Next Article in Journal
Supplementation of SDF1 during Pig Oocyte In Vitro Maturation Improves Subsequent Embryo Development
Next Article in Special Issue
First-Principles-Based Optimized Design of Fluoride Electrolytes for Sodium-Ion Batteries
Previous Article in Journal
Grape Pomace Polyphenols as a Source of Compounds for Management of Oxidative Stress and Inflammation—A Possible Alternative for Non-Steroidal Anti-Inflammatory Drugs?
Previous Article in Special Issue
Modification of the Surface Composition of PTB7-Th: ITIC Blend Using an Additive
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Advances in Hybrid Composites for Photocatalytic Applications: A Review

1
Department of Physics, University of Cagliari, S.P. No. 8 Km 0.700, 09042 Monserrato, Italy
2
Department of Chemical and Geological Science, University of Cagliari, S.P. No. 8 Km 0.700, 09042 Monserrato, Italy
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(20), 6828; https://doi.org/10.3390/molecules27206828
Submission received: 7 September 2022 / Revised: 7 October 2022 / Accepted: 9 October 2022 / Published: 12 October 2022
(This article belongs to the Special Issue Advanced Materials for Energy and Sustainability)

Abstract

:
Heterogeneous photocatalysts have garnered extensive attention as a sustainable way for environmental remediation and energy storage process. Water splitting, solar energy conversion, and pollutant degradation are examples of nowadays applications where semiconductor-based photocatalysts represent a potentially disruptive technology. The exploitation of solar radiation for photocatalysis could generate a strong impact by decreasing the energy demand and simultaneously mitigating the impact of anthropogenic pollutants. However, most of the actual photocatalysts work only on energy radiation in the Near-UV region (<400 nm), and the studies and development of new photocatalysts with high efficiency in the visible range of the spectrum are required. In this regard, hybrid organic/inorganic photocatalysts have emerged as highly potential materials to drastically improve visible photocatalytic efficiency. In this review, we will analyze the state-of-art and the developments of hybrid photocatalysts for energy storage and energy conversion process as well as their application in pollutant degradation and water treatments.

1. Introduction: Importance and Limits

The research and the development of new solutions that strengthen industrial and technological progress, guaranteeing sustainable impacts, is greatly boosting the effort of the scientific community. Modern household combustion devices, motor vehicles, and industrial facilities are common sources of air and water pollution. Particulate matter, nitrogen oxide, sulfur dioxide, volatile organic compounds, dioxins, and polycyclic aromatic hydrocarbons are considered pollutants that are greatly harmful to humans. Outdoor and indoor air pollution cause respiratory and other diseases and are important sources of morbidity and mortality. Policies to reduce air and water pollution represent a winning strategy for both climate and health, lowering the burden of disease attributable to air pollution, as well as contributing to the near- and long-term mitigation of climate change. An interesting review of the environmental and health impact of air pollution is reported by Manisaldis et al., [1] analyzing the sources of environmental pollution in relation to public health and environmental effects. Boosting education, training, public awareness, and public participation are some of the relevant actions indicated in the Paris Agreement of 2015, issued by the UNFCCC (United Nations Climate Change Committee) for maximizing the opportunities to achieve a sustainable impact by the anthropogenic actions, mitigating the climate change and environmental pollution [2]. The raising of environmental technologies can greatly support political/social actions and virtuous daily behaviors. Some examples can be sketched, such as the use of renewable energy obtained by sunlight, wind, rain, waves, and geothermal heat, the definition of smart mass transportation, and highly efficient electric vehicles. However, several factors should be considered in the definition and the use of “green” technologies, starting from the materials and their complete cycle of life, the efficiency, the environmental impact, and the up-grading possibilities. On these bases, photocatalysis is strongly indicated as a promising technology having a fundamental role in applications such as fuel cells, solar conversion, healthy treatments, and environmental radiation (Figure 1). If water splitting has been intensively studied as a possible approach to achieve the effective conversion of solar into chemical energy, photocatalysis is also utilized to decrease the number of organic compounds in the air, water, or soil, representing a high environmental risk for health.
Photocatalysis has been applied to degrade several types of organic waste, such as dyes [3,4], phenolic compounds [5,6], organohalides, and petroleum hydrocarbons [7], but also for the removal of heavy metals such as Chromium, Lead, or Mercury among the others [8,9]. The photocatalytic techniques have been applied even to the degradation of pharmaceutical compounds (PC) that represent a new and emerging source of waste in water and soils [10]. Antibiotics and their by-products have high toxicity, good stability, and great potential to interfere with the environment and ecological environment. Many antibiotics are discharged into the water environment through sewage and animal feces, causing severe water and soil environmental problems [11]. Anti-inflammatory drugs have a high polarity and strong hydrophilicity, but the absorption coefficient in the soil is low, so they are easy to survive underground, surface, and even drinking water resources, causing great pollution to water resources.
The term “photocatalysis” and the base of the effect are already known from the end of the 19th century when Giacomo Ciamician defined the term photocatalysis for the first time. Successively, it this worth mentioning that the first results on photocatalytic water splitting using a TiO2 electrode were reported by Fujishima and Honda in 1972 [12]. After that, the research on this field strongly increased, and in the last 20 years, the number of publications on “photocatalysis” has constantly grown up to 8737 publications in 2021 (Figure 2).
Moreover, the “old” knowledge, the research in this field, and, mostly, the development of new materials with higher efficiency is a mandatory step to overcome the actual technical limits of the processes. Most photocatalysts can utilize the sizeable visible component of the solar spectrum very partially, working on higher energy photons (UV range) that count for only 5% of the solar emission.
Recently different strategies and materials have been utilized to overcome this issue: lowering the band gap of the materials, new energetic levels by doping process, and heterostructures with increased charge separation are just a few examples that we better discuss in the review. On the other hand, the achievement, or at least the improvement in this direction, has a strong potential impact on our daily life and, mostly, on the environmental aspect for a long-time vision of sustainable progress.
Sunlight is a free, perpetual, and clean energy source. If we consider that photocatalysis is a light-induced chemical pathway able to convert organic material into energy, the benefits of this process appear immediate. However, even if the energy produced from the sun is renewable and gives us a plethora of advantages, the materials needed to convert it into practical use should be carefully chosen to guarantee an environmentally friendly process. They should be free from potential impact connected to the extraction of the raw materials and precursors, the use of solvents in the synthesis procedure, and the energy required in the whole life cycle of the photocatalysts should be carefully checked [13,14].
Among the wide choice in terms of possible materials, photocatalysts can be classified into two main categories: semiconductors (organic or inorganic) and metal nanoparticles. Transition metals such as gold, copper, and silver are characterized by localized surface plasmon resonance (LSPR), while semiconducting materials such as TiO2, ZnO, SnO2, MoS2, Bi2MoO6 [15,16,17,18,19,20] assure complete mineralization, no waste disposal problem, low cost and necessity of mild temperature and pressure conditions. Organic materials show mechanical and chemical properties that radically differentiate them from inorganic ones. The absence of metals, the high possibilities to tune the synthesis technique, and the high varieties of optical and structural character of the organic photocatalyst, justify the high interest in this class of materials. Within the organic chemistry panorama, polymers can guarantee very high versatility in terms of structural and optical response with high potential in several applications, but their efficiency is generally hampered by the low thermal and short-time structural stability of the compounds. As already mentioned, different thousands of articles and studies are constantly published, and the interest of the scientific community is strongly active in this field. On these bases, it is important to summarize time by time, the state-of-the-art progress in the design of a variety of semiconductor composite photocatalysts for energy and environmental applications. In recent years different review articles were focalized on photocatalysis, from base properties [21,22,23,24] to materials [25,26,27,28] and to applications.
Inside these very broad categories, several aspects have been considered. If Humayun et al. [29] pointed out the synthesis procedures of composite photocatalysts and the use of microwave is underlined, other works focalize on a specific class of materials: MOF [25], metal halide perovskite [26], graphene [28], and carbon nitride [30].
The use of the photocatalytic process in water treatments [31], for the removal of Volatile Organic compounds (VOC) [32], and the abatement of plastics [33] are just a few recent examples, such as its use for the degradation of persistent herbicides [34] or, more in general, for the development of sustainable agriculture [35].
In this review, we will briefly analyze the state art in inorganic and organic semiconductor photocatalysts, mainly focalizing our discussion on the method to obtain efficient processes using visible light. We will underline the efforts of the scientific community to build efficient heterostructures, and we will analyze the new solutions based on hybrid systems, organic-inorganic, that efficiently work with a relative portion of the solar spectrum.

2. Basic Principles of Photocatalysis

A Photocatalytic process occurs when the light interacts with the surface of semiconductor materials, the so-called photocatalysts, allowing the photo redox process. During this process, there must be at least two simultaneous reactions occurring, oxidation from photogenerated holes and reduction from photogenerated electrons (Figure 3) [36,37,38].
The process can be divided into four fundamental steps: (I) light absorption to generate electron-hole pairs; (II) separation of excited charges; (III) transfer of electrons and holes to the surface of photocatalysts; (IV) utilization of charges on the surface for redox reactions [39]. The holes left in the valence band have high oxidizing power and, reacting with water, generate hydroxyl radicals responsible for pollutant degradation. The electrons in the conduction band, via reaction with dissolved oxygen species, form superoxide ions that promote the reduction process [36,40].
In the oxidation process, the positive holes in the valence band allow oxidizing of the water at the surface of the catalyst, forming hydroxyl radicals (OH.) with a high oxidative decomposing power. These hydroxyl radicals react with the organic pollutant that, at the end of the process, is decomposed into carbon dioxide and water. As well, the organic compounds could react directly with the positive holes created in the valence band, activating an oxidative decomposition.
The reduction process of the oxygen contained in the air is an alternative reaction to hydrogen generation since oxygen is an easily reducible substance. The electrons in the conduction band, reacting with the dissolved oxygen species, form superoxide anions that, after being attached to the intermediate products obtained from the oxidative reaction, form peroxide or hydrogen peroxide and then water as the final product.
The photocatalytic activity of a material depends on several parameters: morphology, size, surface area, reaction temperature, pH, light intensity, amount of catalyst, and concentration of wastewater dependent. The structure of the catalyst plays a fundamental role in the photocatalytic activity: the stability of the catalyst, the position of the conduction band, the degree of hydroxylation and the absorption power, and the size of the particles. In general, the smaller the size of the catalyst nanoparticles, the higher the photocatalytic activity. Cerruto et al. systematically correlated the photocatalytic activity to the average sizes and size distributions of TiO2 nanocrystals [41], while more recently, the study was extended to the hydrogen produced by Al-doped SrTiO3 photocatalysts [42] to the visible photocatalytic activity in β-Bi2O3 [42,43], and the reduction of CO2 in Cu/Cu2O nanocrystals [44]. The surface properties significantly influence the efficiency of the catalyst because, during the photocatalytic process, oxidation and reduction reactions take place on the surface of the catalyst [45,46].
On this topic, He et al. found that the differences in the photocatalytic activity of different shaped samples of ZnO are not caused by the changes in specific surface area, oxygen defects, or particle size, but it was attributed to the pore size. The pore size determines the adsorption efficiency of organic pollutants and dissolved oxygen, and good adsorption performance is the main prerequisite for efficient catalytic degradation [47].
An increase in the temperature during the photocatalytic process induces a decrease in photocatalytic activity. Since the activation is due only to photonic interaction, the photocatalytic systems can operate at room temperature [48]. Even pH plays a key role in photocatalytic reactions because it is responsible for the surface charge properties of the photocatalyst. This could be explained in terms of electrostatic interactions between charged particles and contaminants, influencing the absorption and, consequently, the surface properties [49,50,51]. From multiple studies, it was also evaluated that the degradation rate is directly proportional to the catalyst concentration. An increase in the amount of the catalyst during the photocatalytic process is reflected in the highest number of active sites formed on the semiconductor surface and in an increasing number of OH. and O2. radicals [52].
Another fundamental factor in estimating the degradation rate during a photocatalytic process is the type of pollutant and its concentration. Having different pollutants and different irradiation times to complete mineralization is needed [36,48,52,53].

