwise to 50 mL of 0.1 M Al2(SO4)3·12H2O (Kemira Al-C 2-5, 90%) until pH = 10 was reached (to favor the formation of aluminum hydroxides and oxo-hydroxides). After adjusting the pH, the suspension was kept under stirring at 70 °C for 5 h. Thereafter, the formed colloidal suspension was washed three times with a mixture of ethanol/water 50:50 v/v to remove excess SO42− ions. The wet solid was re-dispersed in 20 mL of ethanol; this suspension was aged for 24 h.
In the second step, 20 mL of HNO3 solution, 0.3 M (Meyer, 65%), was added dropwise to 20 mL of a 15% v/v titanium butoxide Ti[OC(CH3)3]4 (Sigma Aldrich, St. Louis, MI, USA, 97%) in ethanol (Sigma Aldrich, 99.9). The suspension was stirred for 2 h to achieve the completion of the hydrolysis–condensation reaction. Finally, the aluminum oxo-hydroxides obtained in the previous stage were added to the forming [TiO(OH)2]n gel and kept under vigorous stirring until a milky colloidal suspension appeared (Al/Ti ratio of 1.5).
The circular substrates made of AISI 304 SS, with a diameter of 5 cm and orifice diameter of 2 mm, were first washed with detergent soap and then treated with acetone followed by ethanol using sonication at 400 kHz for 20 min. After this, the substrates were stored for 1 h in isopropyl alcohol. The coating was obtained by immersing and withdrawing the substrates at a speed of 20 mm/min. After each deposition, the substrates were dried at 80 °C for 1 h; and this deposition process was repeated five times. The substrate with five depositions was thermally treated at 180 °C for 1 h and annealed at 700 °C for 5 h.

3.2. Structural Characterization Measurements

The structural study of the Al2O3-TiO2 coating was carried out first, with a Bruker D8 Advance Diffractometer (Bruker, Karlsruhe, Germany) using Cu Kα radiation (1.54184 Å) to determine the crystallinity and crystal phase of the produced coating. X-ray diffraction patterns were recorded at room temperature in the 2θ angle interval ranging from 20° to 80° with a step of 0.02 s−1. The analysis of the morphology and size of the particles was performed by employing a field emission scanning electron microscope (FE-SEM) Hitachi SU5000 (Hitachi High-Tech, Tokyo, Japan); for the EDS analysis, a coupled Bruker’s Quantax XFlash 6/60 was used to investigate the compositional aspects of the coating. To characterize the homogeneity of the deposits and texture quality of Al2O3-TiO2, an atomic force microscope NanoScope II (Veeco/Digital Instruments, Santa Barbara, CA, USA) was employed, and the coating thickness was measured by means of a Dektak3 profilometer (Veeco Instruments Inc., Plainview, NY, USA) coupled to the NanoScope II. XRD, SEM, FE-SEM, and AFM characterizations of the Al2O3-TiO2 coating on the 304 SS substrate were conducted after all the tests in this study to demonstrate its stability. The optical characteristics of the Al2O3-TiO2 coating were examined with a Perkin Elmer model Lambda 35 UV-vis spectrophotometer (PerkinElmer, Waltham, MA, USA) from 200 to 900 nm wavelength to register absorbance values of the coating scratch. The bandgap energy was obtained using the Tauc Equation (5), which relates the absorption coefficient to the energy of incident radiation and the bandgap. The intersection of the linear part of the curve α · E r vs. E with the x-axis represents α · E r = 0 , and therefore, E = Eg [79]:
α · E r = A i E E g
where α is the absorbance, arbitrary units; E is photon energy, eV; A i is a proportionality constant, eV; E g is the bandgap, eV; and r is a coefficient that classifies indirect allowed transitions ( r = 2), direct allowed transitions ( r = 1/2), direct forbidden transitions ( r = 3/2), and indirect forbidden transitions ( r = 3).
E = h ν = h c λ
The energy of incident radiation E was calculated from the incident beam wavelengths using Equation (6) where h is Planck’s constant = 4.136 × 10−15 eV s; c is the speed of light in a vacuum = 2.998 × 1017 nm/s; and λ is the wavelength in nm.

3.3. Photocatalytic Degradation

A semicontinuous flow photocatalytic reactor was designed to measure all the degradation reactions with total volume of 0.980 L (internal diameter of 5 cm and height of 50 cm), Figure 7. The six substrates with Al2O3-TiO2 coating were placed in a transverse position, separated from each other by 5 cm long tubes guided by a 50 cm long bar central to the reactor; the bar and tubes were made of 304 SS. An additional substrate, measuring 6 cm in diameter, is positioned at the upper edge of the reactor, serving as a guide stop. The linear ascent of bubbles at an average speed of 2.666 cm/s (Re = 1655) generated the convection currents that kept the system homogenized to favor the adsorption of the contaminant on the photocatalyst surface. The average temperature was 30 ± 0.5 °C. Four Phillips LED lamps (4 W, 400 lumens, MR16 with emission in 410–760 nm range and a maximum peak at around 600 nm) were used for visible light irradiation.

