1. Introduction
In recent decades, the proliferation of industrial activities has led to an alarming increase in environmental pollutants, particularly those stemming from factory operations involving dyes [
1]. While essential in various industries for coloring fabrics, plastics, and other materials, these dyes often contain complex chemical compounds that are discharged as waste [
2]. When improperly managed, these dye-related pollutants can cause significant environmental harm [
3], contaminating water sources, soil, and air and adversely affecting ecosystems and human health [
4].Water pollution has become a pressing issue in developing countries [
5] due to the prevalence of textile industries dominating their economic landscapes, among other sectors [
6]. These industries and agricultural practices employ harmful chemicals that find their way into water bodies [
7], contributing significantly to global water contamination [
8]. The discharge of dyes and toxic organic compounds from industrial and agricultural processes has a profound impact [
9], polluting aquatic environments on a large scale and posing risks to aquatic life and human health through the food chain [
10]. Areas where irrigation with contaminated water occurs face additional challenges [
11], leading to soil pollution and associated health risks [
12]. The persistence of these toxic compounds complicates their biochemical degradation [
13], prompting research efforts to develop alternative and efficient methods for their removal from air and water [
14]. Techniques such as microfiltration [
15], electro-coagulation [
16], and flocculation have been proposed for water treatment but often need to more completely degrade pollutants quickly and affordably [
17].
The challenge of mitigating such environmental damage has spurred the development of innovative technologies to reduce and manage these pollutants [
18]. One promising approach is photocatalysis, which leverages light to accelerate chemical reactions that break down harmful substances [
19]. In photocatalysis, a photocatalyst, typically semiconductor material, absorbs photons and generates electron–hole pairs that drive the degradation of pollutants [
20]. This process holds significant potential for the filtration and detoxification of polluted water and air, making it a key area of research for environmental remediation [
21].
Photocatalysis emerges as a promising solution for addressing water pollution challenges [
22]. This technology leverages light as a catalyst [
23], integrating nanotechnology and materials science to combat environmental pollutants [
24]. Photocatalytic oxidation, a key process in this field, utilizes semiconductor surfaces activated by photon energy to generate radicals that decompose organic pollutants into harmless CO
2 and H
2O [
25].Ceramics have emerged as a valuable component in photocatalysis due to their stability, durability, and catalytic properties [
26]. Among various ceramic materials, kaolin, a type of clay mineral, has garnered attention for its effectiveness as a catalyst [
27]. Kaolin’s advantages include its high surface area, thermal stability, and cost-effectiveness [
28]. Its applications extend beyond photocatalysis, encompassing fields such as ceramics manufacturing and environmental engineering [
29]. Kaolin’s intrinsic properties make it a suitable candidate for developing photocatalytic systems aimed at pollutant degradation [
30].
The choice of raw materials is crucial for developing effective photocatalytic systems [
31], with ceramic membranes presenting a viable option for their stability and cost-effectiveness [
32]. Clay minerals such as kaolinite have been widely studied for their potential in fabricating heterogeneous photocatalysts [
33], often combined with transition metal oxides to enhance catalytic activity [
34]. In this review, kaolin-based materials (DD3 and DD3Z, the latter containing 38% ZrO
2) from Guelma, Algeria, are examined for their suitability in photodegrading organic dyes [
35], particularly using the dye Orange II as a test case [
35]. DD3Z, enriched with zirconium oxide, demonstrates increased open porosity favorable for catalytic applications after suitable heat treatment [
36].To enhance the photocatalytic performance of ceramics, researchers have explored the incorporation of various oxides [
37], including zinc oxide (ZnO) [
38], copper oxide (CuO) [
39], and magnesium oxide (MgO) [
40]. These added oxides offer unique characteristics that improve the efficiency of the photocatalysis process. Zinc oxide, for instance, is known for its strong UV absorption and high electron–hole pair generation [
41]. Copper oxide contributes to effective pollutant degradation through its ability to create reactive oxygen species [
42]. Magnesium oxide enhances photocatalytic activity by improving charge separation and stability. Together, these oxides can significantly boost the performance of photocatalytic systems [
43].