3. Semiconductor Photocatalysts

Photocatalysts are defined as a class of materials that absorb light and can convert it into electrical charges driving the reduction and/or oxidation of species at the catalyst surface. During the last three decades, great attention has been focused on semiconductor metal oxides, selenides, and sulfides as photocatalysts thanks to their bandgap energy between 1.1 and 3.8 eV and their redox potential (Figure 4) [40]. The species with a higher reduction potential have a higher tendency to acquire electrons, while the species with a higher oxidation potential tend to lose them and to be oxidized.
Metal oxides are largely considered the most efficient and widely applied materials, while metal selenides and sulfides are poorly considered for their toxicity and poor stability [46,54]. In general, the photocatalytic mechanism in semiconductors can be sketched with the below-mentioned Equations (1)–(10).
Semiconductor + Light energySemiconductor (ecb + hvb+)
Dye + Semiconductor (hvb+)Oxidation process
Semiconductor (hvb+) + H2OSemiconductor + H+ + OH.
Semiconductor (hvb+) + OHSemiconductor + OH.
Dye + Semiconductor (ecb)Reduction process
Semiconductor (ecb) + O2Semiconductor + O2.−
O2.− + H+HO2.
HO2. + HO2.H2O2 + O2
H2O2 + O2.−OH. + OH + O2
Dye + OH.Degradation Products
Currently, the most attractive photocatalytic materials are n-type semiconductors such as TiO2, WO3, and Fe2O3, because of their high chemical stability and a conduction band edge with a potential level that is negative enough to allow the proton reduction without additional electric bias [55,56,57,58].
As already stated, the first studies of TiO2 as a photocatalyst it is date back to 1972, and then in 1977, Bard applied it for the first time to purify water from CN ions [59].
To date, TiO2 is considered an n-type semiconductor with several advantages (high photoactivity, strong oxidizing capability, cost-effectiveness, low toxicity, and high stability) and represents one of the most widely studied and applied semiconductor photocatalysts. On the other hand, it still suffers, such as all the inorganic semiconductors, of high recombination rate and negligible light harvesting in the visible range.
In order to overcome the bottlenecks above, the research has been focused on improvements in solar spectral absorption and decreasing the recombination rate of these photocatalysts (Figure 5).
The regulation of the energy band structures is the best way to enhance the visible light photocatalytic performance, generally by doping, defects engineering, and surface modification.
Li et al. attribute the increased efficiency of Carbon doped TiO2 nanofiber under visible light to the presence of oxygen vacancies at the surface while the Ti terminals act as reaction sites, resulting in the accumulation of CO* intermediates to produce CH4 [60]. The role of C doping is further analyzed by Nor et al., By DFT calculations. They suggest that C-doped O sites are more stable in bulk than in the subsurface or on the surface, while C-doped Ti sites are more stable on the surface than in the bulk or subsurface. While the presence of C in the O sites introduces impurity states in the band gap that do not affect the band gap energy but induce the formation of electron traps, Carbon in Ti sites induces structural distortions with a reduction in the band gap energy. Both doping processes enhance light absorption in the visible and IR spectrum [61]. The role of lanthanides as doping elements and/or surface modification has been widely studied. The presence of Dy generates a ligand-to-metal charge transfer process in the visible region (500–600 nm), lowering the bandgap to ~2.24 eV [62]. Guetni et al. found that doping TiO2 with Nd allows the material to absorb radiation in the visible range by creating Nd-4f impurity states. The co-doping with Y further increases the Azo dye orange G degradation under a wavelength range of 300–800 nm due to the synergetic effect of intragap levels [63].
The increased visible-light photoactivity of metal-doped TiO2 by intragap levels in the semiconductor band gap requires photons with lower energy to produce carriers in the conduction band. It is well known that in N-doped TiO2, the lowered band gap is due to the formation of N-2p state hybrids with O-2p states in the anatase phase [64]. Further, the improved trapping of electrons at oxygen vacancy sites can inhibit the electron-hole recombination resulting in enhanced photo-activity [65].
A different mechanism is related to the formation of Oxygen vacancies at the surfaces of the semiconductors (related to metallic doping) and, consequently, to F centers with different charges. The loss of an O atom in a metal oxide generates an electron pair trapped in VO (F center), while a positively charged F+ center is due to a single electron in the O vacancy. An unsaturated vacancy generates an F++ center [66]. The presence of charge defects strongly increases the surface reactivity and, as a consequence, the photocatalytic activity, too.
Various metal nanoparticles such as Pt, Au, Pd, Ru, Rh, and Ag have been employed in implementing the semiconductor surface to achieve higher photocatalytic efficiency. High energy irradiation induces Fermi-level equilibration between the semiconductor (like TiO2) and the metal element through charge distribution [67,68], strongly increasing the electron transfer on the semiconductor/metal interfacial system and promoting efficient photocatalytic reactions. The presence of resonance factor related to the reduced dimension of the metal nanoparticle at the surface (mainly Au and Ag) can give act as an added mechanism.
A different approach consists of the use of carbon aerogels materials hybridized with TiO2; the main effect is related to the increase in the surface area of TiO2 due to the great porosity and highly developed hierarchical porous structures. Further, Carbon functionalization results in the formation of new intragap levels that red-shift the absorption in the visible range and, acting as a charge trap, limit the recombination rate [58]. However, there are still some issues to be solved, firstly, the efficiency in the photodegradation should be still increased, but mainly there are some issues related to the synthesis procedure that requires intensive energy consumption and the use of a considerable amount of organic precursor such as formaldehyde [69].
Localized surface plasmonic resonance (LSPR) can be seen as the collective free electron charge oscillation in the metallic nanoparticles optically consistent with the frequency of the incident light wave [70]. This phenomenon usually occurs in nanoparticles below 50 nm and creates an added optical excitation band in the visible region (480 for Ag, 520 for Au) [71].
Yukika et al. coupled small Pt nanospheres with large Au nanocubes on TiO2 to obtain a coupling resonance mode at about 600 nm. Further, the bimetallic coupling allowed charge separation and catalytic reaction at the Pt site and a light-harvesting antenna at the Au site, respectively [72].
It was observed that catalytic reduction of CO2 can be obtained with Au nanoparticles deposited TiO2 under full solar-spectrum irradiation, thanks to the visible contribution associated caused with the LSPR effect of Au NPs [70,73]. A further method to increase the photo-efficiency foresaw the synergic effect of the localized plasmonic features and the formation of metal chiral structures. Indeed, plasmonic materials with chiral configurations have not only emulated the optical properties of their molecular counterparts, but the chiroptical activity can achieve 10 orders of magnitude greater [74,75]. The great enhancement of the electric field, as well as the formation of hot injecting electrons in the semiconductor inside and around the plasmonic nanostructures, are the main player in the mechanism.
Organic semiconductors have recently gained large interest due to the earth’s abundance of their constituent elements, optical and structural properties, and in particular, their tunable energy levels. It is outside the scope of the present paper, but interesting examples of reviews on the use of organic semiconductors in photocatalytic applications are reported in Refs. [76,77].
This class of materials can be divided into two categories: polymers and small molecule materials. Polymeric macro-molecules are constituted by the repetition of a fundamental unit (monomer) and are based on sp2 hybridized carbon atoms, with the alternation of single and double bonds along their building block (Figure 6).
Within the organic chemistry panorama, polymers can guarantee very high versatility in terms of structural and optical response with high potential in several applications [78,79,80], but their efficiency is hampered by the low thermal and short-time structural stability of the compounds (Figure 7).
Covalent organic frameworks (COFs) are a type of purely organic crystalline porous polymeric material with a very high structural tunability due to the high possibility of varying the various functional building blocks [81,82]. COF can be tuned to vary the absorption into the visible range up to infrared with relatively high efficiency. Further, the large specific surface area and regular and customizable micro- and mesopores can offer increased interaction between substrates and catalytic sites. Furthermore, via bottom-up synthesis, post-modification, doping, photosensitization, and hybridization with other photoactive compounds, the construction of the heterojunction structure, the electron-hole recombination can be greatly retarded, the charge transfer processes can be accelerated, and the charge transfer resistance can be greatly decreased [12]. Moreover, the considerably large specific surface area and numerous regular and customizable micro- and mesopores can offer favourable guest-host interaction between substrates and catalytic sites. The building blocks to produce the organic porous architectures are generally rigid and do not possess bulky side groups.
In this scenario, polymeric carbon nitrides have emerged as promising materials for being metal-free semiconductors and metal-free photocatalysts.
First attempts to prepare solid-state carbon nitride materials have been known in the literature since early 1830, and they have eluded rigorous analyses and characterizations for a long time because of their insolubility in water and organic solvents.
Kroke et al., using density functional theory methods (DFT), proposed a polymer based on two-dimensional tri-s-triazine units (heptazines) as the most thermodynamically stable structure of carbon nitride [83]. This polymer, because of its graphitic-like structure, was then called and studied graphitic carbon nitride (g-C3N4) and has attracted strong attention because of its tunable optoelectronic properties and because it is a metal-free semiconductor [84,85,86]. The chemical structure of graphitic carbon nitride has been deeply studied with different techniques, including solid-state nuclear magnetic resonance spectroscopy, Fourier-transform infrared spectroscopy, and x-ray diffractometry [87]. These studies confirmed that g-C3N4 is a layered material, in which van der Walls force holds the stacking layers (C-N bonds), and each layer is composed of tri-s-triazine units, bridged by tertiary amino groups with a high degree of condensation. It is thermally and chemically stable under ambient conditions, biocompatible, eco-friendly, and the graphite-like planar configuration, with a π conjugated system, thereby enabling the transport of charge carriers, and the moderate bandgap makes it workable in the visible part of the solar spectrum [88,89,90,91]. The lone pair electrons of nitrogen are responsible for the formation of the lone pair valence band and the band structure. The combination of the lone pair state and the p bonding stabilizes the lone pair position becoming the key to the electronic structure [92,93].
These appealing properties make g-C3N4 useful for different applications, and from 2000–2019, around 5100 publications have been published on g-C3N4 applications, including photocatalytic water splitting, CO2 reduction, environmental remediation, organic transformation, and white light-emitting diodes [19,94].
Despite these positive attributes, the efficiency of g-C3N4 as a standalone photocatalytic material is diminished by a low quantum efficiency and high recombination rate of photogenerated electron-hole pairs [95]. However, the modification of g-C3N4 could overcome these problems. Recent studies reported that by controlling the morphology and the surface properties of these catalysts, a significant enhancement in photo-reduction is observed. In particular, Porcu et al. studied a new polymer of the family of carbon nitrides with enhanced absorption in the visible part of the solar spectrum [96]. They showed how the protonation method could improve the solubility, dispersibility, and electronic structure, and the exfoliation method of 2D structures, through the breaking of Van der Waals forces, can improve light absorption (Figure 8) [97].

4. Heterostructures

One of the most effective strategies for improving the photocatalytic efficiency of semiconductor materials is the design of heterojunction systems [98,99,100]. A general rational design and fabrication of heterojunction photocatalysts, such as the semiconductor–semiconductor heterojunction, the semiconductor–metal heterojunction, the semiconductor–carbon heterojunction, and the multicomponent heterojunction can be found in the review of Wang et al. [101].
Heterojunction photocatalysts are obtained by coupling two semiconductor photocatalysts and taking advantage of their relative band structure [102,103]. The coupling of the two systems has been classified into Type I, Type II, Z, and S-scheme photocatalysts. (Figure 9) [29,98].
In type I, the heterostructure is formed by two semiconductors, and the photoinduced electron and holes are transferred from one semiconductor to the other, but due to the electron-holes accumulation in one semiconductor, the redox ability is relatively low. Type-II heterojunction has been widely studied and applied, starting in the 1980s [104]. It is based on the simultaneous formation, upon a proper excitation wavelength, of electrons and holes in both photocatalysts. Photogenerated electrons were transferred from one photocatalyst to the other, while photogenerated holes were transferred in the opposite direction. Thus, the photogenerated carriers are spatially separated, and electrons and holes accumulate on different photocatalysts, one for the reduction reactions and the other for oxidation reactions, respectively.
Despite the interesting process and the wide application and studies applied in this scheme [105], the type II scheme presents a low redox ability. Actually, from a dynamic perspective, the repulsion from the existing carrier (electron and carrier) already accumulated in one photocatalyst hampered the transfer process, weakening the overall redox ability of the heterojunction [106,107].
The Z-scheme composites show a high redox potential and very efficient charge carrier separation. The photoinduced holes of semiconductor I recombine with the photoinduced electrons of semiconductor II leaving the photoinduced electrons of semiconductor I and the photogenerated holes of semiconductor II free to be active in the reduction and oxidation process, respectively [108]. These kinds of heterostructures find application in many fields, such as CO2 conversion, water splitting, and pollutant degradation [56,109,110].
The concept of the “direct Z-scheme” has been developed gradually. One of the first successful applications was the dye-sensitized TiO2 by Gratzel [110], where a tandem cell was obtained by coupling nanocrystalline WO3 or Fe2O3 with dye-sensitized TiO2. In this case, holes in the valence band of WO3 or Fe2O3 generated by UV radiation are involved in a water oxidation reaction, while the electrons, transferred to dye-sensitized TiO2, will generate hydrogen in the water reduction reaction. Further, Balayeva et al. developed a surface-grafted WO3/TiO2 hybrid system with high photoactivity under visible light irradiation for indoor air purification [111]. Another Z-scheme application is ZnO-CdS Z-scheme heterojunction for improved H2 evolution performance and, more recently, Cu2O-Pt/SiC/IrOx heterostructure for CO2 reduction and O2 evolution [112]. Indeed, the second generation of Z-scheme photocatalysts considers the use of a mediator between the two materials. The close contact between the photocatalysts and the mediator allows the faster recombination between the photogenerated carriers (electrons from PS II with the photogenerated holes from PS I) due to a lower contact resistance interface [113].
Recently a new scheme has been proposed, the so-called S-Scheme. An S-scheme is like a type-II heterojunction but with a different charge-transfer path. In a typical type-II heterojunction, photogenerated electrons and holes are accumulated on the conduction band and valence band of the two semiconductors resulting in weak redox ability. In an S-scheme heterojunction, the photogenerated electrons and holes are stored in the conduction band of the reduction photocatalysts and the valence band of the Oxidative photocatalyst, respectively, while the other photogenerated charge carriers of the two semiconductors are recombined together [98].
Jang et al. [114] report superior performances for H2O2 production in a heterojunction between ZnO and WO3. The formation of the S-scheme heterojunction (ZnO and WO3) generates electrons on the conduction band of ZnO and holes on the valence band of WO3 that are efficiently spatially separated. Further, CdS/TiO2 heterostructures with a working s-scheme are reported by Wang et al. Due to the higher different working function, electrons from CdS can transfer to TiO2, forming the internal electric field directing from CdS to TiO2. Under illumination, the photogenerated holes in the valence band of CdS recombine with the photogenerated electrons in the conduction band of TiO2 driven, leaving the photogenerated carriers (electron in the conduction band of CdS and photogenerated holes in the valence band of TiO2) free to act to the CO2 reduction and water oxidation reactions, respectively [115].
The S-scheme has been highly utilized in the hybrid organic-inorganic heterostructures, and it will be recalled in the following paragraph.

