Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 3720 2022-08-15 13:45:56 |
2 format Meta information modification 3720 2022-08-16 03:25:40 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Li, X.;  Simon, U.;  Bekheet, M.F.;  Gurlo, A. Mineral-Supported Photocatalysts. Encyclopedia. Available online: https://encyclopedia.pub/entry/26149 (accessed on 26 April 2024).
Li X,  Simon U,  Bekheet MF,  Gurlo A. Mineral-Supported Photocatalysts. Encyclopedia. Available at: https://encyclopedia.pub/entry/26149. Accessed April 26, 2024.
Li, Xue, Ulla Simon, Maged F. Bekheet, Aleksander Gurlo. "Mineral-Supported Photocatalysts" Encyclopedia, https://encyclopedia.pub/entry/26149 (accessed April 26, 2024).
Li, X.,  Simon, U.,  Bekheet, M.F., & Gurlo, A. (2022, August 15). Mineral-Supported Photocatalysts. In Encyclopedia. https://encyclopedia.pub/entry/26149
Li, Xue, et al. "Mineral-Supported Photocatalysts." Encyclopedia. Web. 15 August, 2022.
Mineral-Supported Photocatalysts
Edit

Although they are of significant importance for environmental applications, the industrialization of photocatalytic techniques still faces many difficulties, and the most urgent concern is cost control. Natural minerals possess abundant chemical inertia and cost-efficiency, which is suitable for hybridizing with various effective photocatalysts. The use of natural minerals in photocatalytic systems can not only significantly decrease the pure photocatalyst dosage but can also produce a favorable synergistic effect between photocatalyst and mineral substrate. Owing to their unique structures, large surface area, and negatively charged surface, silicate minerals could enhance the adsorption capacity, reduce particle aggregation, and promote photogenerated electron-hole pair separation for hybrid photocatalysts. Moreover, controlling the morphology and structure properties of these materials could have a great influence on their light-harvesting ability and photocatalytic activity. Composed of silica and alumina or magnesia, some silicate minerals possess unique orderly organized porous or layered structures, which are proper templates to modify the photocatalyst framework. The non-silicate minerals (referred to carbonate and carbon-based minerals, sulfate, and sulfide minerals and other special minerals) can function not only as catalyst supports but also as photocatalysts after special modification due to their unique chemical formula and impurities. The dye-sensitized minerals, as another natural mineral application in photocatalysis, are proved to be superior photocatalysts for hydrogen evolution and wastewater treatment. 

photocatalysis photocatalysts natural mineral photocatalytic supports synergistic effect wastewater treatment dye photodegradation composite photocatalysts

1. Introduction

As environmental pollution issues have taken place more and more frequently in recent years, sustainable development has become the leading target of energy and environmental policy–making worldwide, determining the public concern for scientific research. Since the TiO2-based water-splitting phenomenon was discovered by Fujishima and Honda in 1972, photocatalysis has stepped onto the solar energy conversion stage [1]. Many researchers have verified the extensive potential of TiO2 in the photodegrading of different pollutants, such as bisphenol A [2], tetracycline [3], gaseous formaldehyde [4], S. aureus [5], etc. As a result, photocatalysis is widely regarded as the most promising route to solve both the energy and pollution crises. Several common photocatalysts have been found to possess excellent photocatalytic performance for the degradation of toxic dye and organic pollutants, including ZnO [6][7], In2O3 [8], g-C3N4 [9], BiOBr [10], etc.
However, large-scale industrial applications of pure photocatalysts often suffer several key issues: (1) small specific surface area and resultant poor adsorptive ability; (2) easy photo-generated electron-hole recombination and resultant low quantum efficiency; (3) difficult solid-liquid separation and photocatalyst recycling; and (4) high cost. For example, CdS has superior visible-light absorption and photocatalytic activity [11], while it suffers severe photo-corrosion issues and easily causes secondary pollution problems. Although TiO2 has stable and superb photocatalytic performance, it is only UV-light responsive and hard to recycle [12]. Additionally, self-aggregation and high photogenerated electron-hole recombination rates are the other common issues that the purest photocatalysts encounter. Researchers have been trying many different modification strategies to address these problems. So far, several modification strategies have been developed to control the photocatalyst morphology and suppress photo-induced charge carrier recombination, including heterojunction design, noble-metal deposition, support composites, etc. [13][14][15]. Among these modification strategies, the support composite is the most promising method to overcome the photocatalyst recycling difficulty and thus realize the industrialization of photocatalysts. Common photocatalyst support materials are carbon nanotubes, mesoporous carbon, graphene graphite, zeolite, etc. While these carbon materials generally exhibit more regular morphology and better electrical properties, they are not the ideal support candidates for photocatalyst industrial development considering their cost and recycling issues. Natural minerals are generally defined as abundant natural resources on earth that are obtained directly from nature. Compared with other support materials, natural minerals are well known as suitable photocatalyst support candidates for the following reasons: (1) They are environmentally friendly, low cost, chemical resistant, mechanically stable, easily available, and do not cause secondary pollution to the environment; (2) natural minerals have the characteristics of a large specific surface area, providing more active centers and adsorbing pollutants to concentrate catalysis; (3) natural minerals are also excellent photocatalysts, such as ilmenite and pyrite, which play a good role in the degradation of water pollutants. More importantly, the combination of photocatalysts and natural minerals has become one of the most valuable modification strategies due to the vast availability and cost-effectiveness of natural minerals.