Recent studies have explored bimetallic and mixed metal oxide composites, demonstrating superior efficiency compared to non-metallic counterparts [
44]. Zinc oxide (ZnO) [
45] and copper oxide (CuO) [
46] are prominent examples of transition metal oxides used in photocatalytic applications [
47]. Each exhibits specific band gaps that allow for the effective degradation of organic compounds under UV or visible light irradiation [
48].The preparation of photocatalytic materials involves several methods, each influencing the properties and effectiveness of the final product [
48]. Standard preparation techniques include sol–gel processes/hydrothermal synthesis and co-precipitation/mixing powder methods [
49]. Specifically, for depositing thin layers on ceramic substrates, dipping and autoclave methods are frequently employed [
50]. Dip-coating involves immersing the ceramic substrate into a precursor solution [
35], while autoclave processing uses high pressure and temperature to enhance the deposition quality [
51]; for the incorporation of oxides, deposition and mixing methods are utilized to ensure a uniform distribution and optimal interaction between the ceramic base and the added oxides [
52]. Various composite ceramic/metal oxide powders and thin films were prepared using co-precipitation and traditional mixing methods [
52,
53] alongside Cu-doped ZnO thin layers deposited via sol–gel techniques on ceramic substrates [
33,
51], all evaluated for their efficacy in dye degradation through photocatalysis [
51,
52,
53].
This work presents a systematic review and comparative analysis of our six-year research program (2017–2023) on Algerian kaolinite (DD3)-based photocatalysts. By evaluating four preparation methods (traditional mixing, co-precipitation, sol–gel, and autoclave) and three metal oxides (ZnO, CuO, MgO) under controlled conditions, we identify optimal strategies for photocatalytic applications. Unlike conventional reviews, this analysis leverages internally consistent datasets from our previous studies [
33,
51,
52,
53], all using identical Guelma kaolinite substrates and Orange II dye degradation tests, enabling direct method–performance comparisons.
In this study, we have investigated the application of photocatalysis using ceramic materials, focusing on the role of kaolin as a base catalyst and the impact of added oxides like zinc, copper, and magnesium oxide [
54]. We utilized various preparation methods, including dip-coating and autoclave techniques for applying thin layers on ceramic substrates [
33,
51] and deposition and mixing methods for integrating the oxides [
52,
53]. The objective was to evaluate the effectiveness of these materials and techniques in enhancing photocatalytic performance for environmental pollutant degradation (
Figure 1), thereby contributing to the advancement of sustainable technologies for environmental remediation.
4. Photocatalysis of Dyes
The photocatalytic degradation of organic dyes begins with light absorption by a semiconductor photocatalyst [
94]. When the energy of the absorbed photons equals or exceeds the bandgap energy of the semiconductor, electron–hole pairs are generated within the material. This initial step is crucial for initiating the photocatalytic process [
94]. Once formed, these electron–hole pairs can follow two primary pathways. They may recombine, releasing energy as heat, which is an undesirable outcome as it reduces the efficiency of the photocatalytic process. Alternatively, and more importantly for dye degradation, these charge carriers can migrate to the surface of the semiconductor, where they interact with adsorbed species [
95].
At the semiconductor surface, electrons typically transfer to electron acceptors like oxygen molecules, forming superoxide radicals (•O
2−). Simultaneously, holes can interact with electron donors such as water or hydroxide ions, producing highly reactive hydroxyl radicals (•OH) [
96]. These radicals, particularly hydroxyl radicals, are powerful oxidizing agents capable of breaking down complex organic molecules like dyes [
95,
96]. The organic dye molecules, often adsorbed onto the photocatalyst surface, are then attacked by these reactive species. This leads to a series of oxidation reactions that progressively break down the dye’s molecular structure [
97]. The process typically involves the cleavage of chromophore groups responsible for the dye’s color, followed by further decomposition of the resulting fragments [
97].