3.3.1. MB Degradation

In the reactor, a 6 ppm MB solution was kept under stirring for 30 min in the absence of light, using upward airflow of 5 L/min to allow adsorption–desorption equilibrium between the Al2O3-TiO2 coating surface and contaminant. Six substrates (46.09 cm2 of coating area per substrate) were placed in the reactor. Subsequently, 5 mL of sample was obtained to measure the initial MB concentration, monitoring the absorption band at 668 nm, corresponding to the maximal absorption peak of MB. Four Phillips LED lamps were turned on to start the photocatalytic degradation after collecting the sample. A 2 mL volume of samples was collected every hour to measure the MB concentration. Four experiments were performed under the same conditions.

3.3.2. TOC Quantification

An aliquot of 1 mL was collected every hour from the reactor containing 0.980 L of residual water from a WWTP. The employed residual water had a pH, total solids, and hardness of 7.5 ± 0.3, 550 ± 048 mg/L, and 109.0 ± 0.12 mg/L, respectively. After collecting the first 1 mL sample, the six LED lamps were turned on. The aliquot was filtered to remove suspended solids, and its pH was adjusted to 2 using a 20 v/v% H2SO4 solution to eliminate microorganism reactions and hence exhibited an accurate TOC measurement; a 50 µL volume of each aliquot was injected into a TOC-V CSN (Shimadzu Brand, Kyoto, Japan) with an automatic ASI-V counter at room temperature (25 °C). A calibration curve from 0 to 500 mg/L of potassium biphthalate (C8H5KO4) was used. The procedure was performed according to ISO 8245:1999 [80], which guides the TOC determination in wastewater, applicable for water samples with organic carbon content ranging from 0.3 mg/L to 1000 mg/L. The efficiency of the degradation of contaminants was calculated with Equation (7):
%   d e g r a d a t i o n = C o C C o × 100
where C o is the initial concentration of the contaminant before turning on the lamps, and C is the concentration of the contaminant after time t.

3.3.3. Bactericidal Activity

The bactericidal activity of the Al2O3-TiO2 coating was measured by fecal and total coliform counts according to standard methods for wastewater testing using the most probable number (MPN) method, based on the ability of this microbial group to ferment lactose with acid and gas production when incubated at 35 °C ± 1 °C for 48 h, using a culture medium containing bile salts.
An aliquot of 50 mL was collected every hour from the 0.98 L reactor to inoculate five tubes containing 10.0 mL of sodium lauryl sulfate broth with 10.0 mL of sample each; gas formation (displacement of the medium in a Durham bell) after 48 h of incubation at 35 ± 1 °C was considered a presumptive positive test. Confirmation of the presence of total and fecal coliforms was performed by reseeding two drops taken with a Pasteur pipette from each tube that tested positive in the presumptive test in brilliant green bile broth; the appearance of turbidity (microbial growth) and gas production in the Durham hood following the 48 h incubation period at 35 ± 1 °C were recorded as positive tests. The MPN/100 mL was calculated according to NOM-210-SSA1-2014 [81], appendix H, with the MPN Index, 95% confidence, for the combination of positive and negative results when using 5 to 10 tubes with 10 mL of water sample.
Additionally, the presence of Salmonella spp. was analyzed by determining the plate count (Colony Forming Units, CFU). A 0.5 mL volume of each positive tube from the confirmatory coliform test was spread over the entire surface of plates containing Salmonella-Shigella (SS) agar and incubated for 48 h at 35 °C.

4. Conclusions

Al2O3-TiO2 coatings were successfully deposited on perforated 304 SS substrates using the sol–gel dip-coating technique. Thermal treatments at 180 °C for 1 h and at 700 °C for 5 h, respectively, promoted the TiO2 anatase–rutile phases and Al2O3 α-γ phases. An Al2O3-TiO2 coating monolayer on the 304 SS substrate was elucidated due to the similarity between the particle size distribution (250 ± 50 nm) and the coating thickness (250 ± 50 nm). The Al2O3-TiO2 coating formed a heterojunction with chromium oxide on the 304 stainless steel surface due to the difference in their electronic properties, resulting in better response in the visible spectrum to degrade organic pollutants and obtain improved antibacterial activity. The study demonstrated that five perforated substrates with Al2O3-TiO2 coatings and four 4 Watt lamps in a 0.980 L airlift reactor efficiently reduced MB and total organic contaminants in a WWTP sample by 97.3% with UV-vis absorbance and by 96.51% with TOC determination, respectively. Root mean square roughness of the perforated Al2O3-TiO2 coating on the 304 SS substrate of 85 ± 5 nm enhanced the retention of Salmonella ssp., promoting longer contact between the cell membrane and the Al2O3-TiO2 coating. At the same time, the airflow favored the formation of reactive oxygen species in the aqueous solution, leading to high degradation of organic pollutants and improvement of the antibacterial activity; in this study, a reduction of 94.23% of Salmonella spp. was achieved in 5 h. The low efficiency of Al2O3-TiO2 against total and fecal coliforms is attributed to differences in their bacterial compositions. Mathematical analysis revealed that only 35.5% of total and fecal coliforms have mechanism of action similar to that of Salmonella ssp., which was supported by a square correlation factor of 0.9919 between the calculated and experimental reduction percentage of total and fecal coliforms. As a perspective, future research may involve the utilization of ascorbic acid (OH scavenger), 2-propanol (O2●− scavenger), and ammonium oxalate (AO) (hole h+ scavenger) to elucidate the photocatalytic activity of Al2O3-TiO2 coating in organic contaminant degradation and to identify the oxygen reactive species generated with light exposure.