2. Photocatalyst Modification Methods

Essentially, photocatalysis is a particular catalysis type involving light absorption by several specific active species [16]. For example, a novel inorganic–organic TiO2-g-C3N4 composite could be excited under visible light (λ ≥ 400 nm) irradiation, activating the electron (e) transfer from the valence band (VB) to the conduction band (CB) [17]. In this case, the photo-induced holes of TiO2 could be attracted to the facet-coupled interface between TiO2 and g-C3N4 due to the internal static electric fields of TiO2, and then transferred to VB of g-C3N4, oxidizing RhB directly. Meanwhile, the photo-induced electrons of g-C3N4 transfer to TiO2 due to the higher CB potential, reacting with the adsorbed O2 into superoxide radicals (•O2−1).

2.1. Ion Doping

Doping is one of the most frequently applied methods to effectively extend the light absorption range of semiconductors via directly altering the bandgap structure [18]. Generally, TiO2 photocatalysts could be doped with metal and non-metal dopants to enhance their photocatalytic activity [19][20][21]. For example, cobalt-doped TiO2 samples synthesized via one-step hydrothermal method varied the cobalt dopant concentration from 0.1% to 3.0% to expand its optical response to the visible light region [22]. The photocatalytic phenol degradation reaction reached the highest degradation rate (81.72%) when the cobalt-doped amount was fixed at 0.3%. First principle calculation results indicate the density of the impurity state strengthens with increased dopant concentration, which is beneficial for the visible light absorption range extension and the enhancement of its photocatalytic efficiency. Additionally, the doped ions can also act as shallow traps to promote photoinduced electron-hole separation. However, when the doping concentration exceeds by too much, these doped ions are contrarily exchanged into photo-excited electron-hole recombination centers. Another work found that nitrogen ions could substitute the oxygen lattice of the TiO2 matrix, therefore narrowing the band gap and shifting the absorption edge to the visible light region [23]. The BET surface areas of N–TiO2 and N–TiO2/C nanocomposite were calculated to be 101.4 and 131.6 m2/g, higher than that of TiO2 (87.7 m2/g).
In order to process acidic gas pollution such as CO2 and NOx, special ions, namely alkali or alkaline earth metal ions, are incorporated into photocatalysts for better acidic gas adsorption capacity. For example, alkali (Na, K, Rb, and Cs) or alkaline earth metals (Mg, Ca, Sr, and Ba) were selected to modify anatase TiO2 photocatalyst and investigate their effect on low-temperature NOx trapping efficiency [24]. It was proved that the NOx storage capacity is positively related to the amount of the modified metal species, and Cs-TiO2 and Ba-TiO2 have similar capacities (179 and 187 μmol∙g−1, respectively), which is higher than that of conventional NOx-trapping catalysts. According to in situ diffuse reflectance infrared Fourier transfer (DRIFT) spectra results, adsorbed NOx species were trapped on the Ti sites of alkaline earth–doped TiO2 and the alkali sites of alkali-doped TiO2, while NOx trapped on Ti sites was found to be more efficient than on other sites for its excellent photocatalytic oxidation ability.

2.2. Noble-Metal Deposition

Coupling with one or more noble metal nanoparticles, such as Pt, Au, Ag, and Pd, is another well-developed surface modification approach. First of all, the noble metal nanoparticles function as electron traps and realize the rapid electron migration from photocatalyst to metal nanoparticles through the Schottky barrier formed at the interface between the semiconductor and the noble metals. The surface plasma resonance properties of noble metals can promote photocatalyst excitation under visible light and thus facilitate interfacial electron transfer [25]. Furthermore, the Fermi levels of the noble metal nanoparticles are generally lower than the CB of TiO2; therefore, the band gap structure could be modified. For instance, four noble metals (Pt, Pd, Ag, and Au) were loaded onto the TiO2 surface, and it was found that photoactivity of the noble metal–deposited TiO2 increased with the amount of noble metals [26]. Among these metals, 0.2 wt% Ag–deposited TiO2 exhibited the highest phenol degradation rate (0.51 μmol∙dm−3∙min−1) under 60 min visible light irradiation, which is five times that of pure TiO2. It is pointed out that Ag nanoparticles are deposited on the {101} crystal facet of anatase TiO2 according to TEM analysis, which results in decreasing photoinduced electron-hole recombination and visible light absorption extension via acting as electron traps and the surface plasma resonance effect.