Ideally, the end products of this degradation process are simple, harmless substances like carbon dioxide and water, representing the complete mineralization of the dye [
96,
97]. However, the efficiency and completeness of this process can vary depending on factors such as the nature of the photocatalyst, the specific dye being degraded, and the reaction conditions [
98]. Photocatalysis represents a promising application of ceramics due to their nature abundance, cost-effectiveness, and environmental friendliness [
99]. This technology is particularly significant because it can degrade, reduce, and mineralize hazardous organic compounds into harmless CO
2 and H
2O [
100]. It finds extensive use in reducing toxic metal ions, disrupting waterborne microorganisms, and decomposing air pollutants into volatile organic compounds [
101].
Photocatalysis operates based on advanced oxidation or reduction processes initiated by the excitation of electrons following photon absorption [
102]. In this mechanism, semiconductors serve as active catalysts by facilitating the generation of electron–hole pairs under light irradiation [
103]. While ZnO and CuO have been widely studied, several other metal oxides have also demonstrated strong photocatalytic activity under different lighting conditions. TiO
2, one of the earliest and most extensively explored photocatalysts, functions mainly under UV light by promoting electron excitation from the valence band to the conduction band, leading to the production of superoxide (•O
2−) and hydroxyl radicals (•OH) capable of oxidizing organic molecules [
103]. MgO, though traditionally considered less active under visible light, can enhance photocatalytic efficiency through the introduction of surface defects, which act as active sites for radical generation and improve charge carrier separation [
87]. Fe
2O
3 (hematite), with a narrower bandgap (~2.1 eV), is particularly attractive for visible-light-driven photocatalysis; it facilitates surface oxidation processes but often requires strategies to mitigate fast electron–hole recombination [
81]. WO
3 (tungsten trioxide) offers strong absorption in the visible range and promotes photocatalytic reactions primarily through the activity of valence band holes that generate hydroxyl radicals, contributing significantly to pollutant degradation [
85]. Together, these examples underline the critical roles played by material properties such as bandgap energy, crystallinity, surface morphology, and defect structures in determining photocatalytic performance. A comprehensive understanding of these factors is essential for the rational design of efficient photocatalysts for environmental applications. Overall, the photocatalytic mechanism closely resembles heterogeneous catalysis, where oxidation and reduction reactions are triggered at the semiconductor surface [
52,
103].
Photodynamics in photocatalysis hinges on the creation of electron–hole pairs within semiconductors upon absorbing photons with energy equal to or exceeding the bandgap energy (hν ≤ Eg) (Equation (1)) [
104]. Subsequent to photon absorption and the formation of electron–hole pairs (e−/h+) within the solid mass, these pairs undergo either recombination through heat release or interaction with species absorbed on the semiconductor surface [
104]. Electrons migrate to acceptor molecules (A) in the fluid phase (gas or liquid) based on their redox potential (Equation (2)), while holes transfer to donor molecules (D) (Equation (3)) [
105].
This process generates highly effective free radicals, facilitating oxygen regeneration and water oxidation (Equations (4) and (5)) [
52,
53,
105]. The production of hydroxyl radicals (•OH) is particularly efficient during photocatalysis, effectively reducing the concentration of various chemical compounds and decomposing them with a high oxidation capacity (2.8 eV), surpassing that of other oxidants such as O
2 (2.42 eV) and H
2O
2 (1.78 eV) [
106]. Additional oxidants like HO and H
2O
2 may also arise, potentially leading to intermediate product formation and the mineralization of chemical compounds absorbed on the photocatalyst surface [
107]. The following chemical equations can summarize these mechanisms.
The comparison in
Table 8 highlights the evolution of photocatalytic research, showing a trend towards more diverse and efficient photocatalysts, a deeper understanding of reaction mechanisms, and a more holistic approach to assessing the environmental impact and practical applicability of photocatalytic dye degradation processes [
108].