5. Organic/Inorganic Heterostructures

The development of new materials with higher activity and versatile properties for photocatalytic applications is strongly required. The formation of “hybrid materials”, by coupling an inorganic material (often already in the list of the classical photocatalysts) with new organic structures (often represented by stable polymers) permits to obtain a new class of materials with superior and tunable properties (Figure 10).
Depending on the type of bonding between the two materials, different effects and classifications can be defined. In the first case, the stability of the structure is provided by the presence of physical interactions, such as hydrogen bonding, van der Waals force, or ionic interactions. Proper functional groups at the surface are necessary to stabilize the interaction and stabilize the structure. The second type is defined by stronger interactions generated by chemical bonding, and in general, it defines very stable hybrid materials and a fast charge transfer process.
The organic part in the hybrid system can either be molecular catalysts or organic dyes, which can act as photocatalysts or sensitizers. In this latter case, they will not act as a photocatalyst but support the new system with their properties. The general mechanism is relatively easy: the sensitizer, chemically adsorbed on the semiconductor surface, absorbs the photon with a suitable wavelength passing to an excited state, then the photoexcited carriers are injected into the excited inorganic band (following one of the above descript mechanism, such as Z-scheme and/or S-Scheme). The main scope of the sensitizer is to overcome the physical limitations of the inorganic semiconductors, which generally can absorb only the UV component of the electromagnetic spectrum. Different systems were proposed, where both the inorganic part as well the organic units have been varied. Metal oxides still represent the most common choice, while much more variety can be found in the organic choice.
The π-conjugated systems are promising materials as sensitizers in hybrid systems. Suárez-Méndez et al. [116] report the increased photocatalytic performance of protonated thiophene-based oligomers (OTn+) as sensitizers in an OTn/TiO2 system, while Zhong et al. show high H2 production activity using excitation wavelengths greater than 420 nm in a porous organic-inorganic hybrid system formed by Calix [4] arene dye as sensitizer and titanium dioxide as acceptor units [117].
Coumarins represent a further interesting material for the high absorption cross-section and guarantee an efficient and fast charge transfer to TiO2 through the LUMO level to the conduction bands [118,119], but generally have low structural stability and long-time performances. Hence, among the different organic units, the choice of polymers could generally guarantee increased stability. In particular, the organic-inorganic hybrid system that couples TiO2 to graphitic carbon nitride (g-C3N4) represents an intriguing solution because it meshes the photocatalytic properties of TiO2 with the lower band gap energy (2.7 eV) and thermal and chemical stability of g-C3N4 [120,121]. Two-dimensional conjugated polymers have emerged as a promising class of materials for photocatalytic applications thanks to their photo-physical properties such as: being eco-friendly, consisting of heart-abundant elements, easiness of synthesis, and chemically stable. Additionally, the delocalized conjugated structure of g-C3N4 gives rise to a slow charge recombination rate and a rapid photoinduced charge separation. The formation of a charge-transfer complex at the interface between the organic donor (i.e., g-C3N4) and inorganic acceptor (i.e., TiO2) compound results in a g-C3N4/TiO2 heterojunction system that shows a decrease in the recombination rate of photogenerated electron-hole pairs, as well as an increase in the photocatalytic activity of TiO2 under visible light for wavelengths as low as 450 nm [122,123,124,125]. The band scheme of the band structure of TiO2/g-C3N4 can be seen as an efficient Z-scheme where the internal electric field directing from g-C3N4 to TiO2 was formed because of their different work functions.
Rodrigues et al. reported a photocatalytic activity under visible radiation of the TiO2/g-C3N4 improved by 44.8% compared to pure TiO2, whereas an improvement of 30.5% was obtained under simulated solar radiation, generated by the creation of an effective Z-scheme heterojunction [126]. Further, Liu et al. develop a supramolecular self-assembly approach combined with thermal polycondensation to construct direct Z-scheme-dictated TiO2/g-C3N4 heterojunctions. The hybrid porous nanostructures possess hydrogen-bonded interfacial contact between TiO2 and g-C3N4 with enhanced visible-light photocatalytic activity in the degradation of phenolic pollutants, acetaminophen, and methylparaben, in comparison to the pure TiO2 and g-C3N4 [127].
The heterostructures WO3/g-C3N4 represent the application of an S-scheme, useful for hydrogen production [128]. In this scheme, similar to the previous example, the work function of WO3 (6.23 eV) is larger than that of g-C3N4 (4.18 eV), inducing the formation of an internal electric field from the polymer to the inorganic semiconductor and band-edge bending due to electron redistribution. Upon excitation, the photogenerated holes in the valence band of g-C3N4 recombine with the electrons in the conduction band of WO3, leaving the photogenerated carriers (electron of g-C3N4 and holes of WO3) free to participate in photocatalytic reactions. Following a similar scheme, CeO2 quantum dots (QDs) intimately anchored on a g-C3N4 have been successfully utilized for a heterostructure with efficient photocatalytic performance for bacteria disinfection [129].
ZnS nanoparticles (10–15 nm) hybridized with g-C3N4 showed photocatalytic hydrogen evolution using glucose as the substrate. Interestingly, different efficiency has been observed depending on the synthesis method of Carbon nitride (melamine, urea, and dicyandiamide). The best yield was obtained from g-C3N4 produced from melamine [130].
A further application of g-C3N4 in the hybrid systems is provided by g-C3N4/Ag2CO3 heterostructures. The photocatalytic experiments indicated that the g-C3N4/Ag2CO3 photocatalyst exhibited significantly enhanced photocatalytic activity than the pure g-C3N4 and Ag2CO3 samples toward degrading methyl orange (MO) and catalytic ozonation for oxalic acid mineralization, under visible light irradiation (λ > 420nm) [131,132].
The ZnO/K-CN hybrid exhibits excellent photocatalytic degradation and adsorption efficiency of tetracycline under simulated sunlight, and the degradation rate reaches 90% due to the synergetic action of the formation of efficient type-II heterojunction between the K-doped exfoliated g-C3N4 nanosheet (K-CN) and the ZnO nanorod and the adsorption capacity of tetracycline through electrostatic force [133].
The ZnO/g-C3N4 hybrid system was successfully utilized as a sensor with high selectivity to dimethylamine (DMA), fast response time, and a fast recovery time. The heterojunction is effective for electron transfer and adsorption of oxygen molecules forming reactive oxygen species and thereby promoting the reaction between oxygen and DMA [134].
The use of metal complexes and metal nanoparticles (Au, Pt, Ag, etc.) in exfoliated carbon nitrides was utilized in the recent year for the successful application of photocatalysis. The nanosized metal particle will provide an extra absorption band related to localized surface plasmon resonance in the visible region and the possibility of injecting hot carriers into the reaction. Ag/C3N4 photocatalysts exhibited excellent stability and enhanced visible-light-driven photocatalytic performance both in the degradation of methyl orange (MO) and H2 evolution from water splitting [135]. Wang et al. found that under visible light illumination, the Au/g-C3N4 hybrids catalyst exhibits a superior oxygen evolution reaction performance in alkaline electrolytes. The high activity was attributed to three factors: the hot electrons injection of localized surface plasmon effect induced by light illumination and the formation of Schottky heterostructure [136]. The formation of heterojunctions between AgCl/ZnO and g-C3N4 significantly increases electron-hole transfer and separation compared to pure ZnO and g-C3N4. Thus, AgCl/ZnO/g-C3N4 exhibits increased photocatalytic activity in the removal of tetracycline hydrochloride, reaching 90 % under visible light irradiation. Hydrogen peroxide (H2O2) and superoxide radical (·O2) contributed more than holes (h+) and hydroxyl radicals (·OH) to the degradation process [134]. Zhang et al. observed that, contrary to pure g-C3N4, direct splitting of pure water is achieved with graphitic carbon nitride modified with Pt, PtOx, and CoOx as redox cocatalysts [137]. The photoactivity of Au/g-C3N4 nanocomposites showed higher efficiency in the degradation of both methylene blue dye and the gemifloxacin mesylate compared with that of bare g-C3N4. The efficiency was evaluated as a function of relative concentration between gold and g-C3N4, indicating the 1%Au/g-C3N4 nanocomposites as the best among the samples studied. Under visible excitation, it was able to reach more than 95% degradation of the target dye molecule in 90 min, in contrast to less than 70% obtained by pure g-C3N4 [138].
The structure of the hybrid photocatalysts can be arranged in more articulated paths, like a plasmonic ternary hybrid photocatalyst of Ag/AgBr/g-C3N4, synthesized and used for water splitting under visible light irradiation. Compared to a simple mixture of Ag/AgBr and g-C3N4, the ternary heterostructure showed a higher photoactivity. The performances are increased thanks to different factors: the decreasing of the recombination rate into the polymeric structure due to the high separation of the photoexcited charge, the low dimension of the Ag/AgBr nanoparticles, and the enhanced absorption related to the plasmonic resonance in the visible region [139]. Further, an efficient Ag/Ag2WO4/g-C3N4 ternary plasmonic photocatalyst was successfully synthesized using a facile one-step in situ hydrothermal methods. The strong coupling effect between the Ag/Ag2WO4 nanoparticles and the exfoliated g-C3N4 nanosheets gives excellent photocatalytic results (Rhodamine B and tetracycline degradation) under visible-light excitation thanks to the synergic action of the reduced the recombination rate of photogenerated electrons and holes and plasmonic absorption [140].
The metal can play the role of a cocatalyst, too, lowering the activation energy and increasing the rate of hydrogen production. Metals such as Pt, Au, Ag, Rh, and Ni are widely used as a cocatalyst to increase the efficiency of C3N4 for hydrogen evolution [141].
In Pt/g-C3N4-GO photocatalysts, the H2 evolution activity is strongly influenced by the temperature of the reduction stage, so the maximal rate of H2 evolution was obtained for the catalyst reduced at 400, and Pt concentration of 0.5%. The quantum efficiency was about 3% [142].
The photoactivity of Au/g-C3N4 nanocomposites has been evaluated for the degradation of both methylene blue dye and the drug Gemifloxacin mesylate, and their efficiencies were compared with that of bare g-C3N4. The efficiency increased by over 95% destruction of the target dye molecule in 90 min, in contrast to the 69% achieved with bare g-C3N4 [138].
Surface plasmon resonance plays a fundamental role in the visible photocatalytic property and photoinduced electron-hole pair separation efficiency. Ag-decorated g-C3N4 reports high photocatalytic degradation efficiency of methyl orange (98.7% within 2 h) and the high catalytic reduction property of 4-nitrophenol (100% within 70 s) [143].
Au/g-C3N4/NiFe2O4 exhibited a significant visible-light-driven photoactivity for hydrogen production. The high photocatalytic performances of the ternary hybrid nanocomposites are related to the high visible absorbance for charge carrier production, the fast separation of the electrons captured by AuNPs, and the strong absorption at the localized surface plasmon band in the Au nanoparticles that generate high-density carriers under visible light irradiation [144].
A new class of hybrid photocatalysts is constituted by metal-organic frameworks (MOF), generally constituted by porous crystalline materials with metallic nodes (metal ions or clusters) and organic linkers. MOFs have recently attracted increasing attention in the field of photocatalysis for their ultra-high specific surface area (over 6000 m2g−1), rich topology, and easily tunable porous structure [145].
In some MOFs, the metal clusters and organic linkers can be regarded as isolated semiconductor quantum dots, while the organic framer act as a light-absorbing antenna [146].
Structures such as MOF-5 (Zn4O(BDC)3, BDC: 1,4-benzenedicarboxylate), (Zr6O4(OH)4(BDC)6), and Ti8O8(OH)4(BDC)6, displayed semiconductor-like behavior. Herein, the metal-oxo clusters and organic linkers can be regarded as isolated semiconductor quantum dots and light-absorbing antennae, respectively.
In MOF structures, the HOMO/LUMO levels play similar roles as the conduction and valence band in classical semiconductors. The photogenerated electrons in the LUMO can be transferred to O2, leading to the formation of superoxide radicals (O2˙). Meanwhile, holes in the HOMO can oxidize the surface hydroxyl group/water, generating hydroxyl radicals (HO˙). The band gap of MOF can be modeled in a very wide range, from 3 to 4 eV MOF-5, UiO-66(Zr), and MIL-125(Ti) to a band gap lower than 2 eV in Fe- based MOF. In order to increase the photocatalytic efficiency in the visible region, different strategies have been proposed, mainly devoted to the generation, separation, and transfer of charge carriers [147]. The most studied technique to tune the optical properties is the modification of organic linkers rather than metallic modes. As an example, the use of –NH2 groups in H2BDC decreases the absorption edge down to 2.45 eV, whereas in pure H2BDC, the band gap is 3.6 eV, thereby extending the efficiency in the visible part of the spectrum [148].
Remarkably, the introduction of the –NH2 group leads to increased energy of the HOMO level with no influence on the LUMO. Similar effects have been studied for other functional groups where the –(NH2)2 group is the most active one: –CH3/–Cl < –OH < –(NH2)2 [149]. Similarly, it is possible to use the Covalent organic framework (COF) to obtain efficient visible photocatalysis. In MOF/COF heterostructured photocatalyst (Cu-NH2-MIL-125/TpPa-2-COF), with Cu ions immobilized by -NH2 groups of MOF, while the covalently connected heterojunction facilitates the separation of photogenerated charges. The results show an efficient “amines to imines” process with a high conversion rate (91.2%) and selectivity (>99%) [150].
In addition to functional group modification, the implantation of transition metal ions to complexes with ligands was reported to be a feasible way to enhance photocatalytic performance [151]. The implantation of Fe3+ in porphyrinic MOFs (PCN-224) generates an extension of the optical response to a longer wavelength inhibiting the recombination of electron-hole pairs. Lou et al. report sheet-like sliver vanadate/Bi–Fe MOF (AGV/HBF) heterojunction composite with superior efficiency under visible light irradiation (almost 99% in 2 h for the photocatalytic decomposition of Rhodamine B). The enhanced photocatalytic performance was attributed to the strong visible-light absorption and low recombination of electron-hole pairs due to the as-formed heterojunctions [152].
The dye-sensitizing approach is applied even to the MOF structure. The high specific area and the high number of bonding sites can facilitate a strong interaction and an efficient antenna effect of the dyes. The dye/MOF structure has been applied to H2 production and CO2 reduction, and environmental remediation [152].
Li et al. report porous copper-based metal–organic framework (Cu-MOF) tailored by encapsulating visible-light photosensitive dyes. The study underlying the high-efficiency and diversified performances of the different systems in the photodegradation process reveals significant achievements in tailoring the photocatalytic capabilities via introducing dyes with different photosensitivity into an identical porous copper-organic framework [153].
The coupling of MOFs with other photoactive semiconductors represents an additional strategy to enhance photocatalytic activity. Again, the high specific area of the MOF structure facilitates this approach. Three types of heterojunctions for semiconductors can be identified, depending on the relative position of the Conduction and Valence band as well as the n/p characters of the independent components, namely straddling gap; type-II: staggered gap; type-III: broken gap [101,154]. Until now, many semiconductors have been reported to form composites with MOFs. Metal-containing semiconductors (such as ZnO [155], TiO2 [156], BiVO4 [157], and CdS [158]) and non-metal graphitic carbon nitride [159] have been reported to form composites with MOFs exhibiting highly potential performance in the field of photocatalysis.
Doustkhah et al. utilized the highly porous structure of the MOF (MOF-5) as a precursor to obtaining superior photocatalytic activity in degrading methylene blue (MB) in comparison to other ZnO nanostructures [155]. Similarly, CdS/Cd-MOF nanocomposites were constructed by in situ sulfurizations to form CdS using Cd-MOF as a precursor. The role of the high specific surface is evidenced by the high degradation rate of MB (91.9%) in comparison with the pure Cd-MOF (62.3%) and pure CdS (67.5%) [158].

6. Organic/Inorganic Heterostructures: Applications

Organic-Inorganic hybrids have been widely studied as efficient photocatalysts, degradation of organic pollutants, hydrogen evolution, CO2 reduction, sterilization, and reduction of heavy metals are the most studied. In Table 1 are collected some examples of hybrid photocatalysts for each application.

6.1. Environmental Applications: Pollutants Degradation, Reduction of Heavy Metals, Sterilization

In the field of the degradation of pollutants, Han et al. [161] describe how the π-π stacking in perylene imide/Bi2WO6 hybrid photocatalyst enhances the photocatalytic activity for the degradation of Bisphenol A (BPA), also underlining the good stability and recyclability. The 2D sheet-like structure observed in the SEM measurements promotes the presence of a high concentration of active sites on the surface. The presence of the polymer facilitates the separation of the charge carriers, increasing the photocatalytic activity joined to the nanosheet-like structure that increases the number of active sites for the reaction (Figure 11). The photocatalytic tests show that by increasing the percentage of Bi2WO3 in the hybris first, the photocatalytic activity increases and then decreases, and the best activity has been observed when 30% of Bi2WO3 is present, and it is stable even after five cycles (Figure 12).
The modifications through the modulation of the organic mesh represent a further strategy to increase the photocatalytic activity of the material. The introduction of phenyl rings in the mesh of g-C3N4 was utilized to shift the absorption edge in the visible range and to obtain an efficient hybrid photocatalyst (Phenyl modified-TiO2 hereafter PhCN) up to 600 nm (Figure 13) [96,166]. The absorption spectrum of the hybrid system appears as the sum of the two building blocks.
The phenyl-modified carbon nitride presents lower band gap LUMO levels almost resonant with the conduction band of TiO2; the electrons are promoted from the HOMO ground state into the LUMO excited state of the phenyl carbon nitride and then transferred to the conduction band of the TiO2. The degradation efficiency for Rhodamine B and Methylene Blue under white LED irradiation increased from 4% by using pure compounds to 17% by using only g-C3N4/TiO2, raising to 98% and 88% with the hybrid system PhCN/TiO2 (Figure 14).
New methods for killing antibiotic-resistant bacteria have gained large interest in the last year, and in this field, photocatalysis is widely used. The hybrid photocatalyst studied by Yang et al. [180] appears to be very efficient in capturing bacteria. TEM images of the Sn3O4/PDINH hybrid show an organic amorphous layer with a thickness of about 5nm on the surface of Sn3O4 nanosheets. The formation of the composite has been further confirmed by EDS analysis, which reveals the presence of Sn, O, C, and N (Figure 15). This hybrid photocatalyst obtained by coupling Sn3O4 and perylene diimide, thanks to the π-π bonding interactions, allows generating active oxygen species capable of killing the bacteria because these species interact with proteins, lipids, and other components of the bacteria, causing death. The photocatalytic antibacterial activity has been evaluated by performing the plate-counting method. The highest observed capture rate for this hybrid has been 82%, improved by the presence of the organic polymer on the surface (Figure 16).
Shen et al. synthesized novel three-dimensional Bi2S3 nanocrystals capped by polyvinyl pyrrolidone photocatalyst for the reduction of hexavalent chromium under visible light irradiation. TEM images display flower-like structures consisting of nanosheets (Figure 17). The reduction of metals with such an impressive photocatalyst reaches 95% in only 5 min of reaction (Figure 18).