Author Contributions

Conceptualization, M.A.C.-G. and A.G.-H.; methodology, M.A.C.-G. and M.Q.-C.; software A.G.-H. and J.R.-M.; validation, P.V.-S. and A.G.-H.; formal analysis A.G.-H., P.V.-S. and E.S.-B.; investigation, M.A.C.-G., E.S.-B. and A.G.-H.; resources, M.A.C.-G., A.G.-H., J.R.-M. and M.Á.B.-M.; data curation, A.G.-H., J.R.-M. and M.Á.B.-M.; writing—original draft preparation, M.A.C.-G., A.G.-H. and I.V.L.; writing—review and editing, A.G.-H. and I.V.L.; supervision, A.G.-H. and I.V.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

The authors acknowledge IPN, UTTEC, and UAM for supporting this work.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. ONU-DAES. International Decade for Action Water for Life 2005–2015. 2021. Available online: https://www.un.org/waterforlifedecade/index.shtml (accessed on 15 January 2023).
  2. UNESCO. United Nations World Water Development Report 2020: Water and Climate Change; UNESCO: Paris, France, 2020; Available online: https://www.unesco.org/en/wwap/wwdr/2020 (accessed on 20 January 2023).
  3. Ismail, A.A.; Abdelfattah, I.; Atitar, M.F.; Robben, L.; Bouzid, H.; Al-Sayari, S.A.; Bahnemann, D.W. Photocatalytic degradation of imazapyr using mesoporous Al2O3–TiO2 nanocomposites. Sep. Purif. Technol. 2015, 145, 147–153. [Google Scholar] [CrossRef]
  4. Rueda-Marquez, J.J.; Levchuk, I.; Fernández-Ibañez, P.; Sillanpää, M.A. Critical review on application of photocatalysis for toxicity reduction of real wastewaters. J. Clean. Prod. 2020, 258, 120694. [Google Scholar] [CrossRef]
  5. Švagelj, Z.; Mandić, V.; Ćurković, L.; Biošić, M.; Žmak, I.; Gaborardi, M. Titania-coated alumina foam photocatalyst for memantine degradation derived by replica method and sol-gel reaction. Materials 2020, 13, 227. [Google Scholar] [CrossRef]
  6. Lucas, M.S.; Peres, J.A.; Li-Puma, G. Advanced oxidation processes for water and wastewater treatment. Water 2021, 13, 1309. [Google Scholar] [CrossRef]
  7. Shafiq, I.; Shafique, S.; Akhter, P.; Abbas, G.; Qurashi, A.; Hussain, M. Efficient catalyst development for deep aerobic photocatalytic oxidative desulfurization: Recent advances, confines, and outlooks. Catal. Rev.-Sci. Eng. 2022, 64, 789–834. [Google Scholar] [CrossRef]
  8. Khatri, J.; Nidheesh, P.V.; Anantha Singh, T.S.; Kumar, M.S. Advanced oxidation processes based on zero-valent aluminium for treating textile wastewater. Chem. Eng. J. 2018, 348, 67–73. [Google Scholar] [CrossRef]
  9. Khan, A.H.; Khan, N.A.; Ahmed, S.; Dhingra, A.; Singh, C.P.; Khan, S.U.; Mohammadi, A.A.; Changani, F.; Yousefi, M.; Alam, S.; et al. Application of advanced oxidation processes followed by different treatment technologies for hospital wastewater treatment. J. Clean. Prod. 2020, 269, 122411. [Google Scholar] [CrossRef]
  10. Dewil, R.; Mantzavinos, D.; Poulios, I.; Rodrigo, M.A. New perspectives for advanced oxidation processes. J. Environ. Manag. 2017, 195, 93–99. [Google Scholar] [CrossRef]
  11. Ahmed, S.; Raul, M.G.; Martens, W.N.; Brown, R.; Hashib, M.A. Advances in heterogeneous photocatalytic degradation of Phenols and dyes in wastewater: A Review. Water Air Soil Pollut. 2011, 215, 3–29. [Google Scholar] [CrossRef]
  12. Wawrzyniak, B.; Morawski, A.W. Solar-light-induced photocatalytic decomposition of two azo dyes on new TiO2 photocatalyst containing nitrogen. Appl. Catal. B Environ. 2006, 62, 150–158. [Google Scholar] [CrossRef]
  13. Da Silva, S.W.; Klauck, C.R.; Siqueira, M.A.; Bernardes, A.M. Degradation of the commercial surfactant nonylphenol ethoxylate by advanced oxidation processes. J. Hazard. Mater. 2015, 282, 241–248. [Google Scholar] [CrossRef] [PubMed]
  14. Rey, A.; García-Muñoz, P.; Hernández-Alonso, M.D.; Mena, E.; García-Rodríguez, S.; Beltrán, F.J. WO3–TiO2 based catalysts for the simulated solar radiation assisted photocatalytic ozonation of emerging contaminants in a municipal wastewater treatment plant effluent. Appl. Catal. B Environ. 2014, 154–155, 274–284. [Google Scholar] [CrossRef]