4.1. Photocatalytic Performance
4.1.1. The Catalytic Performance of Powders
Figure 16 illustrates the effectiveness of powders prepared via the mixing method using two ceramic types (DD3 and DD3 + 38% ZrO
2), both with and without additions of ZnO, CuO, and MgO, in purifying and analyzing OII-polluted solutions through absorbance spectra under ultraviolet light [
52,
53]. Initially, the powders without metal-oxide additions achieved a degradation efficiency of 42% for DD3 (
Figure 16a) and 60.3% for DD3 + 38% ZrO
2 (
Figure 16b) [
52] over 7 h and 2 h, respectively [
53]. After incorporating zinc oxide (ZnO) and copper oxide (CuO) via the traditional mixing method to enhance organic dye degradation [
52], DD3-based powders achieved a degradation rate of 93.6% within 30 min, whereas DD3 + 38% ZrO
2 reached the same percentage in just 15 min under visible light exposure [
52]. Additionally, the addition of MgO to ceramic powders demonstrated high efficiency in degrading OII-contaminated solutions, with degradation rates of 77.3% for DD3Z and 74.13% for DD3 after only 5 min [
53].
The pH evolution during photocatalytic degradation was monitored using a calibrated pH meter. The initial Orange II solution in distilled water showed a pH of 6.8 ± 0.2. Under illumination, the pH decreased to 5.2 ± 0.3 after 60 min due to the formation of acidic intermediates (e.g., carboxylic acids) and proton release from the catalyst surface. In dark adsorption tests, the pH remained stable (6.5–6.7), confirming illumination-dependent acidification. This pH shift may influence the catalyst surface charge (PZC ~7.5 for ZnO) and dye molecule ionization. Post-reaction XRD confirmed catalyst stability, showing no phase changes and minimal metal leaching (<0.5 ppm Zn/Cu) after five cycles.
Powders prepared using the co-precipitation method, following the addition of 28 wt.% ZnO/2.8 wt.% CuO [
52] to both ceramic types, exhibited notable degradation results in UV–visible spectra. Specifically, DD3-based powders achieved a degradation rate of 84.1% after 150 min, while DD3 + 38% ZrO
2 powders reached 99.6% degradation within 45 min [
52].
Figure 17 illustrates the photocatalytic degradation of Orange II dye using different additives (ZnO, CuO, and MgO) treated at 500 °C [
52,
53]. The graph shows the degradation efficiency over time for various compositions. The pure DD3 and DD3Z samples exhibit relatively low photocatalytic activity. However, the addition of metal oxides significantly enhances the degradation rate. Notably, the samples containing ZnO show the highest degradation efficiency, with DD3Z + 10 wt.% ZnO performing best, achieving nearly complete degradation within 120 min [
52]. The CuO-containing samples also show improved performance compared to the base materials, while MgO additions appear to have a lesser impact on the photocatalytic activity [
53]. This comparative study demonstrates the effectiveness of ZnO as a photocatalytic enhancer for these ceramic composites in the degradation of Orange II dye [
116]. Also
Table 9 Photocatalysis in previous studies using samples prepared with various methods [
117,
118,
119,
120,
121,
122].
4.1.2. The Catalytic Performance of DD3-Based Layer Samples
The impact of ceramic substrates (DD3, DD3 + ZrO
2) with and without active Cu-doped ZnO thin films, fabricated using both autoclave and sol–gel methods [
33,
51], was investigated for their practical applicability in liquid purification [
33,
51]. Distilled water intentionally contaminated with OII was used as the test solution. The study utilized infrared spectroscopy, as depicted in
Figure 17 [
33,
51], to analyze the evolution of absorbance spectra following exposure to ultraviolet (UV) radiation. For DD3-based substrates, the absorption spectra showed minimal change, with a reduction rate not exceeding 10% regardless of UV exposure time, even with active layers [
33]. Similar behavior was observed for DD3 + ZrO
2 porous substrates before the deposition of thin layers [
33]. However, upon deposition of ZnO thin films, the degradation rate increased to 77.7% for the same substrate type after 6 h of UV exposure [
33].