6.2. Energy: CO2 Reduction and H2 Evolution

The formation of two-dimensional (2D) nanostructures of the polymeric structure is a successful strategy to increase the photocatalytic activity: higher aspect ratio, larger surface area, and increased number of surface groups for anchoring cocatalysts are fundamental properties of the layered structures. Two-dimensional g-C3N4 nanosheets have been successfully obtained, and significantly photocatalytic performance was reported [97,133].
For CO2 reduction, a new and very stable 0D/2D hybrid system has been proposed by Gong et al. [173]. This photocatalyst, with a high surface area, presents high photocatalytic activity because of the presence of a large number of active sites. In SEM images, the hybrids systems show the nanosheet structure and no adhesive nanoparticles on the surface (Figure 19). The observed reduction rate for the CO2 reduction has been 69.8 μmol g−1g−1 under visible light irradiation (Figure 20).
The production of H2 as an energy source has been studied increasingly in the last few years. Hamdy et al. [168] propose a novel polyaniline/ZnO heterojunction, synthesized via sol-gel/polymerization route for photocatalytic hydrogen production under a visible light source.
The amount of polymer has a fundamental role in photocatalytic activity as it plays a uniform deposition of it on the ZnO surface (Figure 21).
Zang et al. showed how the p-conjugation, the flexibility, and the high surface area of pyrene-based polymer coupled with MoS2 enhance photocatalytic H2 production [170]. Under visible light irradiation, the electrons can be quickly transferred from the conduction band of pyrene-based polymer to MoS2 restraining the charge recombination process (Figure 22).
From SEM and TEM measurements, it is possible to observe the different morphology of the two components of the hybrid system. The PyP particle size is 40–100 nm while MoS2 shows a nanosheet-like structure with few nanometers of thickness, as it is well distinguishable in the TEM images (Figure 23). The photocatalytic activity is enhanced by the fast transfer of the excited electrons from the conduction band of the PyP to the MoS2, whose surface is covered with active sites, allowing the photocatalytic process.
Wang et al. designed TiO2-modified g-C3N4 composite photocatalysts using the ball milling synthesis method followed by a calcination process. From spectroscopic measurements and morphological studies, they confirmed the formation of the heterostructure that reflects an enhancement of the separation rate of the photogenerated electrons and holes and, therefore, of the CO2 reduction rate (Figure 24) [177].
From SEM analysis emerges an irregular structure with aggregates of bulk g-C3N4 and TiO2 nanoparticles with 20–50 nm size on its multilayered surface (Figure 25). The formation of the heterostructure promotes the separation between the electron and the holes, and the best photocatalytic result shows the yields of 72.2 and 56.2 μmol g−1 for CH4 and CO, respectively
In summary, hybrid organic/inorganic photocatalysts have emerged as highly potential materials to drastically improve visible photocatalytic efficiency. Our review cannot be fully comprehensive for the large amount of new scientific products that are available daily, but it would represent a guide for readers that want to enter more insights into the field of photocatalysis. More specifically, we presented the high potential impact of the use of visible light for environmental remediation and energy production.

7. Conclusions and Outlook

“Set the controls for the heart of the sun”, sang Pink Floyd; we can now better paraphrase the title as “Set the control for the use of the heart of the sun”. The visible-driven photocatalytic process has become a rapid and important strategy to mitigate the anthropogenic impact and mitigate energy consumption. Even if different solutions have already been proposed and, in some cases, good performances have been achieved, the road is still long. Starting from the wide galaxy of the actual literature, it will be possible to set the control of the best parameters to use solar radiation in the photocatalytic properties.
In this review, we analyzed the state art inorganic and organic semiconductors underlining how the improvement of the absorption in the visible range represents a crucial point in the actual research and how the study of heterostructures can overcome this problem.
New Materials with reduced environmental impact, the use of hybrid solutions to overcome the limitations of pure inorganic and organic materials, surface modification, doping processes, and new heterostructures are important parameters to set the more appropriate energy level and recombination process.
The review aims to provide a useful guideline for designing and developing new photocatalysts useful to find suitable roads to achieve the use of the Heart of the Sun.