  15. Aguinaco, A.; Amaya, B.; Ramírez-Del-Solar, M. Facile fabrication of Fe-TiO2 thin film and its photocatalytic activity. Environ. Sci. Pollut. Res. 2022, 29, 23292–23302. [Google Scholar] [CrossRef] [PubMed]
  16. Chairungsri, W.; Subkomkaew, A.; Kijjanapanich, P.; Chimupala, Y. Direct dye wastewater photocatalysis using immobilized titanium dioxide on fixed substrate. Chemosphere 2022, 268, 131762. [Google Scholar] [CrossRef] [PubMed]
  17. Nakata, K.; Fujishima, A. TiO2 photocatalysis: Design and applications. J. Photochem. Photobiol. C Photochem. Rev. 2012, 13, 169–189. [Google Scholar] [CrossRef]
  18. Behpour, M.; Atouf, V. Study of the photocatalytic activity of nanocrystalline S, N-codoped TiO2 thin films and powders under visible and sun light irradiation. Appl. Surf. Sci. 2012, 258, 6595–6601. [Google Scholar] [CrossRef]
  19. Falk, G.; Borlaf, M.; López-Muñoz, M.J.; Fariñas, J.C.; Neto, J.B.R.; Moreno, R. Microwave-assisted synthesis of Nb2O5 for photocatalytic application of nanopowders and thin films. J. Mater. Res. 2017, 32, 3271–3278. [Google Scholar] [CrossRef]
  20. Nisar, U.; Muralidharan, N.; Essehli, R.; Amin, R.; Belharouak, I. Valuation of surface coatings in high-energy density lithium-ion battery cathode materials. Energy Storage Mater. 2021, 38, 309–328. [Google Scholar] [CrossRef]
  21. Sahu, S.K.; Panthi, D.; Du, Y. Development of Multiple Ceramic Coatings on Porous Tubular Metal Support. ECS Trans. 2023, 6, 249. [Google Scholar] [CrossRef]
  22. Kim, S.G.; Yoon, S.P.; Nam, S.W.; Hyun, S.H.; Hong, S.A. Fabrication and characterization of a YSZ/YDC composite electrolyte by a sol–gel coating method. J. Power Source 2002, 110, 222–228. [Google Scholar] [CrossRef]
  23. Wu, H.; Wu, Y.; Yan, M.; Tu, B.; Li, Y. Microstructure and mechanical properties of surface coating prepared on grade 2 titanium by Ni-B composite electroplating and laser cladding. Laser Technol. 2023, 164, 109498. [Google Scholar] [CrossRef]
  24. Hui, R.; Wang, Z.; Yick, S.; Maric, R.; Ghosh, D. Fabrication of ceramic films for solid oxide fuel cells via slurry spin coating technique. J. Power Source 2007, 172, 840–844. [Google Scholar] [CrossRef]
  25. Ogumi, Z.; Uchimoto, Y.; Tsuji, Y.; Takehara, Z.I. Preparation of thin yttria-stabilized zirconia films by vapor-phase electrolytic deposition. Solid State Ion. 1992, 58, 345–350. [Google Scholar] [CrossRef]
  26. Smeacetto, F.; Salvo, M.; Ajitdoss, L.C.; Perero, S.; Moskalewicz, T.; Boldrini, S.; Ferraris, M. Yttria-stabilized zirconia thin film electrolyte produced by RF sputtering for solid oxide fuel cell applications. Mater. Lett. 2010, 64, 2450–2453. [Google Scholar] [CrossRef]
  27. García-Sánchez, F.; Peña, J.; Ortiz, A.; Santana, G.; Fandiño, J.; Bizarro, m.; Cruz-Gandarilla, F.; Alonso, J.C. Nanostructured YSZ thin films for solid oxide fuel cells deposited by ultrasonic spray pyrolysis. Solid State Ionics 2008, 179, 243–249. [Google Scholar] [CrossRef]
  28. Gołąbiewska, A.; Kobylański, M.P.; Zaleska-Medynska, A. Fundamentals of Metal Oxide-Based Photocatalysis in Book Metal Oxide-Based Photocatalysis Fundamentals and Prospects for Application; Zaleska-Medynska, A., Ed.; Elsevier: Amsterdam, The Netherlands, 2018; pp. 3–50. [Google Scholar] [CrossRef]
  29. Tian, L.; Qian, L.; Yao, X. Preparation of Sn4+-doped titanium dioxide photocatalytic films on 304 stainless steel by duplex treatment. J. Mater. Eng. Perform. 2014, 23, 1790–1798. [Google Scholar] [CrossRef]
  30. Lee, C.S.; Kim, J.; Son, J.Y.; Maeng, W.J.; Jo, D.H.; Choi, W.; Kim, H. Plasma-enhanced ALD of TiO2 thin films on SUS 304 stainless steel for photocatalytic application. J. Electrochem. Soc. 2009, 156, D188–D192. [Google Scholar] [CrossRef]
  31. Soylu, A.M.; Polat, M.; Erdogan, D.A.; Say, Z.; Yildiri, C.; Birer, Ö.; Ozensoy, E. TiO2–Al2O3 binary mixed oxide surfaces for photocatalytic NOx abatement. Appl. Surf. Sci. 2014, 318, 142–149. [Google Scholar] [CrossRef]