Using the autoclave method to deposit the ZnO layer on DD3 + ZrO
2 substrates (
Figure 18) [
51], the absorbance spectra demonstrated significant efficacy, achieving an 81.1% degradation rate over 6 h. Conversely, the photocatalytic activity of layers deposited on DD3 substrates with the same doping and exposure time yielded a lower degradation rate of 36.1%. This discrepancy is attributed to the enhanced open porosity of the DD3 + ZrO
2 substrate, which facilitates the deposition of ZnO and CuO particles during sedimentation [
51].
Overall, both types of ceramics exhibited enhanced photocatalytic performance when the autoclave method was employed compared to the sol–gel method, highlighting the role of substrate porosity and effective deposition of active layers in enhancing photocatalytic efficiency [
123].
Figure 19 compares the degradation of Orange II dye using different preparation methods for the photocatalysts. The graph shows degradation efficiency over time for samples prepared via traditional mixing, co-precipitation, and autoclave methods [
33,
51,
52,
53]. The autoclave method consistently demonstrates superior performance across all compositions, achieving the highest degradation rates. The co-precipitation method generally shows intermediate performance [
52], while the traditional mixing method exhibits the lowest degradation efficiencies [
52,
53]. Notably, the DD3Z + 10% ZnO sample prepared by the autoclave method achieves nearly complete degradation within 60 min, significantly outperforming the same composition prepared by other methods [
52]. This comparison highlights the crucial role of the preparation method in determining the photocatalytic activity of these ceramic composites, with the autoclave method proving the most effective in enhancing the degradation of Orange II dye [
124].
The photocatalytic activities of various photocatalysts were compared based on their effectiveness in degrading methylene blue (MB) and other pollutants under different light sources. Shokraiyan et al. [
125] reported that ZnO/CuO nanocomposites exhibited high photocatalytic performance with 98% degradation of MB and 97% of 4-nitrophenol under LED light irradiation, as measured by decreases in their characteristic absorption peaks at 664 nm and 400 nm, respectively.
Büşra Çinar et al. [
126] studied CuO–TiO
2 p-n heterostructures, finding that the optimal photocatalytic efficiency under UV light was achieved with 1.25 wt.% CuO, leading to a 99% degradation of MB in 45 min. In contrast, visible light irradiation showed lower activity with a maximum of 98% degradation at 0.5 wt.% CuO but only 59% at 1.0 wt.%. Meanwhile, Karina Bano et al. [
92] highlighted that the CuO/ZnO heterojunctions were highly effective in removing tetracycline and ciprofloxacin, achieving 94% and 93% removal, respectively, within 50 min, demonstrating superior photocatalytic activity compared to pure CuO and ZnO [
92]. These results underline the variability in photocatalytic efficiency depending on the material composition and the type of light used.
Table 10 and
Table 11 compares photocatalytic degradation of DD3 using powder and thin-film samples prepared by different methods. Powder samples generally show higher degradation rates than thin films. Adding 38% ZrO
2 to DD3 improves performance across methods. For powders, mixing with metal oxides (ZnO-CuO or MgO) significantly enhances degradation, with ZnO-CuO mixtures showing the highest rates (89.52% in 30 min, reaching 100% in 45 min). The co-precipitation method with ZrO
2 addition yields the best results for powders (99.6% in 45 min), achieving nearly complete degradation quickly. In thin films, sol–gel preparation with ZrO
2 and ZnO:Cu layers performs best, though less effectively than powders. Overall, combining DD3 with ZrO
2 and metal oxides like ZnO and CuO substantially improves photocatalytic activity, with preparation method greatly influencing performance.
4.2. Mechanism of DD3Z/ZnO/CuO and DD3Z/MgO
The increased reliance on fossil fuels like petroleum and coal has led to fuel shortages and significant environmental issues, including global warming and air pollution. Additionally, the global production of organic pollutants such as synthetic dyes, pesticides, and fertilizers continues to harm the environment and living organisms [
120]. In response, eco-friendly energy alternatives, including solar cells, hydropower, and clean hydrogen energy, are being explored to reduce environmental impact [
127]. Among these, clean hydrogen energy stands out for its efficiency in producing clean energy and environmental protection. Various techniques, such as electrolytic processes, photochemical methods, and solar water splitting, are being developed for hydrogen production [
128]. Photocatalysis, in particular, has emerged as a promising, cost-effective method for environmental applications, including water purification and pollutant degradation [
129].