2.3. Heterojunction Design

Heterojunction construction is one of the most commonly applied and effective surface modification approaches, coming from the combination of two or more semiconductors with different VB and CB positions and the resultant narrower band gap structure. The referred heterojunction is the type-II heterojunction and could be classified into four types such as the conventional type-II heterojunction, P–N junction, surface heterojunction and direct Z-scheme heterojunction, depending on the charge carrier separation mechanism [27]. Generally, the photocatalytic activity of TiO2 could be enhanced by designing a heterojunction structure with several oxides and non-oxide materials [28][29][30]. For example, g-C3N4/TiO2 heterojunction composite was synthesized via the in situ hydrothermal method, and the obtained band gap energy was proved to be 2.8 eV compared with 3.0 eV of TiO2 [31]. The g-C3N4/TiO2 sample with a mass ratio of 1:4 exhibited the maximum hydrogen evolution rate (1041 μmol/(g∙h)) under visible light irradiation, which could be attributed to the heterojunction formation and its spatial charge separation effect. The XPS valence spectra results confirm the interaction between g-C3N4 and TiO2 as type-II heterojunction, and the mechanism for photocatalytic performance improvement is explained such that the electron migration from VB of g-C3N4 to CB of TiO2 via C-O-Ti bonding can offer electrons more photocatalytic reaction time. TiO2 is known as a typical n-type semiconductor and can be used to construct p–n heterojunctions by combining it with a p-type semiconductor. As a result, BiOI was selected to couple with TiO2, and it was found that the 20% BiOI/TiO2 heterostructure achieved the highest phenol conversion efficiency (13.5%) and the best selectivity (92.1%) [32]. The enhanced photocatalytic performance is ascribed to a narrower band gap and efficient photoinduced electron-hole separation due to the p–n heterojunction structure as well as more exposed reactive facets.

2.4. Support Material Combination

Over the past decade, loading the well-developed photocatalysts onto special supporting materials with unique chemical properties or morphology has become more popular. Compared with photocatalyst modification methods such as ion doping, support material combinations have different characteristics in promoting photocatalysts. Support material generally only influences the photocatalytic reaction rate instead of the reaction process. On the one hand, when deposited onto the surface of minerals, the aggregation of photocatalysts could be greatly meliorated, which considerably decreases the chance of photo-generated electron-hole recombination [33][34]. For example, the addition of reduced graphene oxide into the titanium oxide matrix could hinder the agglomeration of photocatalyst nanoparticles and efficiently promote photoinduced electron-hole immigration [35]. On the other hand, the combination of minerals and photocatalysts could obtain better adsorptive ability, which is favorable for the concentration of pollutants near the photocatalysts and shortens the transfer route of photo-induced charge carriers. With high specific surface area and porosity, porous diatomite could cooperate with silver phosphate to harvest visible light and adsorb dye molecules [36]. In many cases, minerals have some active species that function differently in photocatalytic reaction systems. Some active species in minerals could be combined with photocatalysts and modify the crystalline and electronic structure of photocatalysts, similar to the mechanism of ion doping and heterojunction construction. While other active species could act as the separation center of photo-induced electron and hole pairs and vary the transfer route of the charge carriers. The most promising directions in recent decades are more postulates of mineral support influences and further elucidations of the known influences.

3. Natural Carbonate or Carbon-Based Mineral-Supported Photocatalysts

3.1. Calcite-Supported Photocatalysts

As the most stable polymorph of calcium carbonate, calcite is extremely common in sedimentary, metamorphic, and igneous rocks throughout the world and is thus considered a rock-forming carbonate mineral. In 2019, Belhouchet et al. made full use of the natural Algerian shale abundantly existing in northeastern Algeria and North Africa to search for highly efficient photocatalytic material [37]. The study results revealed that the natural mineral mainly consisted of calcite coated with TiO2 by the sol-gel method, and the composite CAL30 sample, had the largest pore volume, higher specific surface area, and, therefore, the best photocatalytic activity. It was analyzed that this calcite/TiO2 system followed a pseudo-first-order kinetic model from the photocatalytic experiment results. The introduction of the mineral into CAL30 sample could suppress the electron-hole recombination and reach a close tetracycline degradation performance (the rate constant k = 2 × 10−2 min−1) of pure TiO2, saving 70% of the TiO2 amount.
Additionally, limestone and marble generally make up the Earth’s crust, a significant part with primary constituent as calcite, and serve as one of the largest carbon storages on this planet. Generally speaking, limestone as sedimentary carbonate rock is often formed by the skeletal fragments of marine organisms and the chemical precipitation of calcium carbonate. Because of the water solubility, weak acid properties, and wide distribution, limestone is usually used as a building material, white pigment, and product filler [38]. However, regardless of whether it is a building material or wall painting material, limestone always encounters a cleaning problem. As a result, self-cleaning technology design is put forward to solve this issue economically and efficiently. In 2017, N. Fadhilah could successfully synthesize super-hydrophilic self-cleaning limestone coating paint with TiO2 nanoparticles modified by dispersant polyethylene glycol 6000 (PEG6000) [39]. This new paint with the composition of TiO2:PEG600 as 1:6 could achieve the highest self-cleaning ability (83.1%) with water contact angles of 7°. Up to 2019, Mohsen et al. prepared CaCO3/TiO2 nanostructures with high surface area, mesoporous structure, and lower bang gap energy of 2.05 eV by the precipitation method [40]. Thus, the CaCO3/TiO2 nanocomposite showed photocatalytic desulfurization of 95% dibenzothiophene under 550 Watt linear halogen lamp illumination and in the presence of H2O2 and acetic acid.
Often referred to as magnesium-rich limestone, dolomite is known as an anhydrous carbonate mineral composed of calcium magnesium carbonate and crystallizes in the trigonal-rhombohedral system. With the ideal formula CaMg(CO3)2, dolomite has double carbonate-containing alternating planes of Ca2+ and Mg2+ ions. In 2019, Belarbi et al. calcined dolomite under 800 °C and then modified it by Ca(NO3)2 to obtain the dolomite-based catalyst CaD800, which could degrade 95% pentachlorophenol under UV light irradiation after only 1 h [41].