  32. 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. B Eng. 2014, 58, 147–151. [Google Scholar] [CrossRef]
  33. Montalvo, D.; Corro, G.; Bañuelos, F.; Olivares-Xometl, O.; Arellanes, P.; Pal, U. Selective alcohols production through CO2 photoreduction using Co3O4/TiO2 photocatalyst exploiting synergetic interactions between Ti3+, Co2+ and Co3+. Appl. Catal. B Environ. 2023, 330, 122652. [Google Scholar] [CrossRef]
  34. Parnicka, P.; Lisowski, W.; Klimczuk, T.; Łuczak, J.; Żak, A.; Zaleska-Medynska, A. Visible-light-driven lanthanide-organic-frameworks modified TiO2 photocatalysts utilizing up-conversion effect. Appl. Catal. B Environ. 2021, 291, 120056. [Google Scholar] [CrossRef]
  35. 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 co-catalysts on the facets of rutile TiO2. J. Catal. 2016, 337, 36–44. [Google Scholar] [CrossRef]
  36. 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]
  37. 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]
  38. Magnone, E.; Kim, K.M.; 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]
  39. Phattepur, H.; Hiremath, P.G. Fabrication of Al2O3 supported TiO2 membranes for photocatalytic applications. Mater. Today Proc. 2022, 65, 3694–3699. [Google Scholar] [CrossRef]
  40. Martinez-Gómez, C.; Rangel-Vázquez, I.; Zarraga, R.; Del Ángel, G.; Ruíz-Camacho, B.; Tzompantzi, F.; Vidal-Robles, E.; Perez-Larios, A. Photodegradation and Mineralization of Phenol Using TiO2 Coated γ-Al2O3: Effect of Thermic Treatment. Processes 2022, 10, 1186. [Google Scholar] [CrossRef]
  41. Anderson, C.A.; Bard, A.J. Improved photocatalytic activity and characterization of mixed TiO2/SiO2 and TiO2/Al2O3 materials. J. Phys. Chem. B 1997, 101, 2611–2616. [Google Scholar] [CrossRef]
  42. Lee, S.; Huang, E.-W.; Woo, W.; Yoon, C.; Chae, H.; Yoon, S.-G. Dynamic Strain Evolution around a Crack Tip under Steady- and Overloaded-Fatigue Conditions. Metals 2015, 5, 2109–2118. [Google Scholar] [CrossRef]
  43. 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]
  44. 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]
  45. 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]
  46. Miszczak, S.; Pietrzyk, B. Anatase–rutile transformation of TiO2 sol–gel coatings deposited on different substrates. Ceram. Int. 2015, 41, 7461–7465. [Google Scholar] [CrossRef]
  47. Marques, V.E.C.; Manfroi, L.A.; Vieira, A.A.; de Jesús Pereira, A.L.; das Chagas Marques, F.; Vieira, L. Crystalline Structure, Morphology, and Adherence of Thick TiO2 Films Grown on 304 and 316L Stainless Steels by Atomic Layer Deposition. Coatings 2023, 13, 757. [Google Scholar] [CrossRef]
  48. Abbas, N.; Shao, G.N.; Haider, M.S.; Imran, S.M.; Park, S.S.; Kim, H.T. Sol–gel synthesis of TiO2-Fe2O3 systems: Effects of Fe2O3 content and their photocatalytic properties. J. Ind. Eng. Chem 2016, 39, 112–120. [Google Scholar] [CrossRef]
  49. Da Trindade, C.; 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]
  50. Sridevi, A.; Krishnamohan, S.; Thairiyaraja, M.; Prakash, B.; Yokeshwaran, R. Visible-light driven γ-Al2O3, CuO and γ-Al2O3/CuO nanocatalysts: Synthesis and enhanced photocatalytic activity. Inorg. Chem. Commun. 2022, 138, 109311. [Google Scholar] [CrossRef]
  51. 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. S. Afr. J. Chem. Eng. 2020, 33, 58–66. [Google Scholar] [CrossRef]
  52. Mwema, F.M.; Oladijo, O.P.; Sathiaraj, T.S.; Akinlabi, E.T. Atomic force microscopy analysis of surface topography of pure thin aluminum films. Mater. Res. Express 2018, 5, 046416. [Google Scholar] [CrossRef]
  53. Luo, H.; Xiang, Y.; Tian, T.; Pan, X. An AFM-IR study on surface properties of nano-TiO2 coated polyethylene (PE) thin film as influenced by photocatalytic aging process. Sci. Total Environ. 2021, 757, 143900. [Google Scholar] [CrossRef]
  54. Karunakaran, C.; Abiramasundari, G.; Gomathisankar, P.; Manikandan, G.; Anandi, V. Cu-doped TiO2 nanoparticles for photocatalytic disinfection of bacteria under visible light. J. Colloid Interface Sci. 2010, 352, 68–74. [Google Scholar] [CrossRef]