The determination of semiconductor positions is crucial and is referenced against the standard hydrogen electrode (SHE), coded as NHE (
Figure 20) [
33,
35,
51,
52,
53]. The conduction and valence band positions for ZrO
2, ZnO, CuO, and MgO at the zero-charge point were calculated using the following relationships [
52,
53].
where E
VB represents the edge potential of the valence band, E
CB is the conduction band edge potential, X is the electronegativity value for semiconductors (X
CuO = 5.81 eV, X
ZnO = 5.79 eV, X
ZrO2 = 5.92 eV, X
MgO = 5.2 eV, X
NiO = 5.75 eV, X
Ag2O = 5.2 eV, X
MnO2 = 5.315 eV) [
33,
51,
52,
53], Ee is the energy of the free electrons on the hydrogen scale ~4.5 eV, and Eg is the semiconductor gap energy. After applying the two relations on each element, the conduction and valence band value was found; the results are shown in
Table 12 [
52,
53,
130,
131].
According to S. Nayak et al. [
130], the oxidation potentials of the holes in the valence band need to be sufficiently positive to generate hydroxyl radicals. It is noted that the standard oxidation potential (E
Θ (OH
−/•OH)) required for the conversion of OH
− to •OH is +1.99 eV [
44]. Conversely, the electrons in the conduction band must be negative enough to generate superoxide radicals [
131], with the standard oxidation potential (EΘ (O
2/•O
2−)) needed for the conversion of O
2 to •O
2− being −0.33 eV [
130]. When selecting oxides from various semiconductors, it is important to consider those with valence and conduction band levels most suitable for the mineralization of organic pollutants.
Figure 21 summarizes these findings.
The photocatalytic process relies on light-induced redox reactions to generate active radicals like hydroxyl (•OH) and superoxide (•O
2−) for breaking down pollutants [
131]. Effective photocatalysis depends on selecting materials that efficiently absorb sunlight and promote charge separation and transfer [
132]. Metal-oxide nanomaterials like TiO
2, ZnO, WO
3, and Fe
2O
3 have been studied for their photocatalytic properties due to their diverse structures and optical characteristics [
133]. Among these, TiO
2 and ZnO are particularly noted for their stability, photosensitivity, and non-toxic nature, making them effective for environmental cleanup [
134].
ZnO is known for its suitable band gap (approximately 3.2–3.4 eV), allowing it to utilize visible light effectively [
135]. However, issues such as rapid charge recombination and photo corrosion limit its efficiency. Strategies to enhance ZnO’s photocatalytic performance include doping, forming nanocomposites or heterojunctions with other semiconductors, and modifying its nanostructure to improve light absorption and charge separation [
135]. Despite its potential, the photocorrosion problem can be mitigated by coating ZnO with protective layers like graphene or carbon nanotubes, which enhance its stability and performance [
136].
Overall, developing advanced photocatalytic materials and methods continues to be a key area of research to address environmental challenges and improve the efficiency of clean energy technologies [
134,
135,
136].Based on the provided information, we can compare the photocatalytic performance of various ZnO- and CuO-based nanocomposites and heterojunctions [
137]. The Ag/Ag
2O/ZnO ternary nanocomposite prepared by arching technique showed excellent performance in decomposing methylene blue (MB) under UV-light irradiation, outperforming single- and double-constituent catalysts [
138]. The γ-Fe
2O
3-ZnO–biochar nanocomposite, synthesized via thermal decomposition, demonstrated the highest reported efficiency of 86.2% in degrading Rhodamine B (RhB) within 30 min [
139]. Another notable performer (
Table 13) was the Ag@ZnO/TiO
2 flexible hierarchical heterojunction, produced by hydrothermal and photodeposition processes, which achieved 91.6% degradation of tetracycline hydrochloride (TC-H) in 1 h [
140]. The CuO-TiO
2 heterostructure also showed promising results in MB degradation under visible light, with hydroxyl radicals playing a key role in the photocatalytic process [
140].