Author Contributions

Conceptualization, S.P., P.C.R. and F.S.; writing—original draft preparation, S.P. and P.C.R.; writing—review and editing, S.P. and P.C.R.; supervision, S.P., P.C.R. and F.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the “Fondazione di Sardegna” within the project F75F21001230007 “Paint4Vid—Visible photocatalytic materials for viruses and bacteria inactivation”.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Manisalidis, I.; Stavropoulou, E.; Stavropoulos, A.; Bezirtzoglou, E. Environmental and Health Impacts of Air Pollution: A Review. Front. Public Health 2020, 8, 14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Delbeke, J.; Runge-Metzger, A.; Slingenberg, Y.; Werksman, J. The paris agreement. Towards A Clim.-Neutral Eur. Curbing Trend 2019, 22, 24–45. [Google Scholar] [CrossRef]
  3. Yadav, S.; Shakya, K.; Gupta, A.; Singh, D.; Chandran, A.R.; Aanappalli, A.V.; Goyal, K.; Rani, N.; Saini, K. A review on degradation of organic dyes by using metal oxide semiconductors. Environ. Sci. Pollut. Res. 2022. [Google Scholar] [CrossRef] [PubMed]
  4. Shindhal, T.; Rakholiya, P.; Varjani, S.; Pandey, A.; Ngo, H.H.; Guo, W.; Ng, H.Y.; Taherzadeh, M.J. A critical review on advances in the practices and perspectives for the treatment of dye industry wastewater. Bioengineered 2021, 12, 70–87. [Google Scholar] [CrossRef] [PubMed]
  5. Zhong, N.; Yuan, J.; Luo, Y.; Zhao, M.; Luo, B.; Liao, Q.; Chang, H.; Zhong, D.; Rittmann, B.E. Intimately coupling photocatalysis with phenolics biodegradation and photosynthesis. Chem. Eng. J. 2021, 425, 130666. [Google Scholar] [CrossRef]
  6. Yadav, G.; Ahmaruzzaman, M. New generation advanced nanomaterials for photocatalytic abatement of phenolic compounds. Chemosphere 2022, 304, 135297. [Google Scholar] [CrossRef] [PubMed]
  7. Lotfi, H.; Heydarinasab, A.; Mansouri, M.; Hosseini, S.H. Kinetic modeling of removal of aromatic hydrocarbons from petroleum wastewaters by UiO-66-NH2/TiO2/ZnO nanocomposite. J. Environ. Chem. Eng. 2022, 10, 107066. [Google Scholar] [CrossRef]
  8. Pica, M. Treatment of wastewaters with zirconium phosphate based materials: A review on efficient systems for the removal of heavy metal and dye water pollutants. Molecules 2021, 26, 2392. [Google Scholar] [CrossRef] [PubMed]
  9. Qasem, N.A.A.; Mohammed, R.H.; Lawal, D.U. Removal of heavy metal ions from wastewater: A comprehensive and critical review. Npj Clean Water 2021, 4, 36. [Google Scholar] [CrossRef]
  10. Nguyen, D.N.; Bui, H.M.; Nguyen, H.Q. Heterogeneous Photocatalysis for the Removal of Pharmaceutical Compounds; Elsevier B.V.: Amsterdam, The Netherlands, 2020; ISBN 9780128195949. [Google Scholar]
  11. Yang, X.; Chen, Z.; Zhao, W.; Liu, C.; Qian, X.; Zhang, M.; Wei, G.; Khan, E.; Hau Ng, Y.; Sik Ok, Y. Recent advances in photodegradation of antibiotic residues in water. Chem. Eng. J. 2021, 405, 126806. [Google Scholar] [CrossRef]
  12. Fujishima, A.; Honda, K. Electrochemical Photolysis of Water at a Semiconductor Electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef] [PubMed]
  13. Fernandes, S.; de Silva, J.C.Q.E.; da SIlva, L.P. Life Cycle Assessment of the Sustainability of Enhancing the Photodegradation Activity of TiO2 with Metal-Doping. Materials 2020, 13, 1487. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Fernández-ibáñez, P. Solar-driven Photocatalyzed Hydrogen Production from Water cleavage: Evaluation and Process Life Cycle Assessment. J. Phys. Conf. Ser. 2022, 2305, 012040. [Google Scholar] [CrossRef]
  15. Wang, H.; Li, X.; Zhao, X.; Li, C.; Song, X.; Zhang, P.; Huo, P. A review on heterogeneous photocatalysis for environmental remediation: From semiconductors to modification strategies. Chin. J. Catal. 2022, 43, 178–214. [Google Scholar] [CrossRef]
  16. Kumar, A.; Choudhary, P.; Kumar, A.; Camargo, P.H.C.; Krishnan, V. Recent Advances in Plasmonic Photocatalysis Based on TiO2 and Noble Metal Nanoparticles for Energy Conversion, Environmental Remediation, and Organic Synthesis. Small 2022, 18, e2101638. [Google Scholar] [CrossRef]
  17. Candish, L.; Collins, K.D.; Cook, G.C.; Douglas, J.J.; Gómez-Suárez, A.; Jolit, A.; Keess, S. Photocatalysis in the Life Science Industry. Chem. Rev. 2022, 122, 2907–2980. [Google Scholar] [CrossRef] [PubMed]
  18. Nakata, K.; Fujishima, A. TiO2 photocatalysis: Design and applications. J. Photochem. Photobiol. C Photochem. Rev. 2012, 13, 169–189. [Google Scholar] [CrossRef]
  19. Yao, G.; Liu, Y.; Liu, J.; Xu, Y. Facile Synthesis of Porous g-C3N4 with Enhanced Visible-Light Photoactivity. Molecules 2022, 27, 1754. [Google Scholar] [CrossRef]
  20. Shah, S.A.; Khan, I.; Yuan, A. MoS2 as a Co-Catalyst for Photocatalytic Hydrogen Production: A Mini Review. Molecules 2022, 27, 3289. [Google Scholar] [CrossRef]
  21. Hai, G.; Wang, H. Theoretical studies of metal-organic frameworks: Calculation methods and applications in catalysis, gas separation, and energy storage. Coord. Chem. Rev. 2022, 469, 214670. [Google Scholar] [CrossRef]
  22. Swain, S.; Altaee, A.; Saxena, M.; Samal, A.K. A comprehensive study on heterogeneous single atom catalysis: Current progress, and challenges q. Coord. Chem. Rev. 2022, 470, 214710. [Google Scholar] [CrossRef]
  23. Boumeriame, H.; Da Silva, E.S.; Cherevan, A.S.; Chafik, T.; Faria, J.L.; Eder, D. Layered double hydroxide (LDH)-based materials: A mini-review on strategies to improve the performance for photocatalytic water splitting. J. Energy Chem. 2021, 64, 406–431. [Google Scholar] [CrossRef]
  24. Zhao, J.; Ren, J.; Zhang, G.; Zhao, Z.; Liu, S.; Zhang, W.; Chen, L. Donor-Acceptor Type Covalent Organic Frameworks. Chem. Eur. J. 2021, 27, 10781–107977. [Google Scholar] [CrossRef]
  25. Chen, J.; Abazari, R.; Adegoke, K.A.; Maxakato, N.W.; Bello, O.S.; Tahir, M.; Tasleem, S.; Sanati, S.; Kirillov, A.M.; Zhou, Y. Metal–organic frameworks and derived materials as photocatalysts for water splitting and carbon dioxide reduction. Coord. Chem. Rev. 2022, 469, 214664. [Google Scholar] [CrossRef]
  26. Zhou, Y.; Lu, F.; Fang, T.; Gu, D.; Feng, X.; Song, T.; Liu, W. A brief review on metal halide perovskite photocatalysts: History, applications and prospects. J. Alloys Compd. 2022, 911, 165062. [Google Scholar] [CrossRef]
  27. Al-Azri, Z.H.N.; Aloufi, M.; Chan, A.; Waterhouse, G.I.N.; Idriss, H. Metal Particle Size Effects on the Photocatalytic Hydrogen Ion Reduction. ACS Catal. 2019, 9, 3946–3958. [Google Scholar] [CrossRef]
  28. Kuznetsov, A.E. Review of research of nanocomposites based on graphene quantum dots. Comput. Chem. Appl. New Technol. 2021, 135–160. [Google Scholar] [CrossRef]
  29. Humayun, M.; Wang, C.; Luo, W. Recent Progress in the Synthesis and Applications of Composite Photocatalysts: A Critical Review. Small Methods 2022, 6, 2101395. [Google Scholar] [CrossRef]
  30. Wang, C.Y.; Maeda, K.; Chang, L.L.; Tung, K.L.; Hu, C. Synthesis and applications of carbon nitride (CNx) family with different carbon to nitrogen ratio. Carbon 2022, 188, 482–491. [Google Scholar] [CrossRef]
  31. Jedla, M.R.; Koneru, B.; Franco, A.; Rangappa, D.; Banerjee, P. Recent developments in nanomaterials based adsorbents for water purification techniques. Biointerface Res. Appl. Chem. 2022, 12, 5821–5835. [Google Scholar] [CrossRef]
  32. Almaie, S.; Vatanpour, V.; Hossein, M. Chemosphere Volatile organic compounds (VOCs) removal by photocatalysts: A review. Chemosphere 2022, 306, 135655. [Google Scholar] [CrossRef]
  33. Castilla-Caballero, D.; Sadak, O.; Martínez-Díaz, J.; Martínez-Castro, V.; Colina-Márquez, J.; Machuca-Martínez, F.; Hernandez-Ramirez, A.; Vazquez-Rodriguez, S.; Gunasekaran, S. Solid-state photocatalysis for plastics abatement: A review. Mater. Sci. Semicond. Process. 2022, 149, 106890. [Google Scholar] [CrossRef]
  34. Hazaraimi, M.H.; Goh, P.S.; Lau, W.J.; Ismail, A.F.; Wu, Z.; Subramaniam, M.N.; Lim, J.W.; Kanakaraju, D. The state-of-the-art development of photocatalysts for degrading of persistent herbicides in aqueous environment. Sci. Total Environ. 2022, 843, 156975. [Google Scholar] [CrossRef]
  35. Wang, Y.; Torres, J.A.; Shviro, M.; Carmo, M.; He, T.; Ribeiro, C. Photocatalytic materials applications for sustainable agriculture. Prog. Mater. Sci. 2022, 130, 100965. [Google Scholar] [CrossRef]
  36. Rajeshwar, K.; Osugi, M.E.; Chanmanee, W.; Chenthamarakshan, C.R.; Zanoni, M.V.B.; Kajitvichyanukul, P.; Krishnan-Ayer, R. Heterogeneous photocatalytic treatment of organic dyes in air and aqueous media. J. Photochem. Photobiol. C Photochem. Rev. 2008, 9, 171–192. [Google Scholar] [CrossRef]
  37. Rehman, S.; Ullah, R.; Butt, A.M.; Gohar, N.D. Strategies of making TiO2 and ZnO visible light active. J. Hazard. Mater. 2009, 170, 560–569. [Google Scholar] [CrossRef]
  38. Aslam, M.; Fazal, D.B.; Ahmad, F.; Fazal, A.B.; Abdullah, A.Z.; Ahmed, M.; Qamar, M.; Rafatullah, M. Photocatalytic Degradation of Recalcitrant Pollutants of Greywater. Catalysts 2022, 12, 557. [Google Scholar] [CrossRef]
  39. Ren, Y.; Yu, C.; Tan, X.; Wei, Q.; Wang, Z.; Ni, L.; Wang, L.; Qiu, J. Strategies to activate inert nitrogen molecules for efficient ammonia electrosynthesis: Current status, challenges, and perspectives. Energy Environ. Sci. 2022, 15, 2776–2805. [Google Scholar] [CrossRef]
  40. Saravanan, R.; Gracia, F.; Stephen, A. Basic Principles, Mechanism, and Challenges of Photocatalysis. In Nanocomposites for Visible Light-Induced Photocatalysis; Springer Series on Polymer and Composite Materials; Khan, M., Pradhan, D., Sohn, Y., Eds.; Springer: Berlin, Germany, 2017; pp. 19–40. [Google Scholar]
  41. Cernuto, G.; Masciocchi, N.; Cervellino, A.; Colonna, G.M.; Guagliardi, A. Size and shape dependence of the photocatalytic activity of TiO2 nanocrystals: A total scattering Debye function study. J. Am. Chem. Soc. 2011, 133, 3114–3119. [Google Scholar] [CrossRef] [PubMed]
  42. Zhang, M.; Salvador, P.A.; Rohrer, G.S. Influence of particle size and shape on the rate of hydrogen produced by Al-doped SrTiO3 photocatalysts. J. Am. Ceram. Soc. 2022, 108, 5336–5346. [Google Scholar] [CrossRef]
  43. Leng, D.; Wang, T.; Du, C.; Pei, X.; Wan, Y.; Wang, J. Synthesis of β-Bi2O3 nanoparticles via the oxidation of Bi nanoparticles: Size, shape and polymorph control, anisotropic thermal expansion, and visible-light photocatalytic activity. Ceram. Int. 2022, 48, 18270–18277. [Google Scholar] [CrossRef]
  44. Zheng, Y.; Duan, Z.; Liang, R.; Lv, R.; Wang, C.; Zhang, Z.; Wan, S.; Wang, S.; Xiong, H.; Ngaw, C.K.; et al. Shape-Dependent Performance of Cu/Cu2O for Photocatalytic Reduction of CO2. ChemSusChem 2022, 15, e202200216. [Google Scholar] [CrossRef]
  45. Khan, M.M.; Adil, S.F.; Al-Mayouf, A. Metal oxides as photocatalysts. J. Saudi Chem. Soc. 2015, 19, 462–464. [Google Scholar] [CrossRef] [Green Version]
  46. Saravanan, R.; Gupta, V.K.; Narayanan, V.; Stephen, A. Comparative study on photocatalytic activity of ZnO prepared by different methods. J. Mol. Liq. 2013, 181, 133–141. [Google Scholar] [CrossRef]
  47. He, X.; Yang, Y.; Li, Y.; Chen, J.; Yang, S.; Liu, R.; Xu, Z. Effects of structure and surface properties on the performance of ZnO towards photocatalytic degradation of methylene blue. Appl. Surf. Sci. 2022, 599, 153898. [Google Scholar] [CrossRef]
  48. Chatterjee, D.; Dasgupta, S. Visible light induced photocatalytic degradation of organic pollutants. J. Photochem. Photobiol. C Photochem. Rev. 2005, 6, 186–205. [Google Scholar] [CrossRef]
  49. Castillo-Ledezma, J.H.; Salas, J.L.S.; López-Malo, A.; Bandala, E.R. Effect of pH, solar irradiation, and semiconductor concentration on the photocatalytic disinfection of Escherichia coli in water using nitrogen-doped TiO2. Eur. Food Res. Technol. 2011, 233, 825–834. [Google Scholar] [CrossRef]
  50. Abbasi, S.; Hasanpour, M. The effect of pH on the photocatalytic degradation of methyl orange using decorated ZnO nanoparticles with SnO2 nanoparticles. J. Mater. Sci. Mater. Electron. 2017, 28, 1307–1314. [Google Scholar] [CrossRef]
  51. Reza, K.M.; Kurny, A.; Gulshan, F. Parameters affecting the photocatalytic degradation of dyes using TiO2: A review. Appl. Water Sci. 2017, 7, 1569–1578. [Google Scholar] [CrossRef] [Green Version]
  52. Malato, S.; Fernández-Ibáñez, P.; Maldonado, M.I.; Blanco, J.; Gernjak, W. Decontamination and disinfection of water by solar photocatalysis: Recent overview and trends. Catal. Today 2009, 147, 1–59. [Google Scholar] [CrossRef]
  53. Chen, L.-l.; Zhai, B.-g.; Huang, Y.M. Rendering Visible-Light Photocatalytic Activity to Undoped ZnO via Intrinsic Defects. Eng. Catal. 2020, 10, 1163. [Google Scholar] [CrossRef]
  54. Zhou, H.; Qu, Y.; Zeid, T.; Duan, X. Towards highly efficient photocatalysts using semiconductor nanoarchitectures. Energy Environ. Sci. 2012, 5, 6732–6743. [Google Scholar] [CrossRef]
  55. Solarska, R.; Królikowska, A.; Augustyński, J. Silver Nanoparticle Induced Photocurrent Enhancement at WO3 Photoanodes. Angew. Chem. Int. Ed. 2010, 49, 7980–7983. [Google Scholar] [CrossRef] [PubMed]
  56. Palmas, S.; Castresana, P.A.; Mais, L.; Vacca, A.; Mascia, M.; Ricci, P.C. TiO2-WO3 nanostructured systems for photoelectrochemical applications. RSC Adv. 2016, 6, 101671–101682. [Google Scholar] [CrossRef] [Green Version]
  57. Kay, A.; Cesar, I.; Grätzel, M. New benchmark for water photooxidation by nanostructured α-Fe2O3 films. J. Am. Chem. Soc. 2006, 128, 15714–15721. [Google Scholar] [CrossRef]
  58. Ma, D.; Li, J.; Liu, A.; Chen, C. Carbon Gels-Modified TiO2: Promising Materials for Photocatalysis Applications. Materials 2020, 13, 1734. [Google Scholar] [CrossRef] [Green Version]
  59. Frank, S.N.; Bard, A.J. Heterogeneous Photocatalytic Oxidation of Cyanide Ion in Aqueous Solutions at T1O2 Powder. J. Am. Chem. Soc. 1977, 99, 303–304. [Google Scholar] [CrossRef]
  60. Li, Y.; Ren, Z.; Gu, M.; Duan, Y.; Zhang, W.; Lv, K. Applied Catalysis B: Environmental Synergistic effect of interstitial C doping and oxygen vacancies on the photoreactivity of TiO2 nanofibers towards CO2 reduction. Appl. Catal. B Environ. 2022, 317, 121773. [Google Scholar] [CrossRef]
  61. Nor, N.U.M.; Mazalan, E.; Risko, C.; Crocker, M.; Amin, N.A.S. Unveiling the structural, electronic, and optical effects of carbon-doping on multi-layer anatase TiO2 (1 0 1) and the impact on photocatalysis. Appl. Surf. Sci. 2022, 586, 152641. [Google Scholar] [CrossRef]
  62. Amor, C.O.; Elghniji, K.; Elaloui, E. Improving charge separation, photocurrent and photocatalytic activities of Dy-doped TiO2 by surface modification with salicylic acid. J. Mater. Sci. Mater. Electron. 2020, 31, 20919–20931. [Google Scholar] [CrossRef]