  55. Yadav, H.M.; Kim, J.S.; Pawar, S.H. Developments in photocatalytic antibacterial activity of nano TiO2: A review. Korean J. Chem. Eng. 2016, 33, 1989–1998. [Google Scholar] [CrossRef]
  56. Dhawale, V.P.; Late, D.J.; Kulkarni, S.D. Synthesis, characterization of α-Al2O3 nanoparticles and its application in decolorization of methyl orange azo dye in the presence of UV light. J. Nanosci. Nanotechnol. 2019, 5, 580–583. [Google Scholar] [CrossRef]
  57. Prashanth, P.A.; Raveendra, R.S.; Krishna, R.H.; Ananda, S.; Bhagya, N.P.; Nagabhushana, B.M.; Lingaraju, K. Synthesis, characterizations, antibacterial and photoluminescence studies of solution combustion-derived α-Al2O3 nanoparticle. J. Ceram. Process. Res. 2015, 3, 345–351. [Google Scholar] [CrossRef]
  58. Huang, Z.; Zhao, S.; Yu, Y. Experimental method to explore the adaptation degree of type-II and all-solid-state Z-scheme heterojunction structures in the same degradation system. Chin. J. Catal. 2020, 41, 1522–1534. [Google Scholar] [CrossRef]
  59. Liu, X.; Iocozzia, J.; Wang, Y.; Cui, X.; Chen, Y.; Zhao, S.; Li, Z.; Lin, Z. Noble metal–metal oxide nanohybrids with tailored nanostructures for efficient solar energy conversion, photocatalysis and environmental remediation. Energy Environ. Sci. 2017, 10, 402–434. [Google Scholar] [CrossRef]
  60. Komaraiah, D.; Madhukar Vijayakumar, Y.; Ramana Reddy, M.V.; Sayanna, R. Photocatalytic degradation study of methylene blue by brookite TiO2 thin film under visible light irradiation. Mater. Today Proc. 2016, 3, 3770–3778. [Google Scholar] [CrossRef]
  61. Yu, X.; Kou, S.; Nie, J.; Zhang, J.; Wei, Y.; Niu, N.; Yao, B. Preparation and performance of Cu2O/TiO2 nanocomposite thin film and photocatalytic degradation of Rhodamine B. Water Sci. Technol. 2018, 78, 913–924. [Google Scholar] [CrossRef]
  62. Markovskaya, D.V.; Zhurenok, A.V.; Kurenkova, A.Y.; Kremneva, A.M.; Saraev, A.A.; Zharkov, S.M.; Kozlova, E.A.; Kaichev, V.V. New titania-based photocatalysts for hydrogen production from aqueous-alcoholic solutions of methylene blue. RSC Adv. 2020, 10, 34137–34148. [Google Scholar] [CrossRef]
  63. Reda, S.M.; Khairy, M.; Mousa, M.A. Photocatalytic activity of nitrogen and copper doped TiO2 nanoparticles prepared by microwave-assisted sol-gel process. Arab. J. Chem. 2020, 13, 86–95. [Google Scholar] [CrossRef]
  64. Ganesh, I.; Gupta, A.K.; Kumar, P.P.; Sekhar, P.S.C.; Radha, K.; Padmanabham, G.; Sundararajan, G. Preparation and characterization of Ni-doped TiO2 materials for photocurrent and photocatalytic applications. Sci. World J. 2012, 2012, 127326. [Google Scholar] [CrossRef] [PubMed]
  65. Sun, L.; Zhai, J.; Li, H.; Zhao, Y.; Yang, H.; Yu, H. Study of homologous elements: Fe, Co, and Ni dopant effects on the photoreactivity of TiO2 nanosheets. ChemCatChem 2014, 6, 339–347. [Google Scholar] [CrossRef]
  66. Xu, J.; Wang, F.; Liu, W.; Cao, W. Nanocrystalline N-doped TiO2 powders: Mild hydrothermal synthesis and photocatalytic degradation of phenol under visible light irradiation. Int. J. Photoenergy 2013, 2013, 616139. [Google Scholar] [CrossRef]
  67. Sánchez Mora, E.; Gómez Barajas, E.; Rojas Rojas, E.; Silva González, R. Morphological, optical and photocatalytic properties of TiO2–Fe2O3 multilayers. Sol. Energy Mater. Sol. Cells 2007, 91, 1412–1415. [Google Scholar] [CrossRef]
  68. Zhao, D.; Chen, C.; Wang, Y.; Ma, W.; Zhao, J.; Rajh, T.; Zang, L. Enhanced photocatalytic degradation of dye pollutants under visible irradiation on Al(III)-modified TiO2: Structure, interaction, and interfacial electron transfer. Environ. Sci. Technol. 2008, 42, 308–314. [Google Scholar] [CrossRef]
  69. Hernández-Gordillo, A.; Bizarro, M.; Gadhi, T.A. Good practices for reporting the photocatalytic evaluation of a visible-light active semiconductor: Bi2O3, a case study. Catal. Sci. Technol. 2019, 9, 1476–1496. [Google Scholar] [CrossRef]