Overall, the photocatalytic composites demonstrated strong potential for dye degradation, although further investigations into mineralization pathways and by-product identification will be necessary to fully validate their environmental impact [
120,
121]. While the photocatalytic activity results, based on UV-Vis spectrophotometry, clearly demonstrated a rapid and significant decrease in the characteristic absorption peak of Orange II at 484 nm, it is essential to note that complete mineralization of the dye into CO
2 and H
2O cannot be conclusively confirmed using this technique alone [
122]. The UV-Vis method primarily monitors the disappearance of chromophoric groups but does not provide detailed information about intermediate products formed during degradation [
123,
124].
According to previous studies, Orange II degradation often proceeds through forming intermediate species such as aromatic amines, carboxylic acids, and nitrogen-containing compounds [
125,
126]. Without employing advanced analytical techniques like High-Performance Liquid Chromatography (HPLC), Gas Chromatography–Mass Spectrometry (GC-MS), Total Organic Carbon (TOC) analysis, or Ion Chromatography (IC), the identification and quantification of these intermediates remain unresolved [
127,
128,
129].
Future work will incorporate such techniques to provide a more detailed understanding of the degradation pathway and the extent of mineralization achieved [
130,
131]. Combining photocatalysis with complementary advanced oxidation processes such as photo-ozonation is also a promising strategy [
132,
133]. Ozone could enhance hydroxyl radical production, accelerate oxidation kinetics, and promote deeper mineralization of persistent organic molecules, addressing the limitations of photocatalysis alone [
134,
135,
136].
Some recent studies have successfully combined photocatalysis with ozonation, resulting in higher mineralization rates and shorter treatment times than photocatalysis alone, especially for complex dyes like Orange II [
137,
138,
139,
140,
141].
5. Conclusions
This comprehensive review of photocatalytic material preparation methods yielded several key findings. Adding ZrO2 to DD3 ceramics significantly altered grain structure and increased porosity, enhancing the material’s potential for photocatalytic applications. They have been highlighted due to their natural abundance, affordability, ease of access, and favorable thermal and mechanical properties. Co-precipitation methods generally produce finer particles than traditional mixing, potentially leading to better dispersion and enhanced properties.
Sol–gel and autoclave methods for thin-film deposition resulted in distinct morphologies, with sol–gel producing more uniform flower-like shapes and finer-grained surfaces and autoclave methods yielding larger, more distinct crystallites. XRD analysis revealed the successful incorporation of metal oxides into the ceramic matrix, with observed peak shifts indicating structural distortions due to doping. The analysis confirmed a shift in spectra towards higher values, indicating smaller particle sizes with zinc and copper additions and larger particle sizes with magnesium oxide additions. SEM imaging demonstrated that the addition of ZnO, CuO, and MgO increased porosity and altered surface morphology, creating structures suitable for capturing pollutants. EDX analysis confirmed the successful incorporation of dopant elements, with varying concentrations depending on the preparation method and substrate used.
The photocatalytic testing setup for Orange II dye degradation was described, setting the stage for future performance evaluations. During powder preparation, after adding zinc and copper oxides, the decomposition rate reached 93.63% within a short period of 15 min. When magnesium oxide was added to the ceramic material, the photocatalytic efficiency surpassed that of the previous materials, achieving complete pigment dissolution in under 10 min. Upon depositing Cu thin layers on both types of ceramic samples, the absorption capacity of samples prepared via the autoclave method exceeded that of samples prepared via the sol–gel method.
These findings contribute to understanding how different preparation methods and dopants affect the properties of ceramic-based photocatalysts. This study provides a foundation for optimizing these materials for environmental applications, particularly in the degradation of organic pollutants in water. Future research should focus on quantifying the photocatalytic performance of these materials and exploring their potential for large-scale water treatment applications.