3.2. Other Carbon Material-Supported Photocatalysts

Apart from graphene, other carbon materials with specific morphology are also excellent candidates for photocatalyst support, such as carbon quantum dots, carbon nanotubes, carbon nanofibers, etc. [42]. Similar to graphene, these carbon nanomaterials also possess superior mechanical and electrical properties. In one benzene mineralization study, carbon nanotubes were combined with TiO2 by the wet impregnation method. Carbon nanotubes as support could provide a stable electron reservoir and efficient electron transfer route. Moreover, the existence of carbon nanotubes effectively suppresses the agglomeration of TiO2 particles, and thus more reactive sites could be exposed. Compared with the bare TiO2, the enhanced photocatalytic performance of TiO2/CNT composite was proved to come from tight interfacial contact between carbon nanotubes and TiO2. Due to all these factors, the conversion efficiency from benzene to CO2 over TiO2/CNT is much higher than P25, a kind of commercial TiO2 [43].

4. Natural Sulfate or Sulfide Mineral—Supported Photocatalysts

4.1. Gypsum-Supported Photocatalysts

As the most common sulfate mineral, gypsum is a soft calcium sulfate dihydrate compound that is often used as a fertilizer and as the main constituent in many forms of plaster, blackboard, and drywall. Generally, gypsum deposited from lake, sea, and hot springs crystallizes as a translucent crystal of selenite and forms as an evaporite mineral and hydration product of anhydrite. The previous research mainly focuses on assembling efficient photocatalytic additives with commercial gypsum to develop self-cleaning building materials and exploring their photodegradation efficiency on various pollutants. A novel BiOClxBr1−x@gypsum heterojunction composite was constructed as a self-cleaning surface for building material [44]. The photocatalytic activity results showed that the as-obtained hybrid material could mineralize RhB in 4 min and the recalcitrant Naphthalene contamination in 20 min under visible light irradiation. Additionally, it could maintain superior photocatalytic activity and efficiency for at least 7 months. Janus et al. incorporated N, C co-modified TiO2 into commercial gypsum to obtain a self-cleaning property and exhibit obvious dye removal capacity [45]. They found that N, C-TiO2 photocatalyst functioned as the void space fillers in gypsum matrix could reduce gypsum particle size and over 10% mass content of N, C-TiO2 additive into gypsum causes a considerable decrease of the compressive strength value, pore volume, and specific surface area. The obtained N, C-TiO2/gypsum sample treated at 300 °C exhibited 85% and 70% degradation of azo dye Remazol red 198 after 15 h of UV and visible light irradiation, respectively. V. Binas et al. fabricated photocatalytic cement-based coating mixed with Mn-doped TiO2 and compared the influence of different supporting materials on the photocatalytic activity against NO air pollutants [46]. The substrate materials displayed a slight difference in photocatalytic degradation efficiency, and the mixture of Mn-doped TiO2 composite coating influenced its surface area and porosity.

4.2. Pyrite-Supported Photocatalysts

As a naturally n-type semiconductor material, pyrite is a brass-yellow sulfide mineral with metallic luster and chemical formula of FeS2. Considering its broad existence in igneous, metamorphic, and sedimentary rocks worldwide, pyrite is widely regarded as a promising candidate for wastewater treatment. Diao et al. employed pyrite as a photocatalyst for the degradation of malachite green (MG) and Cr(VI) reduction [47]. Pyrite photocatalysts showed removal efficiency of MG and Cr(VI) reached 96.2% and 53.1%, respectively, from their mixture solution after 120 min UV light irradiation, which is better than that determined for individual separated solutions. They suggested that the presence of malachite green could accelerate Cr(VI) reduction and vice versa, which is attributed to the synergistic effect in the MG and Cr(VI) coexistence system. They also found that the dissolved oxygen could remarkably promote MG degradation efficiency and Cr(VI) reduction that is mainly controlled by direct electrons of pyrite. Furthermore, natural pyrite could be applied to inactivate Escherichia coli in the presence of citrate and EDTA as chelating agents [48]. These chelating agents would increase the iron solubility and reduce the oxidation of the pyrite surface, which promotes the photo-Fenton process and the E. coli inactivation rate. Additionally, natural pyrite could be transformed structurally into a Z-scheme FeS2/Fe2O3 composite via the heat treatment method, and the synergistic effects between FeS2 and Fe2O3 have a remarkable positive influence on its photocatalytic activity [49]. It was proved that the as-obtained composite exhibits increased photocatalytic degradation efficiency from 36% to 65% with the presence of 30 mg/L Cr(VI), and the addition of Cr(VI) ions can also be almost reduced. That is related to the fact that Cr(VI) could capture the electrons and thus improve photo-induced electron-hole pair separation efficiency. These findings make natural pyrite suitable for designing a low-cost and efficient photocatalyst for wastewater treatment and environmental applications.

5. Natural Special Mineral-Based Photocatalysts

5.1. Ilmenite-Based Photocatalysts

As a wide-bandgap anti-ferromagnetic semiconductor, ilmenite is known as a titanium–iron oxide mineral with the idealized chemical formula FeTiO3, containing around 40–60% of TiO2. Generally, pure ilmenite has a rhombohedral crystal structure with the hexagonal packing of the oxygen atoms with cations occupying two-thirds of the octahedral positions, where Fe and Ti occupy alternating layers [50]. Due to its chemical composition and unique properties, ilmenite is always considered the most suitable natural precursor of photocatalyst TiO2. A. Simpraditpan et al. explored the effect of calcination temperature on the properties of titanate nanofibers originated from natural ilmenite minerals and concluded that the temperature increase could result in the phase transformation from thermodynamically metastable anatase phase to the most stable rutile phase with increasing the crystallite size [51]. They found that the nanofibers calcined at high temperatures of 600–1000 °C were mixtures of three crystalline phases, e.g., anatase, rutile, and Fe2O3. Moreover, increasing the calcination temperature induces the increase of crystallite sizes, therefore leading to the absorption edge of the as-synthesized sample shifts to the visible light region. Among these samples, the nanofibers calcined at 400 °C for 2 h exhibited the best photocatalytic activity, even higher than the commercial TiO2. Shao et al. synthesized iron-doped TiO2 nanorods from raw ilmenite ore via a direct acid leaching method. They found that the band gap energy decreased from 2.94 eV to 2.49 eV with higher iron concentration, as compared with 3.02 eV of bare rutile TiO2 [52]. The optimal amount of iron dopant was reported to be 1.69%, leading to the best photogenerated charge carrier separation efficiency. Another study showed that iron species from natural ilmenite minerals could be converted into Fe3O4 in addition to titania phases, providing magnetic recyclability of the photocatalyst under an external magnetic field [53]. This composite was characterized to be composed of face-centered cubic Fe3O4 and body-centered tetragonal anatase TiO2 with superparamagnetic behavior and strong visible-light absorption ability. It could be separated and recycled quickly and easily due to the magnetization effect of Fe3O4 core. Moreover, superoxide radicals were proved to be the main active species of the Rhodamine B degradation process. Xia et al. introduced persulfate into the natural ilmenite reaction system to achieve efficient E. coli inactivation under visible light irradiation [54].

5.2. Monazite-Supported Photocatalysts

A monazite is a group of monoclinic phosphate and arsenate minerals that contains one or several kinds of rare-earth metals. Monazite (Ce), as a thorium-bearing mineral, was selected as TiO2 substrate for its minor radioactivity and its interior framework of a composite on which TiO2 grew by forming Ti–O–Ce microstructure [55]. TiO2 embedded in monazite became a kind of radiation-sensitive photocatalyst, and the degradation principle was similar to photocatalysis in that the band gap of TiO2 is activated by γ-ray irradiation. It was revealed that the radiations of thorium nuclides in TiO2/monazite composite endowed energy to trigger TiO2 excitation, existent cerium ions in monazite served as energy absorber with charge separator, and the TiO2/CeO2 structure could provide a unique electron transfer process which is beneficial to separate the charge-carriers more efficiently and thus improve catalytic activity in dark conditions. Furthermore, radioluminescence (RL) is the phenomenon by which luminescence is produced in a material by the bombardment of α-particles, γ-rays, or β-particles. It was proposed that radio-sensitive Ce could produce interior UV because of RL and TiO2 were excited by RL. This self-activated photocatalytic behavior was similar to the heterojunction reaction mechanism and inner electric field established within TiO2 semiconductor and CeO2 due to Fermi level equilibrium and could transfer electron-hole pairs more effectively.

References

  1. Fujishima, A.; Rao, T.N.; Tryk, D.A. Titanium dioxide photocatalysis. J. Photochem. Photobiol. C Photochem. Rev. 2000, 1, 1–21.
  2. Lu, Y.; Zhao, K.; Zhao, Y.; Zhu, S.; Yuan, X.; Huo, M.; Zhang, Y.; Qiu, Y. Bi2WO6/TiO2/Pt nanojunction system: A UV–vis light responsive photocatalyst with high photocatalytic performance. Colloids Surf. A Physicochem. Eng. Asp. 2015, 481, 252–260.
  3. Rimoldi, L.; Meroni, D.; Cappelletti, G.; Ardizzone, S. Green and low cost tetracycline degradation processes by nanometric and immobilized TiO2 systems. Catal. Today 2017, 281, 38–44.
  4. Zhang, G.; Sun, Z.; Duan, Y.; Ma, R.; Zheng, S. Synthesis of nano-TiO2/diatomite composite and its photocatalytic degradation of gaseous formaldehyde. Appl. Surf. Sci. 2017, 412, 105–112.
  5. Li, C.; Sun, Z.; Zhang, W.; Yu, C.; Zheng, S. Highly efficient g-C3N4/TiO2/kaolinite composite with novel three-dimensional structure and enhanced visible light responding ability towards ciprofloxacin and S. aureus. Appl. Catal. B Environ. 2018, 220, 272–282.
  6. Fu, L.; Guo, Y.; Pan, S.; Huang, J.; Wang, L. Core-shell type composites equipped with carbon dots for high efficiency photocatalyst. Surf. Coat. Technol. 2019, 359, 190–196.
  7. Nada, A.A.; Orimolade, B.O.; El-Maghrabi, H.H.; Koiki, B.A.; Rivallin, M.; Bekheet, M.F.; Viter, R.; Damberga, D.; Lesage, G.; Iatsunskyi, I.; et al. Photoelectrocatalysis of paracetamol on Pd–ZnO/N-doped carbon nanofibers electrode. Appl. Mater. Today 2021, 24, 101129.
  8. Bu, X.; Wu, B.; Long, T.; Hu, M. Preparation, characterization and enhancement of the visible-light photocatalytic activity of In2O3/rectorite composite. J. Alloys Compd. 2014, 586, 202–207.
  9. Li, P.; Huang, L.; Li, Y.; Xu, Y.; Huang, S.; Yuan, D.; Xu, H.; Li, H. Synthesis of dark orange montmorillonite/g-C3N4 composites and their applications in the environment. J. Phys. Chem. Solids 2017, 107, 131–139.
  10. Liu, K.; Tong, Z.; Muhammad, Y.; Huang, G.; Zhang, H.; Wang, Z.; Zhu, Y.; Tang, R. Synthesis of sodium dodecyl sulfate modified BiOBr/magnetic bentonite photocatalyst with Three-dimensional parterre like structure for the enhanced photodegradation of tetracycline and ciprofloxacin. Chem. Eng. J. 2020, 388, 124374.
  11. Xu, H.-Q.; Yang, S.; Ma, X.; Huang, J.; Jiang, H.-L. Unveiling Charge-Separation Dynamics in CdS/Metal–Organic Framework Composites for Enhanced Photocatalysis. ACS Catal. 2018, 8, 11615–11621.
  12. Fatimah, I.; Nurillahi, R.; Sahroni, I.; Muraza, O. TiO2-pillared saponite and photosensitization using a ruthenium complex for photocatalytic enhancement of the photodegradation of bromophenol blue. Appl. Clay Sci. 2019, 183, 105302.
  13. Kumar, S.G.; Devi, L.G. Review on modified TiO2 photocatalysis under UV/visible light: Selected results and related mechanisms on interfacial charge carrier transfer dynamics. J. Phys. Chem. A 2011, 115, 13211–13241.
  14. Wenderich, K.; Mul, G. Methods, mechanism, and applications of photodeposition in photocatalysis: A review. Chem. Rev. 2016, 116, 14587–14619.
  15. Jiang, L.; Yuan, X.; Pan, Y.; Liang, J.; Zeng, G.; Wu, Z.; Wang, H. Doping of graphitic carbon nitride for photocatalysis: A reveiw. Appl. Catal. B Environ. 2017, 217, 388–406.
  16. El Gaidoumi, A.; Doña Rodríguez, J.M.; Pulido Melián, E.; González-Díaz, O.M.; Navío Santos, J.A.; El Bali, B.; Kherbeche, A. Synthesis of sol-gel pyrophyllite/TiO2 heterostructures: Effect of calcination temperature and methanol washing on photocatalytic activity. Surf. Interfaces 2019, 14, 19–25.
  17. Hao, R.; Wang, G.; Tang, H.; Sun, L.; Xu, C.; Han, D. Template-free preparation of macro/mesoporous g-C3N4/TiO2 heterojunction photocatalysts with enhanced visible light photocatalytic activity. Appl. Catal. B Environ. 2016, 187, 47–58.
  18. Shayegan, Z.; Lee, C.-S.; Haghighat, F. TiO2 photocatalyst for removal of volatile organic compounds in gas phase—A review. Chem. Eng. J. 2018, 334, 2408–2439.
  19. Nada, A.A.; Bekheet, M.F.; Viter, R.; Miele, P.; Roualdes, S.; Bechelany, M. BN/GdxTi(1-x)O(4-x)/2 nanofibers for enhanced photocatalytic hydrogen production under visible light. Appl. Catal. B Environ. 2019, 251, 76–86.
  20. Nada, A.A.; Bekheet, M.F.; Roualdes, S.; Gurlo, A.; Ayral, A. Functionalization of MCM-41 with titanium oxynitride deposited via PECVD for enhanced removal of methylene blue. J. Mol. Liq. 2019, 274, 505–515.
  21. Nada, A.A.; El Rouby, W.M.A.; Bekheet, M.F.; Antuch, M.; Weber, M.; Miele, P.; Viter, R.; Roualdes, S.; Millet, P.; Bechelany, M. Highly textured boron/nitrogen co-doped TiO2 with honeycomb structure showing enhanced visible-light photoelectrocatalytic activity. Appl. Surf. Sci. 2020, 505, 144419.
  22. Jiang, P.; Xiang, W.; Kuang, J.; Liu, W.; Cao, W. Effect of cobalt doping on the electronic, optical and photocatalytic properties of TiO2. Solid State Sci. 2015, 46, 27–32.
  23. Jia, T.; Fu, F.; Yu, D.; Cao, J.; Sun, G. Facile synthesis and characterization of N-doped TiO2/C nanocomposites with enhanced visible-light photocatalytic performance. Appl. Surf. Sci. 2018, 430, 438–447.
  24. Tamai, K.; Hosokawa, S.; Asakura, H.; Teramura, K.; Tanaka, T. Low-temperature NOx trapping on alkali or alkaline earth metal modified TiO2 photocatalyst. Catal. Today 2019, 332, 76–82.
  25. Ben Saber, N.; Mezni, A.; Alrooqi, A.; Altalhi, T. A review of ternary nanostructures based noble metal/semiconductor for environmental and renewable energy applications. J. Mater. Res. Technol. 2020, 9, 15233–15262.
  26. Diak, M.; Grabowska, E.; Zaleska, A. Synthesis, characterization and photocatalytic activity of noble metal-modified TiO2 nanosheets with exposed facets. Appl. Surf. Sci. 2015, 347, 275–285.
  27. Low, J.; Cheng, B.; Yu, J. Surface modification and enhanced photocatalytic CO2 reduction performance of TiO2: A review. Appl. Surf. Sci. 2017, 392, 658–686.
  28. El-Maghrabi, H.H.; Nada, A.A.; Roualdes, S.; Bekheet, M.F. Design of Ni/NiO–TiO2/rGO nanocomposites on carbon cloth conductors via PECVD for electrocatalytic water splitting. Int. J. Hydrogen Energy 2020, 45, 32000–32011.
  29. Ghorbanloo, M.; Nada, A.A.; El-Maghrabi, H.H.; Bekheet, M.F.; Riedel, W.; Djamel, B.; Viter, R.; Roualdes, S.; Soliman, F.S.; Moustafa, Y.M.; et al. Superior efficiency of BN/Ce2O3/TiO2 nanofibers for photocatalytic hydrogen generation reactions. Appl. Surf. Sci. 2022, 594, 153438.
  30. Avcıoğlu, C.; Avcıoğlu, S.; Bekheet, M.F.; Gurlo, A. Solar hydrogen generation using niobium-based photocatalysts: Design strategies, progress, and challenges. Mater. Today Energy 2022, 24, 100936.
  31. Alcudia-Ramos, M.A.; Fuentez-Torres, M.O.; Ortiz-Chi, F.; Espinosa-González, C.G.; Hernández-Como, N.; García-Zaleta, D.S.; Kesarla, M.K.; Torres-Torres, J.G.; Collins-Martínez, V.; Godavarthi, S. Fabrication of g-C3N4/TiO2 heterojunction composite for enhanced photocatalytic hydrogen production. Ceram. Int. 2020, 46, 38–45.
  32. Li, B.; Chen, X.; Zhang, T.; Jiang, S.; Zhang, G.; Wu, W.; Ma, X. Photocatalytic selective hydroxylation of phenol to dihydroxybenzene by BiOI/TiO2 p-n heterojunction photocatalysts for enhanced photocatalytic activity. Appl. Surf. Sci. 2018, 439, 1047–1056.
  33. Carriazo, J.G.; Moreno-Forero, M.; Molina, R.A.; Moreno, S. Incorporation of titanium and titanium–iron species inside a smectite-type mineral for photocatalysis. Appl. Clay Sci. 2010, 50, 401–408.
  34. Dong, G.; Zhao, L.; Wu, X.; Zhu, M.; Wang, F. Photocatalysis removing of NO based on modified carbon nitride: The effect of celestite mineral particles. Appl. Catal. B Environ. 2019, 245, 459–468.
  35. Cao, Y.-C.; Fu, Z.; Wei, W.; Zou, L.; Mi, T.; He, D.; Yan, C.; Liu, X.; Zhu, Y.; Chen, L.; et al. Reduced graphene oxide supported titanium dioxide nanomaterials for the photocatalysis with long cycling life. Appl. Surf. Sci. 2015, 355, 1289–1294.
  36. Padmanabhan, S.K.; Pal, S.; Licciulli, A. Diatomite/silver phosphate composite for efficient degradation of organic dyes under solar radiation. Bull. Mater. Sci. 2020, 43, 295.
  37. Belhouchet, N.; Hamdi, B.; Chenchouni, H.; Bessekhouad, Y. Photocatalytic degradation of tetracycline antibiotic using new calcite/titania nanocomposites. J. Photochem. Photobiol. A Chem. 2019, 372, 196–205.
  38. Colangiuli, D.; Lettieri, M.; Masieri, M.; Calia, A. Field study in an urban environment of simultaneous self-cleaning and hydrophobic nanosized TiO2-based coatings on stone for the protection of building surface. Sci. Total Environ. 2019, 650, 2919–2930.
  39. Fadhilah, N.; Etruly, N.; Muharja, M.; Sawitri, D. Self-Cleaning Limestone Paint Modified by Nanoparticles TiO2 Synthesized from TiCl3 as Precursors and PEG6000 as Dispersant. Bull. Chem. React. Eng. Catal. 2017, 12.
  40. Mostafa, M.M.; El saied, M.; Morshedy, A.S. Novel Calcium Carbonate-titania nanocomposites for enhanced sun light photo catalytic desulfurization process. J. Environ. Manag. 2019, 250, 109462.
  41. Belarbi, I.; Çoruh, A.; Hamacha, R.; Marouf-Khelifa, K.; Khelifa, A. Development and characterization of a new dolomite-based catalyst: Application to the photocatalytic degradation of pentachlorophenol. Water Sci. Technol. 2019, 79, 741–752.
  42. Yi, H.; Huang, D.; Qin, L.; Zeng, G.; Lai, C.; Cheng, M.; Ye, S.; Song, B.; Ren, X.; Guo, X. Selective prepared carbon nanomaterials for advanced photocatalytic application in environmental pollutant treatment and hydrogen production. Appl. Catal. B Environ. 2018, 239, 408–424.
  43. Xu, Y.-J.; Zhuang, Y.; Fu, X. New Insight for Enhanced Photocatalytic Activity of TiO2 by Doping Carbon Nanotubes: A Case Study on Degradation of Benzene and Methyl Orange. J. Phys. Chem. C 2010, 114, 2669–2676.
  44. Gnayem, H.; Uvarov, V.; Lahad, O.; Sasson, Y. Hybrid bismuth as self-cleaning composites: Novel aspects of a sustainable photocatalytic technology for solar environmental cleanup. RSC Adv. 2015, 5, 66650–66656.
  45. Janus, M.; Zatorska, J.; Zając, K.; Kusiak-Nejman, E.; Czyżewski, A.; Morawski, A.W. The mechanical and photocatalytic properties of modified gypsum materials. Mater. Sci. Eng. B 2018, 236, 1–9.
  46. Binas, V.; Papadaki, D.; Maggos, T.; Katsanaki, A.; Kiriakidis, G. Study of innovative photocatalytic cement based coatings: The effect of supporting materials. Constr. Build. Mater. 2018, 168, 923–930.
  47. Diao, Z.-H.; Xu, X.-R.; Liu, F.-M.; Sun, Y.-X.; Zhang, Z.-W.; Sun, K.-F.; Wang, S.-Z.; Cheng, H. Photocatalytic degradation of malachite green by pyrite and its synergism with Cr(VI) reduction: Performance and reaction mechanism. Sep. Purif. Technol. 2015, 154, 168–175.
  48. Kalantary, R.R.; Moradi, M.; Pirsaheb, M.; Esrafili, A.; Jafari, A.J.; Gholami, M.; Vasseghian, Y.; Antolini, E.; Dragoi, E.-N. Enhanced photocatalytic inactivation of E. coli by natural pyrite in presence of citrate and EDTA as effective chelating agents: Experimental evaluation and kinetic and ANN models. J. Environ. Chem. Eng. 2019, 7, 102906.
  49. Guo, Q.; Tang, G.; Zhu, W.; Luo, Y.; Gao, X. In situ construction of Z-scheme FeS2/Fe2O3 photocatalyst via structural transformation of pyrite for photocatalytic degradation of carbamazepine and the synergistic reduction of Cr(VI). J. Environ. Sci. 2021, 101, 351–360.
  50. Puthirath Balan, A.; Radhakrishnan, S.; Kumar, R.; Neupane, R.; Sinha, S.K.; Deng, L.; de los Reyes, C.A.; Apte, A.; Rao, B.M.; Paulose, M.; et al. A Non-van der Waals Two-Dimensional Material from Natural Titanium Mineral Ore Ilmenite. Chem. Mater. 2018, 30, 5923–5931.
  51. Simpraditpan, A.; Wirunmongkol, T.; Pavasupree, S.; Pecharapa, W. Effect of calcination temperature on structural and photocatalyst properties of nanofibers prepared from low-cost natural ilmenite mineral by simple hydrothermal method. Mater. Res. Bull. 2013, 48, 3211–3217.
  52. Shao, S.; Yu, J.; Love, J.B.; Fan, X. An economic approach to produce iron doped TiO2 nanorods from ilmenite for photocatalytic applications. J. Alloys Compd. 2021, 858, 158388.
  53. Hong, T.; Mao, J.; Tao, F.; Lan, M. Recyclable Magnetic Titania Nanocomposite from Ilmenite with Enhanced Photocatalytic Activity. Molecules 2017, 22.
  54. Xia, D.; He, H.; Liu, H.; Wang, Y.; Zhang, Q.; Li, Y.; Lu, A.; He, C.; Wong, P.K. Persulfate-mediated catalytic and photocatalytic bacterial inactivation by magnetic natural ilmenite. Appl. Catal. B Environ. 2018, 238, 70–81.
  55. Yu, C.; Wang, C.-F.; Chen, T.; Chang, Y. Synthesis and characterization of radiation sensitive TiO2/monazite photocatalyst. J. Radioanal. Nucl. Chem. 2008, 277, 337–345.
More
Information
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , , ,
View Times: 273
Revisions: 2 times (View History)
Update Date: 16 Aug 2022
1000/1000