  70. NOM-001-SEMARNAT-2021, DOF. Available online: https://www.gob.mx/semarnat/prensa/se-publica-nom-001-semarnat-2021-que-establece-limites-de-contaminantes-en-descargas-de-aguas-residuales (accessed on 19 September 2022).
  71. Dubber, D.; Gray, N.F. Replacement of chemical oxygen demand (COD) with total organic carbon (TOC) for monitoring wastewater treatment performance to minimize disposal of toxic analytical waste. J. Environ. Sci. Health A 2010, 45, 1595–1600. [Google Scholar] [CrossRef]
  72. Netto, G.C.; Sauer, T.; José, H.J.; Peralta Muniz Moreira, R.F.; Humeres, E. Evaluation of relative photonic efficiency in heterogeneous photocatalytic reactors. J. Air Waste Manag. Assoc. 2012, 54, 77–82. [Google Scholar] [CrossRef]
  73. Cao, C.; Hu, C.; Wang, X.; Wang, S.; Tian, Y.; Zhang, H. UV sensor based on TiO2 nanorod arrays on FTO thin film. Sens. Actuators B Chem. 2011, 156, 114–119. [Google Scholar] [CrossRef]
  74. ElMekawy, A.; Hegab, H.M.; Alsafar, H.; Yousef, A.F.; Banat, F.; Hasan, S.W. Bacterial nanotechnology: The intersection impact of bacteriology and nanotechnology on the wastewater treatment sector. J. Environ. Chem. Eng. 2023, 11, 109212. [Google Scholar] [CrossRef]
  75. Yadav, H.M.; Otari, S.V.; Bohara, R.A.; Mali, S.S.; Pawar, S.H.; Delekar, S.D. Synthesis and visible light photocatalytic antibacterial activity of nickel-doped TiO2 nanoparticles against Gram-positive and Gram-negative bacteria. J. Photochem. Photobiol. A 2014, 294, 130–136. [Google Scholar] [CrossRef]
  76. Guillard, C.; Bui, T.H.; Felix, C.; Moules, V.; Lina, B.; Lejeune, P. Microbiological disinfection of water and air by photocatalysis. C. R. Chim. 2008, 11, 107–113. [Google Scholar] [CrossRef] [PubMed]
  77. Nadtochenko, V.A.; Rincon, A.G.; Stanca, S.E.; Kiwi, J. Dynamics of E. coli membrane cell peroxidation during TiO2 photocatalysis studied by ATR-FTIR spectroscopy and AFM microscopy. J. Photochem. Photobiol. A 2005, 169, 131–137. [Google Scholar] [CrossRef]
  78. Thakur, I.; Verma, A.; Örmeci, B. Visibly active Fe-TiO2 composite: A stable and efficient catalyst for the catalytic disinfection of water using a once-through reactor. J. Environ. Chem. Eng. 2021, 9, 106322. [Google Scholar] [CrossRef]
  79. Barajas-Ledesma, E.; García-Benjume, M.L.; Espitia-Cabrera, I.; Ortiz-Gutiérrez, M.; Espinoza-Beltrán, F.J.; Mostaghimi, J.; Contreras-García, M.E. Determination of the band gap of TiO2–Al2O3 films as a function of processing parameters. Mater. Sci. Eng. B 2010, 174, 71–73. [Google Scholar] [CrossRef]
  80. ISO 8245:1999; Water Quality—Guidelines for the Determination of Total Organic Carbon (TOC) and Dissolved Organic Carbon (DOC). International Organization for Standardization: Geneva, Switzerland, 1999.
  81. NOM-210-SSA1-2014, DOF. Available online: https://dof.gob.mx/nota_detalle.php?codigo=5398468&fecha=26/06/2015#gsc.tab=0 (accessed on 28 September 2022).
Figure 1. XRD patterns of the Al2O3-TiO2 coating annealed at 700 °C in air.
Figure 1. XRD patterns of the Al2O3-TiO2 coating annealed at 700 °C in air.
Catalysts 13 01351 g001
Figure 2. Field emission scanning electron microscopy images of Al2O3-TiO2 coating: (a) secondary electron image mode, (b) backscattered electron image mode and elemental mapping, and (c) schematic representation of Al2O3-TiO2 coating and the heterojunction evidenced with XRD results.
Figure 2. Field emission scanning electron microscopy images of Al2O3-TiO2 coating: (a) secondary electron image mode, (b) backscattered electron image mode and elemental mapping, and (c) schematic representation of Al2O3-TiO2 coating and the heterojunction evidenced with XRD results.
Catalysts 13 01351 g002
Figure 3. (a) Atomic force microscopy images of (b) Al2O3-TiO2 coating roughness (Rq) analysis of the Al2O3-TiO2 sample.
Figure 3. (a) Atomic force microscopy images of (b) Al2O3-TiO2 coating roughness (Rq) analysis of the Al2O3-TiO2 sample.
Catalysts 13 01351 g003
Figure 4. UV–visible spectrum of the Al2O3-TiO2 coatings: (a) bands of maximum absorption, (b,c) bandgap calculated with Tauc plot.