  63. Guetni, I.; Belaiche, M.; Ferdi, C.A.; Oulhakem, O.; Alaoui, K.B.; Naimi, Z. Engineering the photocatalytic efficiency of nanoscale TiO2 anatase with doped (Nd/Y) and co-doped (Nd-Y/Nd-Sm) lanthanides to decompose the azo dye orange G under UV-VIS irradiation. New J. Chem. 2022, 46, 10162–10183. [Google Scholar] [CrossRef]
  64. Khan, R.; Rahman, A.U. Band-Gap Narrowing of Titanium Dioxide by Nitrogen Doping. Jpn. J. Appl. Phys. 2001, 40, 561–563. [Google Scholar] [CrossRef]
  65. Irie, H.; Watanabe, Y.; Hashimoto, K. Nitrogen-concentration dependence on photocatalytic activity of TiO2-xNx powders. J. Phys. Chem. B 2003, 107, 5483–5486. [Google Scholar] [CrossRef]
  66. Ricci, P.C.; Casu, A.; Salis, M.; Corpino, R.; Anedda, A. Optically controlled phase variation of TiO2 nanoparticles. J. Phys. Chem. C 2010, 114, 14441–14445. [Google Scholar] [CrossRef]
  67. Chandrasekharan, N.; Kainat, P.Y. Improving the photoelectrochemical performance of nanostructured TiO2 films by adsorption of gold nanoparticles. J. Phys. Chem. B 2000, 104, 10851–10857. [Google Scholar] [CrossRef]
  68. Subramanian, V.; Wolf, E.E.; Kamat, P.V. Catalysis with TiO2/Gold Nanocomposites. Effect of Metal Particle Size on the Fermi Level Equilibration. J. Am. Chem. Soc. 2004, 126, 4943–4950. [Google Scholar] [CrossRef]
  69. Wu, F.; Zhou, Z.; Hicks, A. Life Cycle Impact of Titanium Dioxide Nanoparticle Synthesis through Physical, Chemical, and Biological Routes. Environ. Sci. Technol. 2019, 53, 4078–4087. [Google Scholar] [CrossRef]
  70. Lv, S.; Du, Y.; Wu, F.; Cai, Y.; Zhou, T. Review on LSPR assisted photocatalysis: Effects of physical fields and opportunities in multifield decoupling. Nanoscale Adv. 2022, 4, 2608–2631. [Google Scholar] [CrossRef]
  71. Rycenga, M.; Cobley, C.M.; Zeng, J.; Li, W.; Moran, C.H.; Zhang, Q.; Qin, D.; Xia, Y. Controlling the synthesis and assembly of silver nanostructures for plasmonic applications. Chem. Rev. 2011, 111, 3669–3712. [Google Scholar] [CrossRef] [Green Version]
  72. Aoki, Y.; Ishida, T.; Tatsuma, T. Plasmon-Induced Photocatalysis Based on Pt-Au Coupling with Enhanced Oxidation Abilities. ACS Appl. Nano Mater. 2022, 5, 4406–4412. [Google Scholar] [CrossRef]
  73. Wang, H.; Wang, Y.; Guo, L.; Zhang, X.; Ribeiro, C.; He, T. Solar-heating boosted catalytic reduction of CO2 under full-solar spectrum. Chin. J. Catal. 2020, 41, 131–139. [Google Scholar] [CrossRef]
  74. Liu, T.; Besteiro, L.V.; Liedl, T.; Correa-Duarte, M.A.; Wang, Z.; Govorov, A.O. Chiral Plasmonic Nanocrystals for Generation of Hot Electrons: Toward Polarization-Sensitive Photochemistry. Nano Lett. 2019, 19, 1395–1407. [Google Scholar] [CrossRef] [Green Version]
  75. Golze, S.D.; Porcu, S.; Zhu, C.; Sutter, E.; Ricci, P.C.; Kinzel, E.C.; Hughes, R.A.; Neretina, S. Sequential Symmetry-Breaking Events as a Synthetic Pathway for Chiral Gold Nanostructures with Spiral Geometries. Nano Lett. 2021, 21, 2919–2925. [Google Scholar] [CrossRef]
  76. Dong, J.; Yan, C.; Chen, Y.; Zhou, W.; Peng, Y.; Zhang, Y.; Wang, L.N.; Huang, Z.H. Organic semiconductor nanostructures: Optoelectronic properties, modification strategies, and photocatalytic applications. J. Mater. Sci. Technol. 2022, 113, 175–198. [Google Scholar] [CrossRef]
  77. Holmes, N.P.; Chambon, S.; Holmes, A.; Xu, X.; Hirakawa, K.; Deniau, E.; Lartigau-Dagron, C.; Bousquet, A. Organic semiconductor colloids: From the knowledge acquired in photovoltaics to the generation of solar hydrogen fuel. Curr. Opin. Colloid Interface Sci. 2021, 56, 101511. [Google Scholar] [CrossRef]
  78. Viola, F.A.; Spanu, A.; Ricci, P.C.; Bonfiglio, A.; Cosseddu, P. Ultrathin, flexible and multimodal tactile sensors based on organic field-effect transistors. Sci. Rep. 2018, 8, 1–8. [Google Scholar] [CrossRef] [Green Version]
  79. Cui, Y.; Yao, H.; Zhang, J.; Zhang, T.; Wang, Y.; Hong, L.; Xian, K.; Xu, B.; Zhang, S.; Peng, J.; et al. Over 16% efficiency organic photovoltaic cells enabled by a chlorinated acceptor with increased open-circuit voltages. Nat. Commun. 2019, 10, 2515. [Google Scholar] [CrossRef] [Green Version]
  80. Mazzanti, M.; Milani, M.; Cristino, V.; Boaretto, R.; Molinari, A.; Caramori, S. Visible Light Reductive Photocatalysis of Azo-Dyes with n–n Junctions Based on Chemically Deposited CdS. Molecules 2022, 27, 2924. [Google Scholar] [CrossRef]
  81. Fu, Z.; Wang, X.; Gardner, A.M.; Wang, X.; Chong, S.Y.; Neri, G.; Cowan, A.J.; Liu, L.; Li, X.; Vogel, A.; et al. A stable covalent organic framework for photocatalytic carbon dioxide reduction. Chem. Sci. 2020, 11, 543–550. [Google Scholar] [CrossRef] [Green Version]
  82. Qian, Y.; Ma, D. Covalent organic frameworks: New materials platform for photocatalytic degradation of aqueous pollutants. Materials 2021, 14, 5600. [Google Scholar] [CrossRef]
  83. Kroke, E.; Schwarz, M.; Horath-Bordon, E.; Kroll, P.; Noll, B.; Norman, A.D. Tri-s-triazine derivatives. Part I. From trichloro-tri-s-triazine to graphitic C3N4 structures. New J. Chem. 2002, 26, 508–512. [Google Scholar] [CrossRef]
  84. Li, Y.Y.; Ma, S.F.; Zhou, B.X.; Huang, W.Q.; Fan, X.; Li, X.; Li, K.; Huang, G.F. Hydroxy-carbonate-assisted synthesis of high porous graphitic carbon nitride with broken of hydrogen bonds as a highly efficient visible-light-driven photocatalyst. J. Phys. D Appl. Phys. 2019, 52, 105502. [Google Scholar] [CrossRef]
  85. Li, Y.; Ruan, Z.; He, Y.; Li, J.; Li, K.; Jiang, Y.; Xu, X.; Yuan, Y.; Lin, K. In situfabrication of hierarchically porous g-C3N4 and understanding on its enhanced photocatalytic activity based on energy absorption. Appl. Catal. B Environ. 2018, 236, 64–75. [Google Scholar] [CrossRef]
  86. Mousavi, M.; Habibi-Yangjeh, A.; Pouran, S.R. Review on magnetically separable graphitic carbon nitride-based nanocomposites as promising visible-light-driven photocatalysts. J. Mater. Sci. Mater. Electron. 2018, 29, 1719–1747. [Google Scholar] [CrossRef]
  87. Vaya, D.; Kaushik, B.; Surolia, P.K. Recent advances in graphitic carbon nitride semiconductor: Structure, synthesis and applications. Mater. Sci. Semicond. Process. 2022, 137, 106181. [Google Scholar] [CrossRef]
  88. Huang, D.; Li, Z.; Zeng, G.; Zhou, C.; Xue, W.; Gong, X.; Yan, X.; Chen, S.; Wang, W.; Cheng, M. Megamerger in photocatalytic field: 2D g-C3N4 nanosheets serve as support of 0D nanomaterials for improving photocatalytic performance. Appl. Catal. B Environ. 2019, 240, 153–173. [Google Scholar] [CrossRef]
  89. Zheng, Y.; Lin, L.; Wang, B.; Wang, X. Graphitic Carbon Nitride Polymers toward Sustainable Photoredox Catalysis. Angew. Chem.-Int. Ed. 2015, 54, 12868–12884. [Google Scholar] [CrossRef]
  90. Ong, W.J.; Putri, L.K.; Tan, Y.C.; Tan, L.L.; Li, N.; Ng, Y.H.; Wen, X.; Chai, S.P. Unravelling charge carrier dynamics in protonated g-C3N4 interfaced with carbon nanodots as co-catalysts toward enhanced photocatalytic CO2 reduction: A combined experimental and first-principles DFT study. Nano Res. 2017, 10, 1673–1696. [Google Scholar] [CrossRef]
  91. Cao, S.; Yu, J. G-C3N4-based photocatalysts for hydrogen generation. J. Phys. Chem. Lett. 2014, 5, 2101–2107. [Google Scholar] [CrossRef]
  92. Dong, G.; Zhang, Y.; Pan, Q.; Qiu, J. A fantastic graphitic carbon nitride (g-C3N4) material: Electronic structure, photocatalytic and photoelectronic properties. J. Photochem. Photobiol. C Photochem. Rev. 2014, 20, 33–50. [Google Scholar] [CrossRef]
  93. Liu, G.; Niu, P.; Sun, C.; Smith, S.C.; Chen, Z.; Lu, G.Q.; Cheng, H.M. Unique electronic structure induced high photoreactivity of sulfur-doped graphitic C3N4. J. Am. Chem. Soc. 2010, 132, 11642–11648. [Google Scholar] [CrossRef]
  94. Hayat, A.; Sohail, M.; Ali Shah Syed, J.; Al-Sehemi, A.G.; Mohammed, M.H.; Al-Ghamdi, A.A.; Taha, T.A.; AlSalem, H.S.; Alenad, A.M.; Amin, M.A.; et al. Recent Advancement of the Current Aspects of g-C3N4 for its Photocatalytic Applications in Sustainable Energy System. Chem. Rec. 2022, 22, e202100310. [Google Scholar] [CrossRef]
  95. Zhang, Y.; Pan, Q.; Chai, G.; Liang, M.; Dong, G.; Zhang, Q.; Qiu, J. Synthesis and luminescence mechanism of multicolor-emitting g-C3N4 nanopowders by low temperature thermal condensation of melamine. Sci. Rep. 2013, 3, 1943. [Google Scholar] [CrossRef] [Green Version]
  96. Porcu, S.; Roppolo, I.; Salaun, M.; Sarais, G.; Barbarossa, S.; Casula, M.F.; Carbonaro, C.M.; Ricci, P.C. Come to light: Detailed analysis of thermally treated Phenyl modified Carbon Nitride Polymorphs for bright phosphors in lighting applications. Appl. Surf. Sci. 2020, 504, 144330. [Google Scholar] [CrossRef]
  97. Porcu, S.; Secci, F.; Abdullah, Q.A.; Ricci, P.C. 4-Nitrophenol Efficient Photoreduction from Exfoliated and Protonated Phenyl-Doped Graphitic Carbon Nitride Nanosheets. Polymers 2021, 13, 3752. [Google Scholar] [CrossRef]
  98. Xu, Q.; Zhang, L.; Cheng, B.; Fan, J.; Yu, J. S-Scheme Heterojunction Photocatalyst. Chem 2020, 6, 1543–1559. [Google Scholar] [CrossRef]
  99. Ardizzone, S.; Bianchi, C.L.; Borgese, L.; Cappelletti, G.; Locatelli, C.; Minguzzi, A.; Rondinini, S.; Vertova, A.; Ricci, P.C.; Cannas, C.; et al. Physico-chemical characterization of IrO2-SnO2 sol-gel nanopowders for electrochemical applications. J. Appl. Electrochem. 2009, 39, 2093–2105. [Google Scholar] [CrossRef]
  100. Wang, Z.; Lin, Z.; Shen, S.; Zhong, W.; Cao, S. Advances in designing heterojunction photocatalytic materials. Chin. J. Catal. 2021, 42, 710–730. [Google Scholar] [CrossRef]
  101. Wang, H.; Zhang, L.; Chen, Z.; Hu, J.; Li, S.; Wang, Z.; Liu, J.; Wang, X. Semiconductor heterojunction photocatalysts: Design, construction, and photocatalytic performances. Chem. Soc. Rev. 2014, 43, 5234–5244. [Google Scholar] [CrossRef] [PubMed]
  102. Li, X.; Yu, J.; Jaroniec, M.; Chen, X. Cocatalysts for selective photoreduction of CO2 into solar fuels. Chem. Rev. 2019, 119, 3962–4179. [Google Scholar] [CrossRef] [PubMed]
  103. Deng, Y.; Zhao, R. Advanced Oxidation Processes (AOPs) in Wastewater Treatment. Curr. Pollut. Rep. 2015, 1, 167–176. [Google Scholar] [CrossRef] [Green Version]
  104. Serpone, N.; Borgarello, E.; Grätzel, M. Visible light induced generation of hydrogen from H2S in mixed semiconductor dispersions; improved efficiency through inter-particle electron transfer. J. Chem. Soc. Chem. Commun. 1984, 34, 342–344. [Google Scholar] [CrossRef]
  105. Serpone, N.; Maruthamuthu, P.; Pichat, P.; Pelizzetti, E.; Hidaka, H. Exploiting the interparticle electron transfer process in the photocatalysed oxidation of phenol, 2-chlorophenol and pentachlorophenol: Chemical evidence for electron and hole transfer between coupled semiconductors. J. Photochem. Photobiol. A Chem. 1995, 85, 247–255. [Google Scholar] [CrossRef]
  106. Teranishi, T.; Sakamoto, M. Charge separation in type-II semiconductor heterodimers. J. Phys. Chem. Lett. 2013, 4, 2867–2873. [Google Scholar] [CrossRef]
  107. Bai, S.; Jiang, J.; Zhang, Q.; Xiong, Y. Steering charge kinetics in photocatalysis: Intersection of materials syntheses, characterization techniques and theoretical simulations. Chem. Soc. Rev. 2015, 44, 2893–2939. [Google Scholar] [CrossRef] [PubMed]
  108. Wang, H.; Wang, J.; Zhang, L.; Yu, Q.; Chen, Z.; Wu, S. A New Strategy for Improving the Efficiency of Low-temperature Selective Catalytic Reduction of NOx with CH4 via the Combination of Non-thermal Plasma and Ag2O/TiO2 Photocatalyst. Chem. Res. Chin. Univ. 2019, 35, 1062–1069. [Google Scholar] [CrossRef]
  109. Wang, X.; Liu, G.; Chen, Z.G.; Li, F.; Wang, L.; Lu, G.Q.; Cheng, H.M. Enhanced photocatalytic hydrogen evolution by prolonging the lifetime of carriers in ZnO/CdS heterostructures. Chem. Commun. 2009, 23, 3452–3454. [Google Scholar] [CrossRef]
  110. Grätzel, M. Photoelectrochemical cells. Nature 2001, 414, 338–344. [Google Scholar] [CrossRef] [PubMed]
  111. Balayeva, N.O.; Fleisch, M.; Bahnemann, D.W. Surface-grafted WO3/TiO2 photocatalysts: Enhanced visible-light activity towards indoor air purification. Catal. Today 2018, 313, 63–71. [Google Scholar] [CrossRef]
  112. Wang, Y.; Shang, X.; Shen, J.; Zhang, Z.; Wang, D.; Lin, J.; Wu, J.C.S.; Fu, X.; Wang, X.; Li, C. Direct and indirect Z-scheme heterostructure-coupled photosystem enabling cooperation of CO2 reduction and H2O oxidation. Nat. Commun. 2020, 11, 3043. [Google Scholar] [CrossRef]
  113. Peng, Z.; Yu, J.; Jaroniec, M. All-Solid-State Z-Scheme Photocatalytic Systems. Adv. Mater. 2014, 26, 4920–4935. [Google Scholar] [CrossRef]
  114. Jiang, Z.; Cheng, B.; Zhang, Y.; Wageh, S.; Al-Ghamdi, A.A.; Yu, J.; Wang, L. S-scheme ZnO/WO3 heterojunction photocatalyst for efficient H2O2 production. J. Mater. Sci. Technol. 2022, 124, 193–201. [Google Scholar] [CrossRef]
  115. Wang, Z.; Chen, Y.; Zhang, L.; Cheng, B.; Yu, J.; Fan, J. Step-scheme CdS/TiO2 nanocomposite hollow microsphere with enhanced photocatalytic CO2 reduction activity. J. Mater. Sci. Technol. 2020, 56, 143–150. [Google Scholar] [CrossRef]
  116. Suárez-Méndez, A.; López-Curiel, J.C.; Fuentes, G.A.; Serrano-Rosales, B.; Morales-Zárate, E.; Rivera, V.M. Thiophene-Based Oligomers Formed in-situ: A Novel Sensitizer Material of TiO2/HY Hybrid Material. Top. Catal. 2022, 65, 1218–1224. [Google Scholar] [CrossRef]
  117. Zhong, Y.H.; Lei, Y.; Huang, J.F.; Xiao, L.M.; Chen, X.L.; Luo, T.; Qin, S.; Guo, J.; Liu, J.M. Design of an alkaline pyridyl acceptor-based calix [4] arene dye and synthesis of stable calixarene-TiO2 porous hybrid materials for efficient photocatalysis. J. Mater. Chem. A 2020, 8, 8883–8891. [Google Scholar] [CrossRef]
  118. Palmas, S.; Da Pozzo, A.; Mascia, M.; Vacca, A.; Ricci, P.C. Sensitization of TiO2 nanostructures with Coumarin 343. Chem. Eng. J. 2012, 211–212, 285–292. [Google Scholar] [CrossRef]
  119. Ricci, P.C.; Da Pozzo, A.; Palmas, S.; Muscas, F.; Carbonaro, C.M. Efficient charge transfer process in Coumarin 153-nanotubolar TiO2 hybrid system. Chem. Phys. Lett. 2012, 531, 160–163. [Google Scholar] [CrossRef]
  120. Yan, S.C.; Li, Z.S.; Zou, Z.G. Photodegradation Performance of g-C3N4 Fabricated by Directly Heating Melamine. Langmuir 2009, 25, 11269–11273. [Google Scholar] [CrossRef]
  121. Stagi, L.; Chiriu, D.; Carbonaro, C.M.; Corpino, R.; Ricci, P.C. Structural and optical properties of carbon nitride polymorphs. Diam. Relat. Mater. 2016, 68, 84–92. [Google Scholar] [CrossRef]
  122. Bledowski, M.; Wang, L.; Ramakrishnan, A.; Khavryuchenko, O.V.; Khavryuchenko, V.D.; Ricci, P.C.; Strunk, J.; Cremer, T.; Kolbeck, C.; Beranek, R. Visible-light photocurrent response of TiO2–polyheptazine hybrids: Evidence for interfacial charge-transfer absorption. Phys. Chem. Chem. Phys. 2011, 13, 21511. [Google Scholar] [CrossRef] [PubMed]
  123. Ma, L.; Wang, G.; Jiang, C.; Bao, H.; Xu, Q. Synthesis of core-shell TiO2@g-C3N4 hollow microspheres for efficient photocatalytic degradation of rhodamine B under visible light. Appl. Surf. Sci. 2018, 430, 263–272. [Google Scholar] [CrossRef]
  124. Ni, S.; Fu, Z.; Li, L.; Ma, M.; Liu, Y. Step-scheme heterojunction g-C3N4/TiO2 for efficient photocatalytic degradation of tetracycline hydrochloride under UV light. Colloids Surf. A Physicochem. Eng. Asp. 2022, 649, 129475. [Google Scholar] [CrossRef]
  125. Dong, S.; Chen, S.; He, F.; Li, J.; Li, H.; Xu, K. Construction of a novel N-doped oxygen vacancy-rich TiO2N-TiO2−X/g-C3N4 S-scheme heterostructure for visible light driven photocatalytic degradation of 2,4-dinitrophenylhydrazine. J. Alloys Compd. 2022, 908, 164586. [Google Scholar] [CrossRef]
  126. Rodrigues, K.F.; de Moraes, N.P.; dos Santos, A.S.; Larissa, T.; Rodrigues, A.; Brunelli, D.D. Enhanced 4-chlorophenol Degradation under Visible and Solar Radiation through TiO2/g-C3N4 Z-Scheme Heterojunction. Biointerface Res. Appl. Chem. 2022, 13, 113394. [Google Scholar] [CrossRef]
  127. Liu, Y.; Ren, M.; Zhang, X.; Yang, G.; Qin, L.; Meng, J.; Guo, Y. Supramolecule self-assembly approach to direct Z-scheme TiO2/g-C3N4 heterojunctions for efficient photocatalytic degradation of emerging phenolic pollutants. Appl. Surf. Sci. 2022, 593, 153401. [Google Scholar] [CrossRef]
  128. Fu, J.; Xu, Q.; Low, J.; Jiang, C.; Yu, J. Ultrathin 2D/2D WO3/g-C3N4 step-scheme H2-production photocatalyst. Appl. Catal. B Environ. 2019, 243, 556–565. [Google Scholar] [CrossRef]
  129. Xia, P.; Cao, S.; Zhu, B.; Liu, M.; Shi, M.; Yu, J.; Zhang, Y. Designing a 0D/2D S-Scheme Heterojunction over Polymeric Carbon Nitride for Visible-Light Photocatalytic Inactivation of Bacteria. Angew. Chem.-Int. Ed. 2020, 59, 5218–5225. [Google Scholar] [CrossRef]
  130. Xu, X.; Zhang, J.; Wang, S.; Yao, Z.; Wu, H.; Shi, L.; Yin, Y.; Wang, S.; Sun, H. Photocatalytic reforming of biomass for hydrogen production over ZnS nanoparticles modified carbon nitride nanosheets. J. Colloid Interface Sci. 2019, 555, 22–30. [Google Scholar] [CrossRef]
  131. Yin, H.; Liu, J.; Shi, H.L.; Sun, L.; Yuan, X.; Xia, D. Highly efficient catalytic ozonation for oxalic acid mineralization with Ag2CO3 modified g-C3N4: Performance and mechanism. Process Saf. Environ. Prot. 2022, 162, 944–954. [Google Scholar] [CrossRef]
  132. Pan, S.; Jia, B.; Fu, Y. Ag2CO3 nanoparticles decorated g-C3N4 as a high-efficiency catalyst for photocatalytic degradation of organic contaminants. J. Mater. Sci. Mater. Electron. 2021, 32, 14464–14476. [Google Scholar] [CrossRef]
  133. Jin, C.; Li, W.; Chen, Y.; Li, R.; Huo, J.; He, Q.; Wang, Y. Efficient Photocatalytic Degradation and Adsorption of Tetracycline over Type-II Heterojunctions Consisting of ZnO Nanorods and K-Doped Exfoliated g-C3N4 Nanosheets. Ind. Eng. Chem. Res. 2020, 59, 2860–2873. [Google Scholar] [CrossRef]
  134. Xie, K.; Wang, Y.; Zhang, K.; Zhao, R.; Chai, Z.; Du, J.; Li, J. Controllable band structure of ZnO/g-C3N4 aggregation to enhance gas sensing for the dimethylamine detection. Sens. Actuators Rep. 2022, 4, 100084. [Google Scholar] [CrossRef]
  135. Ding, F.; Ming, T.; Zhang, H.; Gao, Y.; Dragutan, V.; Sun, Y.; Dragutan, I.; Xu, Z. Plasmonic Ag nanoparticles decorated g-C3N4 for enhanced visible-light driven photocatalytic degradation and H2 production. Resour. Chem. Mater. 2022, 1, 1–7. [Google Scholar] [CrossRef]
  136. Wang, H.; Sun, T.; Chang, L.; Nie, P.; Zhang, X.; Zhao, C.; Xue, X. The g-C3N4 nanosheets decorated by plasmonic Au nanoparticles: A heterogeneous electrocatalyst for oxygen evolution reaction enhanced by sunlight illumination. Electrochim. Acta 2019, 303, 110–117. [Google Scholar] [CrossRef]
  137. Zhang, G.; Lan, Z.A.; Lin, L.; Lin, S.; Wang, X. Overall water splitting by Pt/g-C3N4 photocatalysts without using sacrificial agents. Chem. Sci. 2016, 7, 3062–3066. [Google Scholar] [CrossRef] [Green Version]
  138. Faisal, M.; Jalalah, M.; Harraz, F.A.; El-Toni, A.M.; Khan, A.; Al-Assiri, M.S. Au nanoparticles-doped g-C3N4 nanocomposites for enhanced photocatalytic performance under visible light illumination. Ceram. Int. 2020, 46, 22090–22101. [Google Scholar] [CrossRef]
  139. Che, Y.; Liu, Q.; Lu, B.; Zhai, J.; Wang, K.; Liu, Z. Plasmonic ternary hybrid photocatalyst based on polymeric g-C3N4 towards visible light hydrogen generation. Sci. Rep. 2020, 10, 721. [Google Scholar] [CrossRef] [Green Version]
  140. Li, J.; Zhu, X.; Qiu, F.; Zhang, T.; Hu, F.; Peng, X. Facile preparation of Ag/Ag2WO4/g-C3N4 ternary plasmonic photocatalyst and its visible-light photocatalytic activity. Appl. Organomet. Chem. 2019, 33, e4683. [Google Scholar] [CrossRef]
  141. Das, S.; Deka, T.; Ningthoukhangjam, P.; Chowdhury, A. Applied Surface Science Advances A critical review on prospects and challenges of metal-oxide embedded g-C3N4-based direct Z-scheme photocatalysts for water splitting and environmental remediation. Appl. Surf. Sci. Adv. 2022, 11, 100273. [Google Scholar] [CrossRef]
  142. Vasilchenko, D.; Zhurenok, A.; Saraev, A.; Gerasimov, E.; Cherepanova, S.; Kovtunova, L.; Tkachev, S.; Kozlova, E. Platinum deposition onto g-C3N4 with using of labile nitratocomplex for generation of the highly active hydrogen evolution photocatalysts. Int. J. Hydrogen Energy 2022, 47, 11326–11340. [Google Scholar] [CrossRef]
  143. Liu, R.; Yang, W.; He, G.; Zheng, W.; Li, M.; Tao, W.; Tian, M. Ag-Modified g-C3N4Prepared by a One-Step Calcination Method for Enhanced Catalytic Efficiency and Stability. ACS Omega 2020, 5, 19615–19624. [Google Scholar] [CrossRef] [PubMed]
  144. Zeng, J.; Song, T.; Lv, M.; Wang, T.; Qin, J.; Zeng, H. Plasmonic photocatalyst Au/g-C3N4/NiFe2O4 nanocomposites for enhanced visible-light-driven photocatalytic hydrogen evolution. RSC Adv. 2016, 6, 54964–54975. [Google Scholar] [CrossRef]
  145. Bedia, J.; Muelas-Ramos, V.; Peñas-Garzón, M.; Gómez-Avilés, A.; Rodríguez, J.J.; Belver, C. A review on the synthesis and characterization of metal organic frameworks for photocatalytic water purification. Catalysts 2019, 9, 52. [Google Scholar] [CrossRef] [Green Version]
  146. Qian, Y.; Ma, D.; Zhong, J. Metal-Organic Frameworks With Variable Valence Metal-Photoactive Components: Emerging Platform for Volatile Organic Compounds Photocatalytic Degradation. Front. Chem. 2021, 9, 749839. [Google Scholar] [CrossRef] [PubMed]
  147. Zhang, H.; Zhai, G.; Lei, L.; Zhang, C.; Liu, Y.; Wang, Z.; Cheng, H.; Zheng, Z.; Wang, P.; Dai, Y.; et al. Photo-induced photo-thermal synergy effect leading to efficient CO2 cycloaddition with epoxide over a Fe-based metal organic framework. J. Colloid Interface Sci. 2022, 625, 33–40. [Google Scholar] [CrossRef]
  148. Fu, Y.; Sun, D.; Chen, Y.; Huang, R.; Ding, Z.; Fu, X.; Li, Z. An amine-functionalized titanium metal-organic framework photocatalyst with visible-light-induced activity for CO2 reduction. Angew. Chem.-Int. Ed. 2012, 51, 3364–3367. [Google Scholar] [CrossRef]
  149. Hendon, C.H.; Tiana, D.; Fontecave, M.; Sanchez, C.; D’Arras, L.; Sassoye, C.; Rozes, L.; Mellot-Draznieks, C.; Walsh, A. Engineering the optical response of the titanium-MIL-125 metal-organic framework through ligand functionalization. J. Am. Chem. Soc. 2013, 135, 10942–10945. [Google Scholar] [CrossRef] [Green Version]
  150. Han, W.; Shao, L.-H.; Sun, X.-J.; Liu, Y.-H.; Zhang, F.-M.; Yang, Y.; Dong, P.; Zhang, G. Constructing Cu Ion Sites in Mof/Cof Heterostructure for Noble-Metal-Free Photoredox Catalysis. SSRN Electron. J. 2022, 317, 121710. [Google Scholar] [CrossRef]
  151. Shi, L.; Yang, L.; Zhang, H.; Chang, K.; Zhao, G.; Kako, T.; Ye, J. Implantation of Iron(III) in porphyrinic metal organic frameworks for highly improved photocatalytic performance. Appl. Catal. B Environ. 2018, 224, 60–68. [Google Scholar] [CrossRef]
  152. Lou, W.; Wang, L.; Dong, S.; Cao, Z.; Sun, J.; Zhang, Y. A facility synthesis of bismuth-iron bimetal MOF composite silver vanadate applied to visible light photocatalysis. Opt. Mater. 2022, 126, 112168. [Google Scholar] [CrossRef]
  153. Li, Q.; Fan, Z.; Zhang, L.; Li, Y.; Chen, C.; Zhao, R.; Zhu, W. Boosting and tuning the visible photocatalytic degradation performances towards reactive blue 21 via dyes@MOF composites. J. Solid State Chem. 2019, 269, 465–475. [Google Scholar] [CrossRef]
  154. Wang, Q.; Gao, Q.; Al-Enizi, A.M.; Nafady, A.; Ma, S. Recent advances in MOF-based photocatalysis: Environmental remediation under visible light. Inorg. Chem. Front. 2020, 7, 300–339. [Google Scholar] [CrossRef]
  155. Doustkhah, E.; Esmat, M.; Fukata, N.; Ide, Y.; Hanaor, D.A.H.; Assadi, M.H.N. MOF-derived nanocrystalline ZnO with controlled orientation and photocatalytic activity. Chemosphere 2022, 303, 134932. [Google Scholar] [CrossRef] [PubMed]
  156. Zeng, T.; Shi, D.; Cheng, Q.; Liao, G.; Zhou, H.; Pan, Z. Construction of novel phosphonate-based MOF/P-TiO2 heterojunction photocatalysts: Enhanced photocatalytic performance and mechanistic insight. Environ. Sci. Nano 2020, 7, 861–879. [Google Scholar] [CrossRef]
  157. Wei, Z.; Zhu, Y.; Guo, W.; Liu, J.; Jiang, Z.; Shangguan, W. Enhanced photocatalytic overall water splitting via MOF-derived tetragonal BiVO4-based solid solution. Chem. Eng. J. 2021, 414, 128911. [Google Scholar] [CrossRef]
  158. Jing, C.; Zhang, Y.; Zheng, J.; Ge, S.; Lin, J.; Pan, D.; Naik, N.; Guo, Z. In-situ constructing visible light CdS/Cd-MOF photocatalyst with enhanced photodegradation of methylene blue. Particuology 2022, 69, 111–122. [Google Scholar] [CrossRef]
  159. Karimia, M.; Sadeghia, S.; Gavinehroudi, R.G.; Mohebali, H.; Mahjoub, A.; Heydari, A. g-C3N4@Ce-MOF Z-scheme heterojunction photocatalyzed cascade aerobic oxidative functionalization of styrene. New J. Chem. 2021, 45, 6671–6681. [Google Scholar] [CrossRef]
  160. Wu, H.; Sun, Q.; Chen, J.; Wang, G.Y.; Wang, D.; Zeng, X.F.; Wang, J.X. Citric acid-assisted ultrasmall CeO2 nanoparticles for efficient photocatalytic degradation of glyphosate. Chem. Eng. J. 2021, 425, 130640. [Google Scholar] [CrossRef]
  161. Han, J.; Deng, Y.; Li, N.; Chen, D.; Xu, Q.; Li, H.; He, J.; Lu, J. A π-π stacking perylene imide/Bi2WO6 hybrid with dual transfer approach for enhanced photocatalytic degradation. J. Colloid Interface Sci. 2021, 582, 1021–1032. [Google Scholar] [CrossRef]
  162. Cai, T.; Zeng, W.; Liu, Y.; Wang, L.; Dong, W.; Chen, H.; Xia, X. A promising inorganic-organic Z-scheme photocatalyst Ag3PO4/PDI supermolecule with enhanced photoactivity and photostability for environmental remediation. Appl. Catal. B Environ. 2020, 263, 118327. [Google Scholar] [CrossRef]
  163. Zeng, W.; Cai, T.; Liu, Y.; Wang, L.; Dong, W.; Chen, H.; Xia, X. An artificial organic-inorganic Z-scheme photocatalyst WO3@Cu@PDI supramolecular with excellent visible light absorption and photocatalytic activity. Chem. Eng. J. 2020, 381, 122691. [Google Scholar] [CrossRef]
  164. Shirmardi, A.; Teridi, M.A.M.; Azimi, H.R.; Basirun, W.J.; Jamali-Sheini, F.; Yousefi, R. Enhanced photocatalytic performance of ZnSe/PANI nanocomposites for degradation of organic and inorganic pollutants. Appl. Surf. Sci. 2018, 462, 730–738. [Google Scholar] [CrossRef]
  165. Sutar, R.S.; Barkul, R.P.; Delekar, S.D.; Patil, M.K. Sunlight assisted photocatalytic degradation of organic pollutants using g-C3N4-TiO2 nanocomposites. Arab. J. Chem. 2020, 13, 4966–4977. [Google Scholar] [CrossRef]
  166. Porcu, S.; Castellino, M.; Roppolo, I.; Carbonaro, C.M.; Palmas, S.; Mais, L.; Casula, M.F.; Neretina, S.; Hughes, R.A.; Secci, F.; et al. Highly efficient visible light phenyl modified carbon nitride/TiO2 photocatalyst for environmental applications. Appl. Surf. Sci. 2020, 531, 147394. [Google Scholar] [CrossRef]
  167. Wang, J.; Wang, G.; Cheng, B.; Yu, J.; Fan, J. Congo Red photodegradation. Chin. J. Catal. 2021, 42, 56–68. [Google Scholar] [CrossRef]
  168. Hamdy, M.S.; Abd-Rabboh, H.S.M.; Benaissa, M.; Al-Metwaly, M.G.; Galal, A.H.; Ahmed, M.A. Fabrication of novel polyaniline/ZnO heterojunction for exceptional photocatalytic hydrogen production and degradation of fluorescein dye through direct Z-scheme mechanism. Opt. Mater. 2021, 117, 111198. [Google Scholar] [CrossRef]
  169. Liang, Q.; Cui, S.; Liu, C.; Xu, S.; Yao, C.; Li, Z. Self-assembly of triptycene-based polymer on cadmium sulfide nanorod to construct core-shell nanostructure for efficient visible-light-driven photocatalytic H2 evolution. Chem. Eng. J. 2019, 364, 102–110. [Google Scholar] [CrossRef]
  170. Zang, S.; Zhang, G.; Lan, Z.A.; Zheng, D.; Wang, X. Enhancement of photocatalytic H2 evolution on pyrene-based polymer promoted by MoS2 and visible light. Appl. Catal. B Environ. 2019, 251, 102–111. [Google Scholar] [CrossRef]
  171. Zhu, Y.P.; Yin, J.; Abou-Hamad, E.; Liu, X.; Chen, W.; Yao, T.; Mohammed, O.F.; Alshareef, H.N. Highly Stable Phosphonate-Based MOFs with Engineered Bandgaps for Efficient Photocatalytic Hydrogen Production. Adv. Mater. 2020, 32, e1906368. [Google Scholar] [CrossRef]
  172. Hieu, V.Q.; Lam, T.C.; Khan, A.; Thi Vo, T.T.; Nguyen, T.Q.; Doan, V.D.; Tran, D.L.; Le, V.T.; Tran, V.A. TiO2/Ti3C2/g-C3N4 ternary heterojunction for photocatalytic hydrogen evolution. Chemosphere 2021, 285, 131429. [Google Scholar] [CrossRef]
  173. Gong, S.; Teng, X.; Niu, Y.; Liu, X.; Xu, M.; Xu, C.; Ji, L.; Chen, Z. Construction of S-scheme 0D/2D heterostructures for enhanced visible-light-driven CO2 reduction. Appl. Catal. B Environ. 2021, 298, 120521. [Google Scholar] [CrossRef]
  174. Yang, C.; Wang, Y.; Yu, J.; Cao, S. Ultrathin 2D/2D Graphdiyne/Bi2WO6Heterojunction for Gas-Phase CO2Photoreduction. ACS Appl. Energy Mater. 2021, 4, 8734–8738. [Google Scholar] [CrossRef]
  175. Kong, X.J.; He, T.; Zhou, J.; Zhao, C.; Li, T.C.; Wu, X.Q.; Wang, K.; Li, J.R. In Situ Porphyrin Substitution in a Zr(IV)-MOF for Stability Enhancement and Photocatalytic CO2 Reduction. Small 2021, 17, 2005357. [Google Scholar] [CrossRef]
  176. Ke, X.; Zhang, J.; Dai, K.; Fan, K.; Liang, C. Integrated S-Scheme Heterojunction of Amine-Functionalized 1D CdSe Nanorods Anchoring on Ultrathin 2D SnNb2O6 Nanosheets for Robust Solar-Driven CO2 Conversion. Sol. RRL 2021, 5, 2000805. [Google Scholar] [CrossRef]
  177. Wang, H.; Li, H.; Chen, Z.; Li, J.; Li, X.; Huo, P.; Wang, Q. TiO2 modified g-C3N4 with enhanced photocatalytic CO2 reduction performance. Solid State Sci. 2020, 100, 106099. [Google Scholar] [CrossRef]
  178. Yang, H.; He, D.; Liu, C.; Zhang, T.; Qu, J.; Jin, D.; Zhang, K.; Lv, Y.; Zhang, Z.; Zhang, Y. Visible-light-driven photocatalytic disinfection by S-scheme α-Fe2O3/g-C3N4 heterojunction: Bactericidal performance and mechanism insight. Chemosphere 2022, 287, 132072. [Google Scholar] [CrossRef] [PubMed]
  179. Zhang, X.; Tian, F.; Lan, X.; Liu, Y.; Yang, W.; Zhang, J.; Yu, Y. Building P-doped MoS2/g-C3N4 layered heterojunction with a dual-internal electric field for efficient photocatalytic sterilization. Chem. Eng. J. 2022, 429, 132588. [Google Scholar] [CrossRef]
  180. Yang, R.; Song, G.; Wang, L.; Yang, Z.; Zhang, J.; Zhang, X.; Wang, S.; Ding, L.; Ren, N.; Wang, A.; et al. Full Solar-Spectrum-Driven Antibacterial Therapy over Hierarchical Sn3O4/PDINH with Enhanced Photocatalytic Activity. Small 2021, 17, 2102744. [Google Scholar] [CrossRef]
  181. Rathna, T.; PonnanEttiyappan, J.B.; RubenSudhakar, D. Fabrication of visible-light assisted TiO2-WO3-PANI membrane for effective reduction of chromium (VI) in photocatalytic membrane reactor. Environ. Technol. Innov. 2021, 24, 102023. [Google Scholar] [CrossRef]
  182. Shen, H.; Shao, Z.; Zhao, Q.; Jin, M.; Shen, C.; Deng, M.; Zhong, G.; Huang, F.; Zhu, H.; Chen, F.; et al. Facile synthesis of novel three-dimensional Bi2S3 nanocrystals capped by polyvinyl pyrrolidone to enhance photocatalytic properties under visible light. J. Colloid Interface Sci. 2020, 573, 115–122. [Google Scholar] [CrossRef]
  183. Yang, R.; Zhong, S.; Zhang, L.; Liu, B. PW12/CN@Bi2WO6 composite photocatalyst prepared based on organic-inorganic hybrid system for removing pollutants in water. Sep. Purif. Technol. 2020, 235, 116270. [Google Scholar] [CrossRef]
Figure 1. Application fields of photocatalysis.
Figure 1. Application fields of photocatalysis.
Molecules 27 06828 g001
Figure 2. Representative scheme of the publication rate of papers regarding photocatalysis.
Figure 2. Representative scheme of the publication rate of papers regarding photocatalysis.
Molecules 27 06828 g002
Figure 3. Basic scheme of semiconductor−mediated photocatalytic process [38].
Figure 3. Basic scheme of semiconductor−mediated photocatalytic process [38].
Molecules 27 06828 g003
Figure 4. Semiconductor’s band−gaps and redox potentials.
Figure 4. Semiconductor’s band−gaps and redox potentials.
Molecules 27 06828 g004
Figure 5. Strategies to enhance the photocatalytic activity.
Figure 5. Strategies to enhance the photocatalytic activity.
Molecules 27 06828 g005
Figure 6. Chemical structures of polymeric semiconductors.
Figure 6. Chemical structures of polymeric semiconductors.
Molecules 27 06828 g006
Figure 7. Examples of conjugated polymer photocatalysts.
Figure 7. Examples of conjugated polymer photocatalysts.
Molecules 27 06828 g007
Figure 8. 4−Nitrophenol reduction using protonated and exfoliated phenyl carbon nitride (pePhCN) as photocatalyst [97].
Figure 8. 4−Nitrophenol reduction using protonated and exfoliated phenyl carbon nitride (pePhCN) as photocatalyst [97].
Molecules 27 06828 g008
Figure 9. Heterostructures. (A) type−I (B) type−II (C) Z−scheme with redox mediator (D) all solid Z−scheme (E) S−scheme.
Figure 9. Heterostructures. (A) type−I (B) type−II (C) Z−scheme with redox mediator (D) all solid Z−scheme (E) S−scheme.
Molecules 27 06828 g009
Figure 10. Most representative synthetic methods to obtain hybrid systems.
Figure 10. Most representative synthetic methods to obtain hybrid systems.
Molecules 27 06828 g010
Figure 11. SEM images of (a) BWO nanosheets, (b) PI, (c,d) 30% PI@BWO and magnification, (e) FE-SEM images of 30% PI@BWO and (fj) EDX element mapping images of Bi, W, C, O, and N.
Figure 11. SEM images of (a) BWO nanosheets, (b) PI, (c,d) 30% PI@BWO and magnification, (e) FE-SEM images of 30% PI@BWO and (fj) EDX element mapping images of Bi, W, C, O, and N.
Molecules 27 06828 g011
Figure 12. (a) BPA photocatalytic removal performance over as-prepared samples under visible light illumination, (b) pseudo-first-order kinetics equation fitted with the BPA degradation processes over different samples, (c) kinetic rate constants (min−1) of the samples, (d) five consecutive photocatalytic degradation cycles over 30% PI@BWO [161].
Figure 12. (a) BPA photocatalytic removal performance over as-prepared samples under visible light illumination, (b) pseudo-first-order kinetics equation fitted with the BPA degradation processes over different samples, (c) kinetic rate constants (min−1) of the samples, (d) five consecutive photocatalytic degradation cycles over 30% PI@BWO [161].
Molecules 27 06828 g012
Figure 13. SEM images of PhCN/TiO2 hybrid system taken at low (A) and high magnification (B) and EDX analysis for the cross-section denoted by the dashed green. Normalized absorption spectra of TiO2 (anatase), g-C3N4, and PhCN (i.e., the building blocks for the hybrid systems) (A), the hybrid systems g-C3N4/TiO2 and PhCN/TiO2 (B), and the differential absorption of PhCN/TiO2 from building blocks (C) [166].
Figure 13. SEM images of PhCN/TiO2 hybrid system taken at low (A) and high magnification (B) and EDX analysis for the cross-section denoted by the dashed green. Normalized absorption spectra of TiO2 (anatase), g-C3N4, and PhCN (i.e., the building blocks for the hybrid systems) (A), the hybrid systems g-C3N4/TiO2 and PhCN/TiO2 (B), and the differential absorption of PhCN/TiO2 from building blocks (C) [166].
Molecules 27 06828 g013
Figure 14. PhCN/TiO2 photocatalyst and its application for the degradation of RhB and MB (A). Rate constant (k) by employing TiO2 (anatase), g−C3N4/TiO2, and PhCN/TiO2 as photocatalysts respectively (B). The stability of the catalyst is reported for the MB photodegradation (C) [166].
Figure 14. PhCN/TiO2 photocatalyst and its application for the degradation of RhB and MB (A). Rate constant (k) by employing TiO2 (anatase), g−C3N4/TiO2, and PhCN/TiO2 as photocatalysts respectively (B). The stability of the catalyst is reported for the MB photodegradation (C) [166].
Molecules 27 06828 g014
Figure 15. SEM images of (a) PDINH, (b) Sn3O4, and (c) Sn3O4/PDINH−10%. (d) TEM image of Sn3O4/PDINH−10%. HRTEM images of (e) Sn3O4 and (f) Sn3O4 /PDINH-10%. (g) EDS element mapping of Sn3O4/PDINH−10%.
Figure 15. SEM images of (a) PDINH, (b) Sn3O4, and (c) Sn3O4/PDINH−10%. (d) TEM image of Sn3O4/PDINH−10%. HRTEM images of (e) Sn3O4 and (f) Sn3O4 /PDINH-10%. (g) EDS element mapping of Sn3O4/PDINH−10%.
Molecules 27 06828 g015
Figure 16. Antibacterial activity. (a) Photographs of antibacterial results under simulated sunlight illumination (S. aureus). Antibacterial activities of the different materials under simulated sunlight illumination: (b) S. aureus and (c) E. coli. (d) Antibacterial performance of Sn3O4/PDINH heterostructure under NIR irradiation (λ > 700 nm). (e) GSH losing histograms after incubation with materials under simulated sunlight illumination or not [180].
Figure 16. Antibacterial activity. (a) Photographs of antibacterial results under simulated sunlight illumination (S. aureus). Antibacterial activities of the different materials under simulated sunlight illumination: (b) S. aureus and (c) E. coli. (d) Antibacterial performance of Sn3O4/PDINH heterostructure under NIR irradiation (λ > 700 nm). (e) GSH losing histograms after incubation with materials under simulated sunlight illumination or not [180].
Molecules 27 06828 g016
Figure 17. (a) SEM and (b,c) TEM images with different magnifications of Bi2S3−1. (d) HRTEM image and (e,f) EDS elemental mapping of Bi2S3−1.
Figure 17. (a) SEM and (b,c) TEM images with different magnifications of Bi2S3−1. (d) HRTEM image and (e,f) EDS elemental mapping of Bi2S3−1.
Molecules 27 06828 g017
Figure 18. (a) Absorption spectra of RhB solution at different times of photodegradation with Bi2S3-1 (the inset in the right corner is an image of RhB solution before and after photodegradation). (b) Variation of RhB concentration with photodegradation time in the existence of various photocatalysts (Bi2S3−1, Bi2S3−2, and TiO2) under visible light. (c) −ln (Ct/C0) versus time curves of Bi2S3−1, Bi2S3−2 and TiO2. (d) The durability tests of Bi2S3−1 for photodegradation of RhB [182].
Figure 18. (a) Absorption spectra of RhB solution at different times of photodegradation with Bi2S3-1 (the inset in the right corner is an image of RhB solution before and after photodegradation). (b) Variation of RhB concentration with photodegradation time in the existence of various photocatalysts (Bi2S3−1, Bi2S3−2, and TiO2) under visible light. (c) −ln (Ct/C0) versus time curves of Bi2S3−1, Bi2S3−2 and TiO2. (d) The durability tests of Bi2S3−1 for photodegradation of RhB [182].
Molecules 27 06828 g018
Figure 19. SEM image (a), TEM images (b), HRTEM image (c), SAED pattern (d), HAADF−STEM image (e) and corresponding EDX elemental mapping images (f) of InVO4/CN−II.
Figure 19. SEM image (a), TEM images (b), HRTEM image (c), SAED pattern (d), HAADF−STEM image (e) and corresponding EDX elemental mapping images (f) of InVO4/CN−II.
Molecules 27 06828 g019
Figure 20. (a) CO generation and (b) average gas evolution rates of H2, CH4, and CO in MeCN/H2O system over various photocatalysts for 10 h photocatalytic reaction under visible light irradiation (with a UVCUT−420 nm filter). (c) Mass spectrum of 13CO (m/z = 29) produced over InVO4/CN-II in the photoreduction of 13CO2. (d) Cycling production of CO over InVO4/CN−II under visible light irradiation [173].
Figure 20. (a) CO generation and (b) average gas evolution rates of H2, CH4, and CO in MeCN/H2O system over various photocatalysts for 10 h photocatalytic reaction under visible light irradiation (with a UVCUT−420 nm filter). (c) Mass spectrum of 13CO (m/z = 29) produced over InVO4/CN-II in the photoreduction of 13CO2. (d) Cycling production of CO over InVO4/CN−II under visible light irradiation [173].
Molecules 27 06828 g020
Figure 21. (a) Photocatalytic degradation of FLU dye over ZnO, ZnPAN1, ZnPAN3, ZnPAN5, and ZnPAN10, (b) Pseudo-first order kinetic for photodegradation of FLU dye over ZnO, ZnPAN1, ZnPAN3, ZnPAN5 and ZnPAN10, (c) The effect of scavengers on photocatalytic degradation of FLU dye over ZnPAN5, (d) PL spectrum for terephthalic acid on the sample ZnPAN5 (e) Re-cyclic of ZnPAN5 for five consecutive cycles [168].
Figure 21. (a) Photocatalytic degradation of FLU dye over ZnO, ZnPAN1, ZnPAN3, ZnPAN5, and ZnPAN10, (b) Pseudo-first order kinetic for photodegradation of FLU dye over ZnO, ZnPAN1, ZnPAN3, ZnPAN5 and ZnPAN10, (c) The effect of scavengers on photocatalytic degradation of FLU dye over ZnPAN5, (d) PL spectrum for terephthalic acid on the sample ZnPAN5 (e) Re-cyclic of ZnPAN5 for five consecutive cycles [168].
Molecules 27 06828 g021
Figure 22. Schematic illustration of the between PyP and MoS2. D = donor [170].
Figure 22. Schematic illustration of the between PyP and MoS2. D = donor [170].
Molecules 27 06828 g022
Figure 23. SEM images of (a) pure PyP, (b) MoS2 (c) 3 wt. % MoS2/PyP(IM) samples; (df) TEM images of 3 wt. % MoS2/PyP(IM) samples.
Figure 23. SEM images of (a) pure PyP, (b) MoS2 (c) 3 wt. % MoS2/PyP(IM) samples; (df) TEM images of 3 wt. % MoS2/PyP(IM) samples.
Molecules 27 06828 g023
Figure 24. The HR−TEM image (a) and SAED pattern (b) of C3N4/TiO2 composite photocatalysts [177].
Figure 24. The HR−TEM image (a) and SAED pattern (b) of C3N4/TiO2 composite photocatalysts [177].
Molecules 27 06828 g024
Figure 25. SEM (a,b) and TEM (c,d) images of as prepared photocatalysts: (a,c) pure bulk C3N4, (b,d) C3N4/TiO2 composite photocatalysts.
Figure 25. SEM (a,b) and TEM (c,d) images of as prepared photocatalysts: (a,c) pure bulk C3N4, (b,d) C3N4/TiO2 composite photocatalysts.
Molecules 27 06828 g025
Table 1. Organic-Inorganic photocatalysts and applications.
Table 1. Organic-Inorganic photocatalysts and applications.
ApplicationPhotocatalystPhotocatalysts TargetSource of LightEfficiencyRef.
Degradation of organic pollutantsCitric acid/CeO250 mgL−1 glyphosphateUV lamp 800 W/m2, 365 nm and visible lamp 1800 W/m2, 400–800 nm100%[160]
Perylene imide/Bi2WO650 mL Bisphenol A 10 ppm300 W Xe lamp cold light source100%[161]
Ag3PO4/PDI20 mgL−1 tetracycline hydrochloride300 W xenon lamp with a cut filter (>420 nm)82%[162]
WO3/Cu/PDI20 mgL−1 tetracycline hydrochloride300 W Xe lamp with a cut-off filter (λ > 420 nm)75%[163]
ZnSe/Polyaniline10 mgL−1 tetracycline hydrochlorideLED with 50 W powers and 4300 LM as a visible-light source90%[164]
g-C3N4/TiO2200 mL Bisphenol A, Brilliant green and mixed dyes 20 mgL−1Direct sunlight95%[165]
PhCN/TiO250 mL RhB/MB 10 mgL−1white LED Philips 13 W with optical power 100 mW98%[166]
S-doped g-C3N4/TiO2100 mL Congo Red 50 mgL−1300 W Xenon lamp>90%[167]
Hydrogen evolutionPolyaniline/ZnOH2OXenon lamp9.4 mmolh−1g−1 H2[168]
Cds/TCPH2O300 W Xe-lamp UV cut-off filter λ ≥ 420 nm104.51 mmolh−1g−1 H2[169]
Pyrene/MoS2H2OA 300 W Xeon lamp with a working current of 15 A2.7 mmolh−1 g−1 H2[170]
Ti-phosphonate/MOFH2Ovisible-light irradiation (λ > 400 nm)1260 mmolh−1g−1 H2[171]
gC3N4/TiO2/Ti3C2H2Osimulated visible light (Xe 300 W lamp)2592 μmolh−1 g−1 H2[172]
CO2 reductionInVO4/g-C3N4CO2 and H2Ovisible light irradiation (λ > 420 nm)69.8 mmolh−1g−1 CO[173]
Graphydine/Bi2WO6CO2 and H2Osimulated sunlight irradiation2.13 mmolh−1g−1 CH3OH[174]
Zr(IV)/MOF/CoCO2 and H2Ovisible-light irradiation70.8 mmolh−1g−1 CH4[175]
SnNb2O6/CdSe-DETACO2 and H2Ovisible-light irradiation (λ  >  400 nm)36.16 mmolh−1g−1 CO[176]
g-C3N4/TiO2CO2 and H2O8 W UV-light lamp56.2 μmolh−1 g−1 CO[177]
SterilizationCeO2/polymeric CNStaphylococcus aureusvisible-light irradiation (λ ≥ 420 nm)88% sterilization[129]
a-Fe2O3/g-C3N4Escherichia coli300 W Xenon lamp, 400 nm ultraviolet cut-off filter>80% sterilization[178]
P-doped MoS2/g-C3N4Escherichia coliVisible light100% sterilization[179]
Sn3O4/PTCDIStaphylococcus aureus and Escherichia coliSimulated sunlight>90% sterilization[180]
Reduction of heavy metalsTiO2/WO3/PANICr(VI) 10 mgL−1400 W visible lamp (2600 Lux)68%[181]
PVP/Bi2S3Cr(VI) 10 mgL−1Visible light95%[182]
PW12/CN/Bi2WO6Cr(VI) 20 mgL−11KW xenon lamp (λ > 420 nm)98%[183]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Porcu, S.; Secci, F.; Ricci, P.C. Advances in Hybrid Composites for Photocatalytic Applications: A Review. Molecules 2022, 27, 6828. https://doi.org/10.3390/molecules27206828

AMA Style

Porcu S, Secci F, Ricci PC. Advances in Hybrid Composites for Photocatalytic Applications: A Review. Molecules. 2022; 27(20):6828. https://doi.org/10.3390/molecules27206828

Chicago/Turabian Style

Porcu, Stefania, Francesco Secci, and Pier Carlo Ricci. 2022. "Advances in Hybrid Composites for Photocatalytic Applications: A Review" Molecules 27, no. 20: 6828. https://doi.org/10.3390/molecules27206828

Article Metrics

Back to TopTop