Figure 4. UV–visible spectrum of the Al2O3-TiO2 coatings: (a) bands of maximum absorption, (b,c) bandgap calculated with Tauc plot.
Catalysts 13 01351 g004
Figure 5. Photocatalytic activity of the Al2O3-TiO2 coating: (a) MB degradation and (b) total organic carbon removal.
Figure 5. Photocatalytic activity of the Al2O3-TiO2 coating: (a) MB degradation and (b) total organic carbon removal.
Catalysts 13 01351 g005
Figure 6. Combined effect of the Al2O3-TiO2 coating to inhibit bacterial growth.
Figure 6. Combined effect of the Al2O3-TiO2 coating to inhibit bacterial growth.
Catalysts 13 01351 g006
Figure 7. Airlift reactor, dimensions, arrangement of substrates, and placement of lamps; on the left upper corner, the M16 LED lamp spectrum is shown.
Figure 7. Airlift reactor, dimensions, arrangement of substrates, and placement of lamps; on the left upper corner, the M16 LED lamp spectrum is shown.
Catalysts 13 01351 g007
Table 1. Results of surface coating on 304 SS obtained with Rietveld method.
Table 1. Results of surface coating on 304 SS obtained with Rietveld method.
PhasesQuantity
(%)
Lattice Constants (nm)
abc
α-Al2O37.40.4748-1.294
γ-Al2O33.00.8017--
Anatase-TiO26.90.3801-0.9595
Rutile-TiO23.70.4620-0.2958
Fe Austenite720.35901--
Ferrite40.5581.4650.542
Magnetite10.8353--
Cr2O32-0.50091.3734
Table 2. Comparison of the degradation efficiency of different photocatalyst systems with TiO2.
Table 2. Comparison of the degradation efficiency of different photocatalyst systems with TiO2.
PhotocatalystConditions of Photocatalysis Degradation Efficiency
(%)
TypeDopingPresentationLoadingPollutantConcentrationLight SourceTime
(h)
Volume
(L)
Ref.
TiO2--Spherical, 70 nmFilms
(glass substrate)
---
MB1 × 10−6 M200 W Tungsten filament bulb. UV light (λ = 386 nm)40.01592.00[62]
CuNTiO21:1:1
atomic ratio
Spherical, ~20 nm.0.1 gMB3 × 10−5 M500 W Xenon lamp (upon blocking
UV light by a 400 nm glass filter)
60.591.00[63]
CuO/TiO21% wtnanorods0.05 gMB1.8 to 18
mg L−1
386-LED or 450-LED light emitting diode20.1092.00[60]
Ni_TiO20.5 Ni/TiSpherical, 50 nm0.24 gL−1MB3.2 mg L−1Osram
1000 W Xenon short arc display/optic lamp, XBO,
40.2099.00[64]
Ni_TiO20.75 Ni/TiNanosheets
90–100 nm, thickness: 15–20 nm
0.03 gMB3 mg L−1500 W power Xenon lamp (LSXS500) with a visible cut-off filter (l < 400 nm)30.0396.00[65]
N_TiO21.6 N/TiSpherical, 10 nm0.25 gPhe25 mg L−1luminous intensity
15.54 mWcm−2 (400 < λ < 650 nm)
100.1081.00[66]
TiO2-Fe2O31:1porous
diameter
< 1 µm
Coating
(glass substrates)
MB3 mg L−1150 W, halogen lamp100.00367.00[48]
TiO2-Fe2O35 Fe/TiGranular, 9–17 nm0.4 gMB1 × 10−5 M100 W Halogen Lamp, 100 PAR30/FL 240 V (350 < λ < 1050 nm) INTENSITY: 85 Mw cm−220.0287.00[67]
Al2O3-TiO20.1 Al/TiGranular,
10–20 nm
0.025 gRh(B)2.0 × 10−5 M500 W halogen lamp, cut-off filter (λ < 450 nm)0.160.0589.25[68]
Al2O3-TiO21.5 Al/TiTetrahedral, and
rice-like, 210 ± 90 nm
Coating
(304SS substrates)
MB6 mg L−14 LED lamps (4.5 W, 400 Lumens,
MR16 Philips)
410 < λ < 760 nm
60.9897.43*
MB: methylene blue dye, Rh(B): rhodamine B dye, Phe: phenol, and * this study.
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.; 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. https://doi.org/10.3390/catal13101351

AMA Style

Camacho-González MA, Lijanova IV, Reyes-Miranda J, Sarmiento-Bustos E, Quezada-Cruz M, Vera-Serna P, Barrón-Meza MÁ, Garrido-Hernández A. High Photocatalytic Efficiency of Al2O3-TiO2 Coatings on 304 Stainless Steel for Methylene Blue and Wastewater Degradation. Catalysts. 2023; 13(10):1351. https://doi.org/10.3390/catal13101351

Chicago/Turabian Style

Camacho-González, Mónica Araceli, Irina Victorovna Lijanova, Joan Reyes-Miranda, Estela Sarmiento-Bustos, Maribel Quezada-Cruz, Pedro Vera-Serna, Miguel Ángel Barrón-Meza, and Aristeo Garrido-Hernández. 2023. "High Photocatalytic Efficiency of Al2O3-TiO2 Coatings on 304 Stainless Steel for Methylene Blue and Wastewater Degradation" Catalysts 13, no. 10: 1351. https://doi.org/10.3390/catal13101351

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop