Next Article in Journal
Optimization of a Reference Kinetic Model for Solid Oxide Fuel Cells
Previous Article in Journal
Continuous-Flow Biocatalytic Process for the Synthesis of the Best Stereoisomers of the Commercial Fragrances Leather Cyclohexanol (4-Isopropylcyclohexanol) and Woody Acetate (4-(Tert-Butyl)Cyclohexyl Acetate)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Photosensitive Hybrid Nanostructured Materials: The Big Challenges for Sunlight Capture

by
Giuseppina Luciani
1,
Claudio Imparato
1 and
Giuseppe Vitiello
1,2,*
1
Department of Chemical, Materials and Production Engineering, University of Naples Federico II, p.le V. Tecchio 80, 80125 Naples, Italy
2
CSGI, Center for Colloids and Surface Science, via della Lastruccia 3, 50019 Florence, Italy
*
Author to whom correspondence should be addressed.
Submission received: 13 December 2019 / Revised: 6 January 2020 / Accepted: 8 January 2020 / Published: 10 January 2020
(This article belongs to the Section Nanostructured Catalysts)

Abstract

:
Solar radiation is becoming increasingly appreciated because of its influence on living matter and the feasibility of its application for a variety of purposes. It is an available and everlasting natural source of energy, rapidly gaining ground as a supplement and alternative to the nonrenewable energy feedstock. Actually, an increasing interest is involved in the development of efficient materials as the core of photocatalytic and photothermal processes, allowing solar energy harvesting and conversion for many technological applications, including hydrogen production, CO2 reduction, pollutants degradation, as well as organic syntheses. Particularly, photosensitive nanostructured hybrid materials synthesized coupling inorganic semiconductors with organic compounds, and polymers or carbon-based materials are attracting ever-growing research attention since their peculiar properties overcome several limitations of photocatalytic semiconductors through different approaches, including dye or charge transfer complex sensitization and heterostructures formation. The aim of this review was to describe the most promising recent advances in the field of hybrid nanostructured materials for sunlight capture and solar energy exploitation by photocatalytic processes. Beside diverse materials based on metal oxide semiconductors, emerging photoactive systems, such as metal-organic frameworks (MOFs) and hybrid perovskites, were discussed. Finally, future research opportunities and challenges associated with the design and development of highly efficient and cost-effective photosensitive nanomaterials for technological claims were outlined.

1. Introduction

Solar energy, as a renewable and low-cost energy source, represents the most reliable and green solution to address energy and environmental issues [1,2,3,4,5]. This has prompted research efforts to effectively harvest and store solar energy into chemical fuels and fine chemicals. With this aim, photocatalytic processes have been developed and employed in different technological fields. Particularly, they exploit semiconductor materials to carry out photo-driven redox reactions for a huge number of applications, such as hydrogen production by water splitting, CO2 capture and reduction to fuel molecules, pollutants degradation, as well as organic syntheses [4,5,6,7,8]. From a mechanistic point of view, irradiating a photocatalyst by UV/visible light with energy equal or higher than its bandgap, electrons (e) can jump from the valence band (VB) into the conduction band (CB) and leave holes (h+), generating electron-hole pairs. These photo-generated charge carriers migrate to the catalyst surface and are then involved in redox reactions (Figure 1a).
Several parameters determine the performances of a photocatalyst. First, light-absorption properties set the energy of generated excited states, thus driving the reactions with adsorbed species. Furthermore, the recombination of excited electron/hole pairs can occur, limiting the reactivity (Figure 1a). Light absorption by inorganic semiconductor photocatalysts is usually restricted to the ultraviolet (UV) range, and it is hard to extend their absorption features to the visible light range, only through the modification of their structures [6,7,8]. Also, the charge-recombination phenomena are recurrent. On the other hand, light-harvesting features of organic photocatalysts can be tuned through molecular design. Indeed, the mono-dispersed status of organic molecules endows the intramolecularly separated charges within the discrete organic dyes, to directly transfer out for the destruction and reformation of chemical bonds (Figure 1c) [1]. However, their performances are limited by self-quenching of excited states and photo-bleaching due to their sensitive nature. To overcome these limits, a promising approach, combining inorganic and organic components at the molecular scale, leads to hybrid materials, exhibiting peculiar intrinsic photoelectronic properties and improved photocatalytic performances (as schematized in Figure 1b).
A large variety of strategies to design hybrid materials and nanomaterials with improved photocatalytic activity has been conceived and developed based on (i) the chemical nature of both organic and inorganic components and/or their structural organization and (ii) the different approaches to combine the photoactive molecular species [1,2,3,4].
The aim of this review was to describe the most promising strategies for the design and development of hybrid nanostructured materials for efficient sunlight capture and solar energy exploitation in different applications. Specifically, a large class of photocatalysts includes conventional semiconducting oxides, among all TiO2, ZnO, SnO2, WO3, Nb2O5, and ZrO2, whose absorption range is limited to the UV radiation, because of their bandgap larger than 3 eV (Scheme 1). Therefore, huge efforts have been made to achieve their photosensitization to visible and possibly even to infrared light. Other interesting oxides, such as Fe2O3 and CuxO, have a narrower bandgap, but their band potentials maybe not suitable for some reactions, so the formation of hybrid structures can be useful not only to broaden their light absorption but also to tune their energy levels.
The most attractive solutions to improve the photocatalytic performances of oxide-based systems rely on: (i) dye sensitization, (ii) charge transfer complex sensitization, (iii) heterostructures formation, frequently including carbon-based components, and (iv) metal and non-metal doping [7,8,9,10]. Doping of TiO2 and other oxides has already been extensively investigated and reviewed. At the same time, plasmonic nanostructures have been showing an important role in the development of visible light photocatalysts for various applications [11]. In the last few years, huge progress has been made in the coupling of inorganic semiconductors with organic compounds, polymers, and carbon materials, which was the main focus of the present review. In addition, other inorganic-organic hybrid materials have recently emerged as new types of forthcoming photocatalytic materials, including metal-organic frameworks (MOFs), hybrid perovskites, and semiconducting polymers [1,12,13]. Furthermore, combinations of the previously described systems were explored to enhance photocatalytic performances. This review included the advancements reported since 2015 on all these kinds of hybrid systems for photocatalytic and related applications. Finally, future research opportunities and challenges associated with the development of highly efficient and cost-effective photosensitive nanomaterials for technological claims were envisioned.

2. Dye Sensitization

A dye is a colored substance that binds to a substrate to give it coloration. Currently, dyes are mostly synthetic and found in a variety of applications, including textiles, leather, food, organic dye lasers, contrast imaging, image sensors, and other optical devices. The use of dye compounds as sensitizers of semiconductors gained interest since the 1960s when the electron transfer from a dye to a metal oxide electrode was recognized as the origin of photoinduced currents in such systems [14]. In the early 1990s, dye-sensitized solar cells (DSCs) started to be developed, especially by Grätzel’s group [14], then dyes became widespread in diverse light-harvesting systems. The working principle of n-type sensitization is based on the following mechanism: the dye absorbs visible light, exciting an electron from the ground energy state (highest occupied molecular orbital, HOMO) to an excited state (lowest unoccupied molecular orbital, LUMO), then the electron is injected into the conduction band (CB) of the semiconductor (Scheme 2). The cycle has to be completed by restoring the initial state of the dye by means of a suitable electron donor in solution [14,15].
Along with the desired pathway of electron transfer, some processes occur, which contrast the efficiency of the sensitization, generally indicated as recombination processes. The first is direct decay from the excited state, linked to its lifetime. The injected electron may also recombine with the oxidized dye (back electron transfer) or with acceptors in solution. The detailed mechanism of charge transfer is rather complex. It can involve also empty surface states of the semiconductor, and its efficiency depends on both kinetic and energetic factors. The time scale of injection should be largely faster than that of electron decay to the ground state. Then, the electron transport in the semiconductor particle or film should be efficient enough to allow the electron to reach a surface trap, active site, or co-catalyst particle, so it can take part in a redox reaction. The relative position of energy levels at the dye/semiconductor interface (HOMO, LUMO, CB edge, possible surface states) is obviously important, and not straightforward to be experimentally determined, as these energy levels can depend on the conditions, such as dye-surface interaction and solution environment (pH, dissolved species, etc.). The efficiency of charge transfer and the stability of the photosensitizer/semiconductor system are dependent on the linking mode between the two components. Different interactions can be established, ranging from physical, hydrophobic, and electrostatic interactions to hydrogen and covalent bonding [16,17]. Although physical adsorption is the most used approach, sensitizers bound through weak physical interactions are susceptible to desorption in aqueous solution. Moreover, a stronger interaction with the surface, such as chemisorption or covalent bonding, is generally reported to favor charge transfer. Even ester anchoring bonds formed between carboxyl groups of the dye and surface hydroxyl groups revealed a favorable electronic coupling but also a tendency to hydrolyze under the photochemical reaction conditions [16]. Furthermore, the binding mode and stability are strikingly dependent on solution pH, the concentration of dye and catalyst, and the functional groups of the dye [16,18]. Different types of dyes have been employed as sensitizers. The two main groups are metal complexes and organic dyes, but quantum dots are gaining a growing interest as well. Ruthenium complexes, such as Ru-bipyridyl and Ru-polypyridyl, are the most widely tested dyes, also in solar cells, in order to extend the light absorption spectrum, increase the charge separation efficiency, and enhance the stability of the linkage with the semiconductor [15]. Organometallic complexes of several other transition metals have been studied, e.g., Zn, Sn, Cu, Fe, Co, Os, Ir, Pt. However, some of these metals and the related complex dyes are expensive, including Ru, and some of them imply concerns about their toxicity. Therefore, there is a strong interest in employing metal-free organic dyes, which are usually cheap and exhibit a variety of advantages, especially the possibility to widely control their photophysical properties through structural modifications. For example, the absorption spectra can be tuned by careful consideration of HOMO and LUMO potentials via changing the length of the conjugated system and the electron-donation ability of the donor moiety [15]. Consequently, organic dyes have become the current trend in the field and were considered in this review.

2.1. Organic Dyes in Photocatalytic Systems

Organic dyes can be distinguished in emissive (fluorescent) and non-emissive dyes. Several emissive dyes have a donor-acceptor structure, consisting of three sub-molecular units: an electron-donor group (D), a π-spacer (π), and an electron-acceptor group (A). The D-π-A framework is extensively employed in dye-sensitized solar cells (DSCs) and has also been used in photocatalytic H2 production systems because it is associated with efficient charge separation, modulation of the intramolecular charge-transfer nature and the absorption properties [15,19]. In this class of dyes, coumarin, merocyanine, triphenylamine, phenothiazine, carbazole, and their derivatives are the most popular electron donor moieties. Some of the principal groups of organic dyes employed as metal-free sensitizers are listed in Table 1. Porphyrins and their analogs, like metal-porphyrin complexes, are found in natural systems: a relevant example are chromophores in chlorophyll, ensuring an efficient solar light harvesting for photosynthesis. This role attracted the interest to implement porphyrins in DSCs and photocatalytic materials, supported by their wide range of redox, photochemical, and photophysical properties and good photostability. These macrocyclic molecules comprise a rigid two-dimensional 18-π electrons structure (Figure 2), giving high absorption coefficients in two regions of visible light, 400–450 nm (Soret band, blue) and 500–700 nm (Q-bands, red) [16]. Porphyrin- or metalloporphyrin-based nanosystems, either alone, coupled with graphene and semiconductors, or embedded in metal- or covalent organic frameworks (see Section 5) have been studied in photocatalytic processes, such as the reduction of CO2 to CO, CH4, or acetylene [20]. Phthalocyanines are structurally related to porphyrins but include more conjugated cycles, extending the Q-band light absorption to near IR, over 800 nm. Similar to porphyrin, the energy of the excited state of phthalocyanines favors the electron injection into the CB of TiO2; however, they are more rarely investigated as metal-free sensitizers [15].
Xanthene-based dyes are among the most employed in the field of dye-sensitized semiconductors. They typically absorb solar light in the 400–600 nm range and include rhodamine, eosin, erythrosine, fluorescein, rose Bengal, and a variety of derivatives and related compounds (Figure 2). Several advances in the understanding of these dye-sensitization mechanisms have been made, studying xanthene dyes in conjunction with TiO2 and many other semiconductors. Similarly, some cationic organic dyes, such as methylene blue, thiazole orange, and thionin acetate, have also been found to exhibit photosensitization activity, principally addressing them to the use in H2 production [15].
Moreover, a variety of natural pigments derived from plants, fungi, or bacteria have been presented, the most common families being flavonoids (especially anthocyanins), betalains, carotenoids, chlorophylls, tannins, and anthraquinones. Natural dyes are inexpensive, abundantly available, and can be obtained from waste biomass; anyway, their generally low stability suggests the need for protection strategies, e.g., with a conductive layer serving as both a charge transport material and a stabilizer. Several dyes extracted from biomass have been proposed as valid sensitizers in DSCs and in the photocatalytic or photoelectrochemical water splitting, with promising photon-to-current conversion efficiencies [21].
The limited stability under the operative photocatalytic conditions is the main drawback of organic dye sensitization. Most dyes tend to undergo gradual degradation following the electron transfer to the semiconductor, which can represent the first step in the radical mechanisms, leading to partial or complete oxidation of the compound. In fact, dyes are very frequently employed as model contaminants for the evaluation of the photocatalytic oxidation activity of novel materials. On the other hand, the possible sensitization effects, arising from radiation absorption by the dye, may activate different modification or degradation pathways, which might hinder a clear understanding of the action of the photocatalyst [22,23]. The stability of dyes can depend on several factors, besides the nature of the molecule and the catalyst: solution pH, irradiation wavelength and intensity, presence and concentration of suitable electron donors for the regeneration of the dye initial state, and so on. The loading of the photosensitizer itself is another parameter to control, affecting the photocatalytic performance: if the dye concentration is too high, it may tend to aggregate or saturate the semiconductor surface, resulting in activity decrease [16,17]. A larger number of anchoring groups increases the long-term photochemical and linkage stability, which eventually permits higher repeatability of the performances. Lately, many researchers have been making efforts to improve the stability of dyes, primarily through their structural design and the insight into their interaction with the semiconductors.

2.2. Organic Dye-Sensitized TiO2

Titanium dioxide (TiO2) is by far the most widely studied metal oxide photocatalyst due to its chemical and photochemical stability, low cost, biocompatibility, and excellent photoinduced activity. It exhibits three primary crystal phases—anatase, rutile, and brookite—stable to ambient conditions. Anyway, its large bandgap (about 3.0–3.2 eV, depending on the crystal phase) and the fast electron-hole recombination are the main limitations to its photocatalytic performances [24,25]. Coupling TiO2 with dyes is a classical strategy to counteract these issues. Porphyrins and phthalocyanines have been used for the sensitization of TiO2, principally as metal complexes, with Zn, Cu, Co, Ru, Sn [15]. However, applications of various metal-free porphyrins were reported in the last years, mainly in the degradation of water contaminants under visible light irradiation. Tetrakis(4-carboxyphenyl) porphyrin (TCPP) was bound on TiO2 nanoparticles by wet impregnation, giving satisfying and repeatable photocatalytic efficiency, even with 0.1 wt% dye loading [26]. It was suggested that the carboxyl groups anchoring the dye to the surface facilitated electron injection and charge separation lifetime [26]. Many researchers synthesized novel porphyrins, introducing different functional groups with the aim to improve their sensitizing performances. A halogenated porphyrin with 4-chlorophenyl groups (H2TClPP) was adsorbed on a photoelectrode made of a TiO2 nanotubes array [27], two porphyrins with carboxyl-functionalized phenyl groups were anchored to TiO2 nanorods by a solvothermal procedure [28], hematoporphyrin was combined with N-doped TiO2 [29], while sulfonated and fluorinated porphyrins with reactive –SO3H groups were conjugated on anatase TiO2 nanoparticles [30,31]. All these works reported an enhanced visible light activity in the decomposition of dyes or other model pollutants or the inactivation of bacteria. The main properties and best photocatalytic activity results of most reported dye-sensitized systems are summarized in Table 2.
The sensitizer-surface interaction is a crucial point in photoinduced procedures and has recently been studied in detail. The experimental and theoretical investigation of the adsorption of phenylporphyrins on (110) crystal facet of rutile TiO2 observed a self-metalation reaction, i.e., the formation of Ti-porphyrin complexes (Figure 3) [32,33]. The effect of the anchoring mode of porphyrin derivatives was evaluated by means of surface silanization by APTES (3-aminopropyl triethoxysilane), allowing an amide covalent grafting on TiO2 and SiO2 nanoparticles [34], or surface modification by maleic anhydride of a Fe3O4@SiO2@TiO2 magnetic composite, achieving in both cases stable anchoring and photocatalytic activity [35]. Moreover, using a dimer or trimer of porphyrin with a pyridyl anchoring group, which formed H-bonds on the TiO2 surface, enhanced both hydrogen generation rate under visible light and stability [36]. A few applications of specifically synthesized metal-free phthalocyanines have also been reported. Different substituted phthalocyanine derivatives were introduced on bare TiO2 [37,38] or co-doped TiO2 [39] by hydrothermal or sol-gel process, and an enhanced Cr(VI) reduction or dye degradation was recorded. These are among the rare cases in which the dye is linked to the metal oxide during the synthesis procedure, obtaining hybrid catalysts stabilized by covalent bonds, as reported for a calixarene-based dye introduced in the sol-gel synthesis of porous TiO2 [40]. In most works, the sensitizers are attached on the surface of nanoparticles by simple adsorption from solution.
Xanthene-based dyes include some of the traditionally most widely employed sensitizing organic dyes. Eosin Y (EY), a brominated dye with a benzoic group and an absorption band between 450 and 550 nm, results to be the most widespread in the last years. It was used in combination with TiO2 for visible light processes, e.g., diclofenac degradation [41] and hydrogen generation, with either Pt [42] or a Co(III) complex [43] as co-catalyst and triethylamine (TEA) as a sacrificial electron donor. Comparable H2 evolution rates were also obtained on titanate nanotubes modified with Ni [44] and on TiO2-MoS2 photocatalyst [45]. A promising, yet less widely investigated application of photocatalysis is in selective organic synthesis reactions. TiO2 sensitized by EY was tested in the selective aerobic oxidation of different benzyl ethers to benzoates [46] and of benzylic alcohols to aldehydes [47]. Other systems based on semiconductors different than TiO2 have employed EY, as described below.
Alizarin (1,2-dihydroxy anthraquinone), a red dye naturally found in plant roots, is another common photosensitizer along with its derivatives, such as alizarin red S (ARS), containing a sulfonate group. Recent papers by Lang, Wang, Hao, and co-workers have dealt with TiO2/ARS systems for the selective aerobic oxidation of organic sulfides to sulfoxides [48,49], amines to imines [50], and alcohols to aldehydes or ketones [51] under visible light irradiation. Most of these reactions were carried out, taking advantage of TEMPO ((2,2,6,6-tetramethylpiperidin-1-yl)oxyl) as a redox mediator, allowing to achieve excellent yields and selectivities and to improve dye stability (Figure 4). ARS was found to be more efficient than EY, thionine, and Flavin in benzyl alcohol oxidation [51], and than alizarin blue S and nuclear fast red in benzylamine oxidation [50]. The ARS-TiO2 charge transfer pathway was proposed to be a direct injection rather than the indirect one typical of dyes [49]. The oxidation of sulfides was also performed on anthraquinones-sensitized TiO2 (alizarin precursors), evidencing the key role of the chelation of Ti ions through the diol group for this family of compounds [52]. TiO2-ARS was successfully tested in another organic synthesis reaction—the cyclization of N,N-dimethylanilines and maleimides to the corresponding tetrahydroquinolines—under visible light irradiation [53].
Donor-acceptor (D-π-A) dyes provide plenty of options to tune the electronic and physicochemical characteristics of the molecules by modifying the electron donor, acceptor, or spacer group, or employing multiple donor or acceptor groups. Indeed, several works were devoted to understanding the effect of substituents, functionalities, and alkyl chain size on the photocatalytic efficiency and stability of dye-sensitized TiO2. All the following cases applied the sensitized photocatalysts for H2 generation. For instance, in alkoxyphenyl-substituted carbazole dyes [54] and thiophenothiazine dyes [55], the longest alkyl chain gave the best H2 production rates, thanks to the higher hydrophobicity, extending the charge separation lifetime; in triphenylamine-based dyes, bulkier substituents and additional phenyl donating groups [56] or a thiazolothiazole unit [57] gave higher stability and yields. Phenothiazine-based dyes, featuring thiophene-based spacers and different anchoring groups, have been extensively studied [19], determining an improvement of visible light H2 generation of Pt/TiO2 by the effect of the spacer structure [58] and the insertion of a peripheral glucose unit as co-adsorbent (Figure 5) [59]. Other phenothiazine- and carbazole-based dyes were designed for the same application [60], including a D-π-A-π-A photosensitizer, with benzothiadiazole as auxiliary electron withdrawing unit [61], and, in particular, a dibenzopyrromethene (BODIPY)-conjugated phenothiazine, which gave Pt/TiO2 high activity up to near-infrared (NIR) region [62].
Different types of photoactive compounds can be included in the design of a dye with specific structure: some D-A-π-A perylene dyes, featuring cyanoacrylic acid and dicyanomethylene rhodamine as the acceptor/anchoring group, combined with a N-annulated perylene donor and a quinoxaline auxiliary acceptor [63]; two triphenylamine-benzimidazole-based dyes bound on TiO2, with Cu2WS4 nanocubes as an alternative water splitting co-catalyst to Pt [64]; a particular composite dye, consisting of π-conjugated indoline–rhodanine with a chlorophyll derivative, which induced panchromatic absorption in the visible range, efficient electron transfer, and prolonged stability [65]; a series of dyes with triphenylamine donor, vinyltiophene bridge, and a cationic pyridinium acceptor (cationic D-π-A photosensitizers) [66]. A superior H2 evolution rate was recently reached with a thioalkyl-substituted tetrathiafulvalene dye on mesoporous TiO2 [67].
Squaraine dyes, based on squaric acid (diketocyclobutenediol), are gaining interest in several applications, including red light and NIR harvesting in DSCs and organic solar cells, owing to their high absorptivity in the region around 600–800 nm and to their easy structure tunability [68]; nonetheless, their use in photocatalytic systems has been limited. Two examples are: a zwitterionic squaraine dye, anchored through π–π interaction to graphene in TiO2/rGO nanocomposites, active in H2 production [69], and squaraines featuring dodecyl spacers and carboxylic or phosphonate anchoring groups, coupled with Pt/TiO2 or Re(I)/TiO2 to obtain photocatalysts tested in H2 evolution and CO2 reduction [68].
As mentioned above, a current trend is a search for effective natural dyes extracted from plants or microorganisms. Recent reports on this kind of dyes as TiO2 sensitizers include curcumin, a simple conjugated diketone found in turmeric (Curcuma longa) [70], and anthocyanin [71]. In several works, natural dyes are extracted from the species containing them, e.g., chlorophyll and flavonoid from parsley leaves and turmeric roots [72], chlorophyll and other pigments from cyanobacteria [73], anthocyanin-rich mixture from Vaccinium meridionale Swartz (Andean blueberry) [74], extract from Picramnia sellowii tree [75], or Chlorella pyrenoidosa, a green unicellular alga used as both sensitizer and biotemplate for Fe3O4–TiO2 composites [76]. Further examples of organic dyes that have been applied in TiO2 photocatalysis are malachite green (also known as brilliant green) [77] and Sudan black B, a diazo compound [78], for the degradation of polyethylene and phenol, respectively.
Relatively few studies so far have faced the theoretical modeling of the structural and electronic features of the dye-semiconductor interface and its effect on photocatalytic activity. A dedicated DFT investigation considered TiO2 nanoclusters with different adsorbed dyes: a small rigid molecule based on coumarin, a larger rigid oligomethine cyanine dye with indole groups, and the penicillin V antibiotic, with a flexible backbone [79].

2.3. Other Organic Dye-Sensitized Semiconductors

Many other semiconducting oxides received the most attention so far in view of their photocatalytic potential and photosensitization. Zinc oxide (ZnO) is one of the most considered, thanks to the suitable properties as very tunable material with high-sensitivity, large specific area, non-toxicity, good compatibility, and high isoelectric point. Also, ZnO showed an excellent chemical stability, as well as a thermal stability, as a n-type semiconducting material, favoring a wide range of applications: from photocatalysis to production of solar cells and supercapacitors, as well as luminescent material for biomedical, biosensors piezo-electrics, and optoelectronic devices, in the form of bulk crystal, thin-film, and pellets [92]. However, ZnO nanostructures resulted suitable for absorption in ultraviolet rather than visible light because of the wide bandgap (3.3 eV). Hybrid ZnO was synthesized with organic dye-sensitizers, such as 4-((2-hydroxynaphthalen-1-yl)diazenyl)benzoic acid (4ABBN), by precipitation or sol-gel route, for enhanced H2 generation from water-methanol solution without co-catalysts [84], as well as curcumin by precipitation, for the degradation of polycyclic aromatic hydrocarbons in daylight conditions [82]. ZnO nanorods with adsorbed fluorescein, perylene-3,4,9,10-tetracarboxylic dianhydride (PTCDA), or protoporphyrin provided enhanced visible-light degradation of dyes and resorcinol [81]. A ZnO/alizarin red S system was examined in the presence of TEMPO and Ag+-Na+ electrolyte for the oxidation of aromatic and aliphatic alcohols [86]. Eosin Y and rose Bengal showed a synergistic sensitization of ZnO nanorod array with deposited Pt in H2 evolution under visible light [85]. Natural dyes, in particular, anthocyanin extracted from Karkade (Hibiscus) flowers, have been associated with ZnO as an effective sensitizer to eliminate E. coli bacteria and mineralize their organic components under simulated solar light [83]. Attempts to extend the ZnO sensitization to the near-infrared (NIR) region have been reported, thanks to conjugation with cyanine dye (IR820) [93], protoporphyrin (PPIX), or squaraine dye (SQ2) [80]. Finally, the interfacial charge transfer between a coumarin dye and ZnO quantum dots was studied by transient absorption and DFT, identifying a rather direct electron transfer mechanism [94].
Zirconium dioxide (ZrO2) has a bandgap larger than 5 eV, which poses a strong limitation on its radiation absorption range; on the other hand, the highly negative redox potential of its conduction band makes it interesting for photocatalytic reductions. Very few attempts of ZrO2 sensitization have been reported; a less recent work studied the charge transfer between baicalein, a flavonoid dye, and ZrO2 nanoparticles, in comparison with TiO2 [95].
Tin dioxide (SnO2) is another oxide, which is attracting interest in photocatalysis and photovoltaics (as an electron transport layer in solar cells). Anyway, there are almost no reports about the photosensitization of SnO2 with organic compounds. Yang et al. demonstrated that Eosin Y was able to sensitize Sb-doped SnO2, achieving visible light activity in H2 production from the TEOA solution with 0.1 g EY/g SnO2 [87]. Scarce information is available regarding other metal oxides, to the best of the authors’ knowledge. Iron oxides show visible light responsivity as a result of their relatively small bandgap; they require modifications to enhance the photocatalytic performances. Anchoring of TCPP porphyrin on Fe3O4 nanoparticles was reported to improve the removal of methylene blue (MB), although it appears not clear if that result was to actually correlate to sensitized degradation or prevalently to enhanced adsorption [96].
Some mixed metal oxides have been tested as sensitized photocatalysis as well: zinc tungstate (ZnWO4) nano rod-like samples prepared by hydrothermal method showed sensitization by methylene blue (MB) in the visible light reduction of Cr6+ [88]. Bismuth-based semiconductors have been lately gaining interest: BiPO4, BiOCl, and BiOF were sensitized, respectively, by Eosin Y [89], Rhodamine B [90], and quercetin (a natural flavonoid) [91], showing enhanced photodegradation activity under visible light.

3. Ligand-to-Metal Charge Transfer Complexation

Dye sensitization can be regarded as an indirect photosensitization process since it occurs in two steps, starting with light absorption by the dye (Scheme 2). A different sensitization mechanism (type II) has been observed with compounds that do not absorb visible light by themselves but are able to induce an extended light absorption when coordinated with metal ions on the surface of a semiconductor. In this case, the electron transfer occurs in a single step, directly from the ground state (HOMO) of the sensitizing ligand to the conduction band of the semiconductor, without involving an excited state [10,18,97]; hence, the common definition, ligand-to-metal charge transfer (LMCT), or, more in general, molecule-to-band or ligand-to-band charge transfer (Scheme 2).
Several organic compounds, including relatively small molecules, and some metal complexes have been shown to give this kind of direct sensitization when coordinated or sometimes even only adsorbed on the semiconductor. Catechol and its derivatives are the most studied, but also some carboxylic acids, diketones, and inorganic metal complexes (e.g., hexacyanoferrate, [Fe(CN)6]4−) have shown this property. In principle, when a sufficient degree of electronic coupling is established between the molecular orbitals of the ligand and the CB of the semiconductor, a direct electron injection can occur. These sensitizers have a rather wide HOMO-LUMO gap, which makes the involvement of their excited state in the charge transfer less probable.
The characteristics of LMCT complexes (e.g., binding mode, electronic structure, optical response, charge transfer efficiency, and photochemical stability) are closely related to the chemical structure and electron energy levels of the ligands [10]. The establishment of a strong electronic coupling is favored in case of electron-rich compounds, often containing one or more aromatic rings, with suitable functional groups (e.g., enediol, carboxylate, nitrile, and alcohol) to form stable interactions, such as bidentate coordination, with surface atoms, rather than a mere adsorption [10,18]. These electron-rich ligands tend to determine a large redshift in the absorption edge of wide bandgap semiconductors, usually in the visible range, reducing their effective bandgap. In this case, the position of the absorption band of the complex depends on the relative energy levels of the ligand’s HOMO and the semiconductor’s CB edge. It should be underlined that the HOMO and LUMO potentials of organic ligands are not absolute values, but the result to vary according to the semiconductor they are bound to [98,99].
The distinction between type I (indirect) or type II (direct) charge transfer (CT) mechanism occurring in a sensitized system is not always clear and straightforward. Indeed, in some cases, both pathways have a similar probability to be active, even if one can be kinetically favored. In addition, although dyes are commonly associated with indirect CT, some visible light-absorbing molecules are predicted to prefer a direct electron injection [49,69,100]. Direct photosensitization should result in a more efficient electron transfer to the conduction band; on the other hand, the back electron transfer (recombination) might also be favored [97]. Although a stronger electronic coupling is considered to imply a more efficient charge injection, a slightly lower coupling may provide a partially delocalized LMCT, allowing longer charge separation lifetimes and resulting in overall higher efficiency [101]. To date, a limited number of studies have dealt with the detailed evaluation of the electron transfer processes (absorption and emission on the fs-ps timescale) in CT complexes by transient spectroscopic techniques or computational methods.
Semiconductors absorbing only UV radiation, like TiO2, happen to show activity under visible light in the degradation of some colorless organic substrates (e.g., phenol, chlorophenols). The reason can be individuated in the LMCT occurring between the substrate, even if only weakly adsorbed, and the photocatalyst. The electron injection from the adsorbate is indeed the first step of its oxidation. The organic radical can accept an electron from a donor in solution, restoring its original form, or it can undergo successive oxidation steps; for example, other electron transfer or the attack by a reactive oxygen species (ROS), eventually leading to its complete degradation [10,18,102]. As a consequence, similarly to dyes, the stability of LMCT sensitizers in the photoreaction conditions may represent an issue that must be carefully evaluated. Nonetheless, these ligands, especially the simplest as catechol, are often reported to be quite resistant to self-oxidation and often added in the reaction system as electron donors, forming CT complexes by in situ adsorption on the semiconductor’s surface [103]. Cheap and renewable-derived compounds, e.g., glucose and other saccharides, might be used in this way as sacrificial agents to foster visible light photocatalytic activity.
In the synthesis of hybrid semiconductors with LMCT complexes, various factors have to be taken into account: the type of linking groups of the ligand, the exposed crystal planes of the semiconductor, the accessibility of its metal ions for the coordination, pH of the reaction environment, etc. Bidentate bridging or chelating complexation modes give stronger bonding of the ligand than monodentate coordination, usually reflected in a more efficient and stable CT complex. It should be taken into account that the investigation of the photoinduced CT mechanisms involving small molecules may also turn helpful in the evaluation of the behavior of other compounds with the same structural units or functional groups (dyes and/or polymers), useful for photoinduced applications.

3.1. LMCT-Sensitized TiO2

The large majority of works on photocatalysts sensitized by direct charge transfer complexes refer to TiO2. Catechol (1,2-dihydroxybenzene) and its derivatives are the most extensively studied molecules in the branch of CT complexes. The versatile functions of catechol-based compounds mainly arise from the unique properties and reactivity of the enediol (o-dihydroxyaryl) group, making them ideal bidentate ligands, able to interact with all kinds of surfaces and strongly chelate many transition metals. The adsorption and coordination of catechol and derivatives on the surface of TiO2 have been studied both experimentally and theoretically, and different binding modes have been proposed, depending also on the considered TiO2 crystal polymorph and facet (Figure 6a): bidentate chelating, bidentate bridging, monodentate, hydrogen bonding, or mixed configurations [97,104]. The most recent DFT and experimental studies on the structural and electronic characteristics of the TiO2/catechol interaction consider catechol as ligand or anchoring group not only for the application in photocatalytic processes [102] but also for type II dye-sensitized solar cells [104,105,106].
Catechol complexation shifts the absorption onset of TiO2 up to over 600 nm (Figure 6b). The presence of substituents on the catechol ring can enhance the visible light absorption and the photochemical response of the complexes; electron-donating or electron-withdrawing groups are reported to have different effects on the charge separation lifetime and photocatalytic efficiency of the sensitized system [106,107,108,109]. For example, electron-withdrawing substituents, such as cyano- or sulfonic groups, fostered H2 production on Pt/TiO2 [109]. Other related molecules with enediol functional moiety have been proposed, such as 2,3-dihydroxynaphthalene (or 2,3-naphthalenediol, DHN) [110,111,112,113] and ascorbic acid [114]. Salicylic acid [102,113,115,116], as well as the correlated 1-hydroxy-2-naphthoic acid [117] and 5-aminosalicylic acid [116], have shown good sensitization potential, with results comparable to catechol. Interestingly, also glucose, a safe, cheap, and available biomass-derived compound, is able to sensitize TiO2 [103] or TiO2/reduced graphene oxide [118] by surface adsorption, allowing visible light activity in Cr(VI) reduction and O2 reduction to H2O2. Monodentate ligands, such as phenol and its derivatives, induced a remarkable extension of the TiO2 absorption range as well [119]. Even a carboxylate-rich porous carbon obtained by gluconate carbonization on TiO2 was hypothesized to give visible light activity because of LMCT processes [120].
The main properties of recently reported systems, featuring LMCT complexes and their photocatalytic performances under visible or solar light irradiation, are summarized in Table 3. Despite the number of fundamental and characterization studies, the reports on the photocatalytic properties of such sensitized hybrid materials have started to increase only in the last years. Shi et al. realized the selective oxidation of amines to imines with air on TiO2 sensitized in situ with different substituted catechols [107] or salicylic acids [115], under blue LED illumination, achieving excellent yields with 0.8 wt% sensitizer, thanks to the cooperation of TEMPO as co-catalyst. Not only oxidation but also reduction reactions were performed by means of sensitized photocatalysts: Fukui and co-workers used TiO2 impregnated with DHN for the visible-light-driven conversion of aromatic aldehydes to the corresponding alcohols [111] or nitroaromatics to amino aromatics [112]. The visible light activity of the TiO2 surface modified with 2-naphthol was evaluated in the removal of RhB, As(III) [121], or Cr(VI) [122].
It is interesting to note that most catecholic compounds can undergo oxidative self-polymerization in suitable conditions, such as alkaline pH or light irradiation. For instance, polydopamine is receiving great attention because of its adhesive and hydrophilic properties. Core-shell TiO2 nanoparticles with a thin polydopamine layer have been found very active in visible light degradation of RhB [123]. Dopamine and its derivatives norepinephrine and nitrodopamine were polymerized on TiO2, inducing visible activity in the removal of pollutants, especially through photoreduction reactions [124]. Contrary to dopamine, complex polydopamine appears to act through a dye sensitization mechanism (Scheme 3a). Even catechol has revealed the ability to polymerize on TiO2 surface: such sensitized TiO2 powders were active in Cr(VI) reduction and H2 production [125], while dip-coated films worked for the inactivation of Escherichia coli and Staphylococcus aureus in water [126]. Employing polymerized sensitizing ligands should turn out to be an effective strategy to improve the stability of the sensitizer against self-degradation [10,18].
Other classes of organic compounds have been demonstrated to be capable of LMCT sensitization [10]. Recent reports include: carboxylic acids, such as oxalate [128], amino acids [129], nitrogen heterocyclic compounds, such as pyridine [130], 8-hydroxyquinoline [110,131], and 2-methylisoquinolinium [132], zwitterionic N-substituted aminosulfonic acids [133], or 2-methoxyethanol [134]. Curiously, even apolar aromatic hydrocarbons, lacking the common coordinating functional groups, have exhibited the activation of LMCT with metal oxides: benzene adsorbed on Pd/TiO2 gave cross-coupling reaction with cyclohexane under visible light, thanks to the activation of a C–H bond [135]. Moreover, acetylene derivatives, such as octyne and ethynylpyrene, used as capping ligands on metal oxide nanoparticles (e.g., TiO2, ZnO, RuO2), create efficient electronic coupling through their π electrons (M–O–C≡C– core–ligand linkages) and then interfacial charge delocalization, enhancing sensitized photoluminescence and photocatalytic activity [136].
The peculiar complexation of catecholic compounds on TiO2 surfaces allows the production of materials with interesting properties besides those directly depending on LMCT sensitization. Dopamine can be used as a powerful and flexible bridge, providing good interfacial contact and charge separation between different components of nanocomposites, as demonstrated for TiO2 grafted on a glycidyl methacrylate-ethylene glycol dimethacrylate copolymer [137], on N-doped carbon in core/shell nanostructures [138], or melamine-graphene nanocapsules [139]. DOPA (3,4-dihydroxyphenylalanine) and DHICA (dihydroxyindole carboxylic acid) are easily polymerized as conjugated on the surface of oxide colloidal nanoparticles, determining the formation of a Ti-indole LMCT, which induced the oxidative polymerization of monomers to different melanin polymers. The final hybrid TiO2-melanin nanostructures have shown excellent antimicrobial activity even without irradiation, owing to their ability in generating reactive oxygen species (ROS) [140,141,142,143]. Another interesting family of compounds, yet still poorly studied in the frame of CT complexes, are β-diketones: acetylacetone (2,4-pentanedione) has been shown to considerably shift the absorption onset of TiO2 [101,144] and ZrO2 [145] through LMCT. Moreover, hybrid oxide-acetylacetonate materials prepared by a sol-gel procedure exhibit the peculiar ability to generate and stabilize on their surface the superoxide radical anion (a pivotal ROS), with a consequent activity in organic pollutants degradation even in the dark [144,145,146,147]. In most reports, sensitized hybrid materials are prepared by surface adsorption on commercial or synthesized TiO2 nanoparticles, whereas LMCT sensitized thin films have been seldom investigated, besides some TiO2 coatings whose surface was functionalized with catechol [126,148]. An alternative preparation method for hybrid bulk materials or films consists of a bottom-up sol-gel route with the introduction of the complexant in a Ti alkoxide solution, obtaining hybrid gels [144] or sols used for dip coating [110] or spin coating [149].

3.2. Other LMCT-Sensitized Metal Oxides

Very few reports on the charge transfer complex sensitization of other semiconductors for photocatalytic purposes are available. Despite the extensive research on its applications in photocatalysis, zinc oxide has not been the object of much research in this view. Only some studies on the interaction of ZnO with sensitizing ligands were found, e.g., with catechol [152], 1,2-dihydroxynaphthalene, and 8-hydroxyquinoline [131], or dopamine and modified catechol with porphyrins [153]. ZrO2 and Fe2O3 were previously proven to extend their absorption bands following the adsorption of dopamine, ascorbic acid, and alizarin [98]; no recent updates regarding direct sensitization of these oxides were found, although amorphous ZrO2 complexed with acetylacetone showed absorption extended up to the visible range [145]. Niobium oxide (Nb2O5) has been proposed as a photocatalyst as well, and its photoactivity with visible light has been connected with the formation of LMCT with the substrates. This was shown for aromatic hydrocarbons on vacuum heat-treated Nb2O5: toluene adsorption under visible radiation leads to its oxidation by a ROS-mediated radical mechanism [151]. Another work reported visible light degradation of 4-chlorophenol and concurrent enhancement of Cr(VI) reduction on Nb2O5 nanorods/graphene composite [154]. Titanates, wide bandgap semiconductors with the perovskite crystal structure, also show interesting photocatalytic properties. Studies on the effect of catechol, salicylic acid, and 2,3-dihydroxynaphthalene on interfacial CT on TiO2, SrTiO3, and BaTiO3 have been performed in relation to their band edge potentials [99,113], underlying a broad shift of the absorption onset and enhancement of the photoelectrochemical and photocatalytic properties, not only under visible light but also with UV irradiation. Orthoferrites are less known as perovskite materials in the photocatalysis field; ErFeO3 functionalized with tartrate ligands was found to show multiple photoluminescence and a considerable photocatalytic activity [150]. Finally, the emerging bismuth-based semiconductors, such as BiOIO3, were lately investigated with many different dihydroxybenzoic acids as complexing ligands: compounds with an ortho-dihydroxyphenyl structure chelate Bi(III) surface sites forming CT complexes (Scheme 3b) cause the formation of surface oxygen vacancies, affecting the photocatalytic properties over the hybrid semiconductors [127].

4. Carbon-Based Hybrid Nanomaterials

In the last years, carbon-based nanostructured materials, such as graphene and its derivatives [155,156], fullerenes, activated carbons [26,157], biochar [158,159], carbon nanotubes (CNTs) [13,160], and nanoparticles [161], have been of great interest because of their unique structural, electrical, and mechanical properties, offering promising opportunities for photocatalytic applications (Figure 7).
Indeed, carbon-based nanostructured materials can play important roles in the design and synthesis of novel hybrid photocatalysts for different applications [160,162], thanks to a good electron conductivity (related to the characteristic zero bandgap), large surface area, excellent physicochemical stability, and facile synthesis protocols based on a wide variety [163]. Also, the morphologies and crystal structures of carbon-based nanostructured materials can be opportunely defined and produced according to the specific requirements of different applications (Figure 8).
The combination of carbon-based nanomaterials with a wide variety of inorganic materials of different sizes, shapes, and compositions leads to hybrid nanostructured materials with tunable catalytic, electronic, optical, magnetic, and redox properties [162]. For example, the vastly increased surface-area-to-volume ratio of catalytic metal oxides or alloy nanoparticles opens new possibilities in electro- and photocatalysis. Such coupling of different nanomaterials has attracted considerable recent attention since the resulting nanohybrids can be potentially applied in many technological areas. Hetero nanostructures of carbon and inorganic materials usually enhance the assets of both components and, due to cooperative effects, can lead to novel functionalities and properties with respect to those of the initial components. Such improvements can reflect in an enhancement of the electrical, thermal, chemical, and/or mechanical properties. In this context, by following different synthetic approaches, carbon-based components have been combined with many inorganic semiconducting oxides, principally TiO2 and ZnO, as well as BiVO4 or WO3, to produce hybrid nanostructured materials for advanced photocatalytic applications. Indeed, a great number of examples have appeared in the literature, where carbon-based materials have been used as additives of wide bandgap semiconductors, acting either as an electron acceptor or/and sensitizer.

4.1. Carbon-Based/TiO2 Hybrids

TiO2, a wide bandgap semiconductor, presents a light absorption limited to the UV region and a very fast charge recombination. In this way, one of the most studied strategies to improve the photocatalytic activity is based on its combination with a carbon-based material (p-type semiconductor) in order to induce an enhancement of charge separation yield and extension of the electron lifetime [13]. By following different synthetic approaches, a large variety of hybrid nanostructured materials were recently designed and prepared, combining TiO2 with different types of carbon-based materials (Figure 9), such as graphene, graphene oxide (GO) [164,165] and reduced graphene oxide (rGO) [69,118,166,167,168], CNTs [164,169], carbon particles [161,164,170,171,172], fullerene [169,173], and biochar [158,159,174], working as adsorbers, electron acceptors, and transporters.
These hybrid nanostructured materials showed an enhanced visible light absorption, also exhibiting both excellent adsorption capability and photocatalytic activity, for the removal of pollutants [169,175,176] and heavy metal ions from wastewater [177,178,179], photo-oxidative treatment of organics [167,168,180], as well as water splitting and hydrogen production [178,181,182,183,184,185]. In all cases, hybrid heterostructures have shown a remarkably improved activity, not only due to the bandgap narrowing but also to the large surface area of the hybrid nanostructures in terms of charge transport, proposing them as good photoactive nanomaterials. A summary of the carbon-based hybrid materials considered in this review is reported in Table 4.

4.2. Carbon-Based/ZnO Hybrids

ZnO has been widely considered as a photocatalytic semiconductor. Actually, its photocatalytic efficiency and recycling catalytic utilization efficiency are significantly limited because of the quick recombination of photoelectron-hole pairs. Therefore, improving the photocatalytic and recycling efficiency of ZnO nanostructures is the key to practical photocatalytic applications. Recent research efforts have focused on ZnO-based heterojunctions, enhancing the photocatalytic properties of hybrid nanostructured materials [186,187,188,189]. Among them, heterostructures with carbon-based materials have stimulated a very particular interest [188,190]. In this context, many examples of hybrid nanostructured materials have been designed. Particularly, graphene (and its derivatives)/ZnO nanocomposites significantly expand the photocatalytic degradation activity of these host materials [191,192]. Although several methods have been employed to fabricate photocatalytic ZnO nanoparticles adorned with graphene, the synthesis of such nanohybrids with well-ordered particle size without agglomeration for enhanced photocatalytic ability is a challenge. ZnO nanoparticles and graphene can be synthesized by various techniques in order to obtain ZnO nanoparticles decorated graphene nanocomposite for wastewater decontamination. Another important class of carbon-based heterostructures for photocatalysts is constituted by CNTs/ZnO hybrid nanomaterials due to the peculiar molecular geometry and excellent optoelectronic properties of CNTs [193]. These materials could be tailored to maximize their photocatalytic degradation abilities due to their continuous conducting pathways and extensive interfacial contact. Many ZnO/CNTs heterostructures have been prepared by following different synthetic approaches [193,194], inducing the formation of self-assembly hybrid nanomaterials in which multi-morphology ZnO nanostructures have been tightly combined with CNTs. This approach improved the light absorption efficiency of ZnO nanostructures, as well as the separation efficacy of the photogenerated electron-hole pairs.
Finally, modifications of ZnO photocatalyst with carbon quantum dots (CQDs) have been mostly considered due to low toxicity, aqueous stability, enhanced surface area, economic feasibility, good biocompatibility, and chemical inertness of CQDs [161,172]. Many strategies have been developed to enhance the photocatalytic activity of conventional ZnO via coupling with CQDs, by principally following green and bio-sustainable chemical approaches [195,196,197,198,199]. Therefore, the unique properties of carbon-based/ZnO nanocomposites were used for charge separation, as they reduce the e/h+ pair recombination [200]. The photogenerated e/h+ pairs are involved in oxidation and reduction reactions on the surface of photocatalysts, producing reactive oxygen species (ROS), such as superoxide (•O2) and hydroxyl radicals (•OH). The formed radicals interact with molecular structures of contaminants, which decompose into less hazardous materials. In all these cases, the use of carbon-based hybrid materials represents the main strategy to enhance the photodegradation of various pollutants in the visible light region. However, the main goal of the researchers is actually focused on the possibility to optimize the particle size and surface area via many easier synthesis methods, in order to make carbon-doped ZnO an exclusive photocatalyst with the use of solar light.

4.3. Other Carbon-Based/Metal Oxide Hybrids

The photocatalytic activity of many other semiconductor oxides was greatly enhanced by conjugation with carbon-based nanomaterials. In this context, iron oxides, such as Fe2O3 or Fe3O4, represented promising candidates for the development of visible-light photocatalysis. These oxides were n-type semiconductors, with a small bandgap (∼2 eV), high stability under ambient conditions and aqueous solution at pH > 3, low toxicity, and a good tendency to absorb visible light. For example, hematite, Fe2O3, and nanoparticles with large surface area and porous structure have been prepared and revealed distinguished photocatalytic activity. Particularly, hybrid heterostructure of α-Fe2O3 with highly hydrophilic carbon dots has been recently reported as a photocatalyst to convert longer-wavelength radiation, e.g., NIR light, into shorter wavelengths, such as visible light, making them of great potential for water splitting [201] and pollutants degradation [202,203]. Similarly, hybrids formed by three-dimensional (3D)-reduced graphene oxide nanosheets containing Fe3O4 nanoparticles have demonstrated superior electrochemical performance as supercapacitors electrode [204], as well as graphene/Fe3O4 nanocomposites, showed improved photocatalytic activity [205,206]. Also, CuxO-based heterostructures received great attention. For example, nanostructured hybrids with carbon materials, such as CNTs, graphene, C60, carbon quantum dots, carbon fibers, activated carbon, carbon black, have been produced to enhance the performance of CuxO photocatalysts for H2 production by photocatalytic water splitting under visible light irradiation [207,208,209]. Similarly, carbon-based nanomaterials were combined with tungsten oxides [209,210], indium oxide [211], or bismuth vanadate [212], showing high photocatalytic activity in water splitting or antibiotics degradation.
As summarized in Table 4, a lot of hybrid carbon-based/semiconducting oxides photocatalysts have been designed and developed in the most recent years, by using different synthetic strategies. At the same time, many technological applications have been considered till now; principally, the main claim was the photodegradation of standard organic molecules, such as rhodamine or methylene blue, also showing promising efficiency results. High degradation efficiencies, in the range of 70%–100% and within two hours on average, have obtained in most of the considered studies, strongly depending on the type of produced heterostructure. In this context, the CNTs-based nanohybrids appeared as the most efficient ones, also under simulated sunlight. However, it still appears necessary to hopefully enhance their application in the photodegradation of the most diffuse pollutants in an aqueous environment, i.e., anti-inflammatory and/or antibiotics, to resolve an important environmental question. Also, interesting results have been recently obtained for H2 production, although many steps forward must be taken to enhance productivity under sunlight.
Finally, an important perspective for the future of this class of hybrid photoactive materials is the possibility to expand the use of bioinspired approach and bio-sustainable sources (i.e., organic wastes) to produce efficient photocatalysts, with the final aim to strengthen the development of green and sustainable chemical processes.

5. Metal-Organic Frameworks (MOFs)

Among organo-inorganic hybrid materials, metal-organic frameworks (MOFs) display very intriguing properties for photocatalytic applications [214]. They are crystalline porous networks obtained, connecting metal-heteroatom clusters and organic ligands through coordination bonds. Also known as porous coordination polymers (PCPs), due to the nature of interaction generating the lattice, these materials show extremely versatile design flexibility as well as modular features. Both inorganic component and organic ligands can be selected among a huge class of available moieties and assembled, following a lego-like approach, to tailor pores size, light absorption properties, band gap width, as well as nature and number of active sites. Furthermore, their huge surface area and tunable pore size and distribution promote substrate-catalyst interaction, providing at the same time size and shape selectivity towards the desired reaction [1,214,215]. Owing to these superior features, MOFs arise great interest in many photocatalytic processes, including hydrogen evolution [216,217,218], CO2 reduction [175,219,220,221], and pollutant degradation [222,223,224], as well as organic selective oxidations [225]. In most MOFs, inorganic metal nodes are the active component towards photo-redox reactions, exhibiting photoelectronic properties, typical of inorganic semiconductors (Figure 10a), yet with the better photocatalytic performance [1]. Actually, they can be considered ‘quantum-dot’ semiconductors connected by organic ligands. Due to their nanometric size, charge carriers cover short distances to reach the surface, thus limiting recombination phenomena [1,221]. Furthermore, photo-responsive molecules can be directly employed to build the network, and, due to overlap of electronic states, they can act as molecular antennas to improve visible light-harvesting properties of the semiconductor nodes (Figure 10b), according to a dye sensitizing or a ligand-to-metal charge transfer mechanism (see Section 2 and Section 3). Alternatively, organic binders can take part in redox processes through π-π * transition mechanism (Figure 10c). Finally, the presence of semiconductor nodes and photoactive dye molecules into the same MOF allows dual excitation phenomena of both components, as well as inter-ligand charge transfer, due to the overlapping of frontiers orbitals, thus boosting photocatalytic activity (Figure 10e) [1].
Moreover, the presence of a great number of functional groups and coordination sites, as well as porous structure, allow incorporation of foreign photoactive species, metals, or dyes, which further improve photocatalytic properties and confer multifunctionality [1,215,221]. Indeed, despite the huge potential, MOFs performances in photocatalytic processes are still limited by poor absorption in the visible range, as well as charge recombination phenomena. To date, different strategies have been proposed to improve the photocatalytic activity of these systems under solar light irradiation [12,221]. These include: (i) high absorption photoactive ligands, (ii) ligand functionalization, (iii) dye sensitization, (iv) engineering of inorganic building units, and (v) encapsulation of photocatalytic active species (MOF composites photocatalysts). A summary of the most promising strategies and systems developed thereof is reported in Table 5.

5.1. High Absorption Photoactive Ligands

Using ligands with highly efficient light absorption as organic building blocks is the most effective approach to increase light-harvesting properties and decrease charge recombination phenomena. Notably, high photo-responsive ligands in the visible light range include porphyrins/metallo-porphyrins [226,269], moieties containing specific functional groups (–NH2, –NO2, –SH, and –OH) [227], conjugated systems [228], as well as metal complexes [12]. Obtained systems are usually active under visible light irradiation, allowing higher performance than bare TiO2 under UV-light excitation. Porphyrins, metallo-porphyrins, and metal complexes have been widely employed to build MOFs with efficient visible light harvesting. MOF PCN-22, obtained by assembling Ti–O carboxylate clusters with tetrakis(4-carboxyphenyl)porphyrin (TCPP), shows a wide absorption range from 200 nm to 640 nm, as well as a narrow bandgap (1.93 eV). Similar features are disclosed by zirconium-porphyrin MOF, PCN-222, (also named as MOF-545 or MMPF-6). In both systems, the dye acts as an antenna for the metal nodes (Figure 7) and allows both high charge separation and intensive light absorption. This results in high photocatalytic yield under visible light conditions (Figure 11), even higher than those measured on UiO-66-NH2 and MIL-125-NH2 under similar conditions [229,230].
Bipyridyl-containing cadmium–MOFs, Cd(dcbpy) show high light-harvesting properties as well as high activity and selectivity towards photo-oxidation of benzylamine to N-benzylbenzaldimine under visible light irradiation at room temperature, without any photosensitizer or co-catalyst [231]. Actually, they show higher activity than conventional inorganic semiconductors, such as TiO2, Nb2O5, WO3, even doped with noble metals, under similar conditions. Ligand-to-metal charge transfer mechanism enables MOFs photoactivity under solar irradiation. This occurs in zinc-imidazolate MOF (ZIF-8), disclosing high antibacterial activity and huge potential in air pollution remediation and personal care (Figure 12) [232]. Antibacterial performance of ZIF-8 is far higher than highly active photocatalysts MOF MIL-100(Fe), NH2-MIL-125(Ti), and NH2-UIO-66(Zr) in photoinduction.
Metal conjugation to aromatic ligands markedly enhances their intrinsic visible light absorption [12]. Introducing Cu2+ cations into the porphyrin ring of porphyrin Zr MOF PCN 224 boosts light absorption at 660 nm and, at the same time, hinders electron-hole recombination, ultimately enhancing photocatalytic yield. Obtained systems can be efficiently employed as antibacterial formulations in wound healing [226]. Notably, single Pt atom coordinated porphyrin has been employed as a ligand to produce MOF nanosheets with an impressive H2 production rate of 11,320 μmol g−1 h−1 through water splitting under visible range [218]. Similarly, a Ru(II) complex with bis(4′-(4-carboxyphenyl)-terpyridine), (Ru(tpy)2), has been exploited for the synthesis of Ti-Ru(tpy)2 MOF, which shows intensive absorption and, consequently, enables photocatalytic hydrogen production through photoreforming in the visible light region [233].
Organic ligands, containing unsaturated carbons and conjugated species, can also play a key role in modulating the bandgap and, consequently, MOFs’ photocatalytic properties. As a proof of concept, using π-conjugated 1,4-bis(2-[4-carboxyphenyl]ethynyl) benzene (cpeb) as a linker for Zr6O4(OH)4(CO2)12 clusters leads to a Zr-based MOF (VNU-1) with absorption edge in the visible range, at about 540 nm, and higher photocatalytic activity than Degussa P-25, as well as UiO-66(Zr) and UiO-67(Zr) MOFs, with the same structure [228]. Similarly, 2,5-dihydroxyterephthalic acid employed as a linker to build Ti-based MOF NTU-9 significantly influences both the valence band (VB) and the conduction band (CB), resulting into a wide absorption in the visible range up to 750 nm and small bandgap (1.72 eV) [227].
An effective approach to design highly photoactive MOFs consists in the simultaneous use of two or more linkers. This strategy is useful to include bulky metalloorganic ligands [12] into the framework and allows tuning porosity and functions of obtained MOFs. The mixed-ligand strategy was applied to incorporate porphyrin into an imidazole-functionalized Zr-based MOF, enlarging visible absorption up to 750 nm [234]. A mixed linker Zr-MOF was also synthesized by assembling 2D layers, composed of Zr6 clusters and primary linker benzene tribenzoate (BTB), with TCPP porphyrin. The obtained MOF revealed a 3D layer-pillar structure, intensive absorption in a wide range of the solar spectrum, as well as high photocatalytic activity towards diclofenac degradation [270].

5.2. Ligand Functionalization

MOFs photo-electronic properties can be improved by the modification of organic ligands. This can be carried out either during synthesis in-situ or following a post-synthetic approach [12,242]. Actually, electron-donating/withdrawing features of substituting groups influence the energies levels of both valence and conduction bands and, consequently, optical features as well as redox potentials. Furthermore, they can also affect charge recombination phenomena. As a general rule, electron-donating substituents can extend the absorption range to the visible light region, thus promoting photocatalytic performance [12]. Notably, semiconducting features are strongly influenced by the resonance effects in the organic ligands [271]. Among the investigated substituents, a huge number of studies prove the great efficacy of amino groups in extending MOFs absorption to the visible range. Particularly, amino-functionalized Ti-MOFs, including NH2-MIL-125(Ti), disclose high performance in a great number of photocatalytic processes, including water purification [235], H2 production [236], CO2 reduction [272], as well as organic synthesis [237,238].
Recent studies proved that, under light irradiation, photogenerated electrons could migrate to Ti nodes, whereas holes are localized on the amino group (Figure 10b); this hinders charge recombination and promotes photocatalytic performance [273]. In amino-functionalized Fe-based MOFs, the ligand to metal electron transfer mechanism coexists with charge separation phenomena in semiconductor Fe-O clusters under visible light (Figure 10e), concurring to impressive photocatalytic activity towards CO2 as well as Cr(VI) reduction [12,239]. Adding a higher number of amino functional groups can further shift the light absorption of MOFs to longer wavelengths, but it decreases the oxidation power due to the shift of VB potential to more negative values [274].
As reported by recent studies, even amide functionalization can improve photocatalytic properties [240]. Moreover, ligand functionalization can be carried out to anchor metal ions with catalytic activity. In this way, Zr-based MOFs were modified, including thiol groups that acted as binding sites for Co(II) ions. Obtained Co-thiolate moieties acted as co-catalysts toward efficient and selective CO2 reduction under visible light [241]. A novel promising approach to produce visible-light responsive MOFs consists of post-synthesis covalent modification with a covalent organic framework (COF). This strategy was successfully applied to NH2-MIL-125(Ti) or NH2-UiO-66(Zr) systems, which were modified with triazine-based frameworks (B-CTF-1). Compared to bare MOFs and even physical mixtures of MOFs and COFs, obtained MOF/COF hybrid materials showed higher stability, as well as hydrogen production rate, due to the presence of amide bonds between B-CTF-1 and MOFs, ensuring wide absorption range as well as long-lived photogenerated charge carriers [225,242], as shown in Figure 13.
The coating of NH2-MIL-125 with shell via a covalent bond linkage modifies the intrinsic electronic properties and reduces the recombination rate of photogenerated electron-hole pairs, resulting in excellent performances. Introducing a dye into the semiconductor-like network usually shifts its absorption edge to the visible range and further improves the photocatalytic activity. This approach can be implemented through post-synthesis conjugation using amino groups in the ligand [243]. Following this strategy, Ru(II)-polypyridyl complexes were inserted into MIL-125 MOF, obtaining a stable system with high photoactivity for the selective oxidative reactions of amines [244]. As a more promising alternative, dye sensitization can be carried out during the synthesis process (in-situ strategy). Following this approach, erythrosin B dye was included in Zr-based MOFs (UiO-66), which were not active in the visible range. Contrarily, obtained dye-sensitized Pt-UiO-66(Zr) systems showed enhanced photocatalytic activity towards H2 production under visible light, owing to the charge transfer from excited erythrosin to the MOF and subsequent electron migration to Pt NPs [245].

5.3. Engineering of Inorganic Building Units

MOFs photocatalytic properties can be also tuned through an accurate design of the inorganic cluster. Notably, metal ion substitution in the inorganic node can improve light harvesting and depress charge recombination phenomena, thus enhancing photocatalytic performance. In this regard, partial replacement of Zr with Ti atoms in NH2-MIL-125(Ti) MOF results in high activity for remediation of pharmaceutical pollutants under solar irradiation [246].
Ce(III) doping into the MIL-101(Cr) network promotes charge migration through Ce4+/Ce3+ redox pair and allows hydrolysis of borazane under visible light [247]. Alternatively, in MOFs-based semiconductors, even the size of inorganic building blocks can be exploited as a parameter to tune photoelectronic properties. Notably, larger assembling units account for smaller band gaps and allow photocatalytic response in the visible region [12].

5.4. Encapsulation of Photocatalytic Active Species (MOF Composites Photocatalysts)

Encapsulating metal, metal-containing species, or carbon nanomaterials into MOF frameworks can further improve their photocatalytic properties and is currently considered one of the most promising strategies to design high-performance photocatalysts. Actually, MOF nanocomposites combine the advantages of both host nanostructures (optical, catalytic, electrical, magnetic, etc.) and MOFs matrix (high surface, tunable ordered structure, etc.), present multifunctional behavior, and often unravel unique properties due to intimate interaction between the components. Noble metal nanoparticles (MNPs) play a key role in enhancing MOFs photocatalytic properties [1]. Intrinsic properties of MOFs, such as high surface area, allow fine dispersion of MNPs and avoid their aggregation. Due to surface plasmon resonance (SPR) effects, MNPs can improve visible light absorption. Moreover, they can accept photogenerated electrons, thus limiting charge recombination [1,221], ultimately leading to superior catalytic performance.
Pt nanoparticles can be deposited on ZrO2-based MOF by atomic layer deposition (ALD), increasing the photocatalytic activity. This could be ascribed to the promotion of charge separation phenomena and limitation of electron/hole recombination [248]. A combination of photoactive ligands with MNPs significantly further promotes photocatalytic performance. This occurs in UiO-66-NH2 and even in MIL-125(Ti)−NH2 MOFs, whose already relevant properties can be improved by encapsulating Pd [249], Pt [250], or Au [251]. Accepting photogenerated electrons from inorganic semiconductor clusters, Pt and Au can also act as co-catalysts in redox processes [272]. As an alternative to noble metal nanoparticles, doping with less expensive metals, such as Cu, can also significantly enhance both stability and photocatalytic activity, as a result of the plasmonic phenomenon and metal-semiconductor junction, improving charge mobility and depressing recombination (Figure 14) [252].
Active metal species can be also encapsulated as molecular catalysts to enhance MOF photocatalytic efficiency. In this regard, Ni@Ru-UiO-67, as well as Ag@Re-UiO-67, systems showed efficient CO2 conversion to CO under visible light [253,254].
Another valid approach to reduce charge recombination rate and improve MOFs photocatalytic performance is based on their coupling with semiconductor nanoparticles, such as TiO2, ZnO, CdS, C3N4 [255]. Cu(II)-porphyrin and TiO2 nanoparticles into MOF structure (PCN-224(Cu)) leads to a marked improvement of light-harvesting features and promotes separation of photogenerated charges, reaching superior photocatalytic performance for CO2 reduction [219,256]. Also, composites or heterostructures of titanium oxide or titanate fibers and MOFs, such as TiO2@NH2-MIL-88B(Fe), disclose high photocatalytic activity [257,258,259]. Recently, novel heterojunctions, based on bismuthoxyhalide-MOFs composites, were successfully produced by controlled growth of BiOX (X=Cl, Br) nanosheets on the surface of NH2-UiO-66 and NH2-MIL-125 (Ti). These composites showed impressive photocatalytic activity under solar irradiation due to the absorption of visible light, as well as the efficient charge separation and transfer at the interface between BiOX and MOF components [260,261,275]. Less conventional nanostructures, such as silver vanadate (Ag3VO4) and silver tungstate (Ag2WO4) nanoparticles, can be included in Ti-containing MOF (MIL-125-NH2) to enhance its photocatalytic performance, which proved even better than previously investigated samples [262]. CdS, as well as MoS2, have been raising great interest for photocatalytic processes due to their low bandgap (around 2.4 eV) that enables activity under visible light, but they suffer from poor stability caused by photo-corrosion. Their inclusion into a MOF framework, however, allows a fine dispersion of nanoparticles, avoids aggregation, improves stability, promotes sunlight absorption and photo-generated electron-hole exchange between the components, and ultimately improves photocatalytic activity [263,264].
Similarly, including lead halide perovskite quantum dots (QDs) into Fe-porphyrin-based MOF confers stability in water-containing systems. Furthermore, the close interaction between QDs and the Fe catalytic site in the MOF allows the fast exchange of photogenerated electrons, enhancing the photocatalytic activity towards CO2 reduction, which was 38 times higher than on the bare components [265]. Apart from inorganic semiconductors, integrating MOFs with conductive carbon materials is an effective approach to promote the migration rate of photogenerated carriers [12,221]. For example, composites obtained combining activated carbon (AC) with MIL-88B (Fe) MOF showed higher photocatalytic activity than bare components [266]. Furthermore, reduced graphene oxide (RGO) can be coupled to MOF frameworks, such as UiO-66(NH2) [267] and RGO-MIL-53(Fe) [268] to improve their photocatalytic properties. Interestingly, incorporating both semiconductor nanoparticles, such as CdS quantum dots, and RGO nanosheets into a MOF appears as a promising strategy to (i) expand absorption properties of the MOF to visible light, (ii) match heterojunction bandgap, (iii) allow selective adsorption to enhance catalytic activity, (iv) protect CdS nanoparticles from leaching. All these features promote sunlight harvesting, photo-generated charge migration, as well as selectivity and, as a result, the photocatalytic activity [223]. As a general rule, the interfacial between MOFs and nanostructures, semiconductors, and RGO in MOF nanocomposites plays a key role in defining the photocatalytic behavior.
Certainly, many strategies for MOFs design have been developed (as summarized in Table 5) to increase the photocatalytic activity under visible irradiation or solar light irradiation. Several applications have been considered: significant performances (in the range of 80%–100% after not-long time) have been shown in the photodegradation of organic dyes, while promising H2 productivity values (between 0.75–10000 μmol·h−1·g−1) have been obtained. Therefore, new approaches must be explored to design heterojunctions with intimate interaction and bandgap matching with the aim to increase the photocatalytic activity.

6. Conductive Polymers-Based Hybrid Nanomaterials

Conductive polymers are very interesting organic macromolecules, widely considered for the design of advanced hybrid materials with photocatalytic properties [276], thanks to their good processability, especially through solution processes [277]. Conductive polymers can provide matched band structures with other inorganic semiconductors, such as metal oxides nanostructures, reducing the recombination of photogenerated electron-hole pairs for obtained hybrid composite photocatalysts [278,279]. More specifically, the loading of conductive polymers improved the photostability of semiconductor photocatalysts, also inhibiting their photo-corrosion. At the same time, the presence of π-conjugated long chains in the conductive polymers makes them able to trap the photoinduced holes, thus hindering the self-oxidation of the same photocatalysts. On the other hand, a conductive polymer combined with a semiconductor photocatalyst has shown different roles under different light irradiation: under UV-visible light irradiation, the conductive polymer can act as a co-catalyst, which is excited by irradiated light and a hole transporting material, while under visible light irradiation, only the conductive polymer can be excited as a photosensitizer [277,279]. The most used conductive polymers, which have been loaded into various semiconductor metal oxides, are polyaniline (PANI), polypyrrole (PPy), poly(3,4-ethylenedioxythiophene) (PEDOT), and polythiophene (PTh), schematized in Figure 15.
Many different nanostructured materials have been systematically produced by using various preparation methods, determining different morphologies and photocatalytic performances. Particularly, the combination of conductive polymers and semiconductor metal oxides determined an enhancement in the separation of photoexcited charge carriers, also extending the light absorption range, increasing the adsorption of reactants (Figure 16), inhibiting photo-corrosion, and reducing the formation of large aggregates. Consequently, the synergic effect between the conductive polymers and the semiconductor nanomaterials resulted in a significant improvement of the visible-light-driven photocatalytic activity of these hybrid nanostructured materials, which have been principally used for the degradation of hazardous chemicals as well as hydrogen production [277,279]. The main properties of the semiconducting polymers/metal oxide nanostructured photosensitive hybrid materials considered in this review are summarized in Table 6.
For example, a large portion of the conductive polymers-based hybrid photocatalysts has been prepared by using polyaniline (PANI), highly considered for its high conductivity, simple preparation, and reasonable stability [280,281]. PANI has been combined with many semiconductor metal oxides, such as TiO2 [282,283], ZnO [284,285,286], ZrO2 [287], CoFe2O4 [288], or SnO2 [289] to prepare the composite photocatalysts developed to address the issue of water contamination by different mono- and multivalent heavy metal ions in recent years. A schematic illustration of the action mechanism of ZnO/PANI is represented in Figure 17.
Another important conductive polymer is represented by polypyrrole (PPy), mostly used to produce nanohybrid photocatalysts for photodegradation of organic pollutants, and the photocatalytic performances of PPy-based materials have been affected by the fabrication method and type of irradiated light [277]. Many examples of hybrid nanomaterials with photocatalytic properties have been recently reported, in which PPy has been combined with different semiconductor metal oxides, such as TiO2 [290,291], for pollutants degradation or hydrogen production, ZnO [292,293,294], and its derivatives doped with other metals, ZnWO4 [295], Fe2O3 [296], or Fe3O4 [297,298,299].
Poly(3,4-ethylenedioxythiophene) (PEDOT) is a poorly soluble polymer characterized by excellent electronic and optical properties, good optical transparency in its conducting state, high stability, and moderate bandgap, as well as a low redox potential [300]. Thanks to these properties, PEDOT has been reported to demonstrate high electrocatalytic activities for the oxygen reduction reaction [301], as well as an efficient catalyst in hydrogen generation and storage [302,303]. For this reason, PEDOT has been combined with many semiconducting metal oxides, such as TiO2 [304], ZnO [305], or a combination of both metal oxides [306], determining an increase in the photocatalytic activity.
Similarly, polythiophene (PTh) has also attracted much great attention, as one of the most promising conducting polymers, able to significantly enhance the photocatalytic activity of the most common semiconducting inorganic photocatalysts. Indeed, PTh has been observed to be capable to increase the adsorption ability of the inorganic catalyst, also widening the light absorption region towards the visible light. However, only a few polythiophene (PTh)-based composite photocatalysts have been actually reported and principally used for the photodegradation of organic pollutants and the CO2 photoreduction [307].
Definitely, the incorporation of conductive polymers has been demonstrated to significantly improve the visible-light-driven photocatalytic activity of hybrid nanostructured materials, by enhancing the separation of photogenerated charge carriers and extending the light absorption range. Based on these significant advantages, many nanostructured hybrids composed by semiconductor oxides conjugated with conductive polymers in different heterostructures have been designed, showing a promising photocatalytic activity, principally for water purification. Very high efficiencies in the degradation of organic molecules have been observed, with productivity values >80%–90%, and, in many cases, of a complete degradation within 1 or 2 h. Furthermore, the principal advantages for the use of photoactive polymeric-based materials are easy removal and long life, and control of the formation of secondary contamination is avoided. However, further steps appear necessary to better understand the mechanisms of action of these materials and their physicochemical features, with the aim to better improve the photoactivity, principally against the more insidious organic pollutants, as well as to extend their application field.

7. Hybrid Perovskites

Organo-inorganic hybrid perovskites have been largely explored to carry out semiconductor-based photocatalytic processes. In these compounds, the inorganic component is identified by the general formula ABX3, where A and B are transition metal cations, and X is a non-metallic anion, usually a halogen/oxygen anion. Actually, they hold great potential to harness solar light efficiently due to their peculiar photoelectronic properties, including strong light absorption, low exciton binding, as well as long photo-excited charges lifetime and migration. Moreover, the composition of both inorganic (A, B, and O) and organic components can be adjusted to tune bandgap and optimize photocatalytic efficiency towards the desired reaction [308,309,310]. Particularly, Pb-based organohalide perovskites, such as methylammonium lead iodide (MAPbI3, MA = methylammonium), exhibit excellent photocatalytic properties due to marked light absorption, fast charge carrier migration properties, and poor recombination phenomena [309,311,312]. These systems have been deeply investigated in the last years as light-absorbing layers in perovskite-based solar cells, an emerging photovoltaic technology, which rapidly achieved record efficiencies, and showed promising performances in the fields of hydrogen production and organic synthesis (Figure 18) [312].
Nevertheless, their low-stability in an aqueous environment and lead-induced high toxicity strongly limit their applications [311,313]. Several strategies have been proposed to improve the stability of these systems. Notably, hybrid mixed halide perovskites—MAPb(I1−xBrx)3 [314] and MAPbBr3−xIx [315]—prepared by one-pot crystallization or post-synthesis ion exchange in a mixed halide solution, combine enhanced stability with improved photocatalytic activity for H2 evolution in aqueous HX solution (X = I, Br), under visible light (λ ≥ 420 nm). Alternatively, stable photocatalytic hydrogen iodide (HI) splitting to H2 can be achieved, exploiting the dynamic equilibrium of the dissolution and precipitation of MAPbI3 [316]. Isolating MAPbX3 perovskites from solvent molecules can be an effective method to improve their stability.
To this purpose, bulky capping ligands, such as aniline, can be bound to the surface of MAPbBr3 system, exposing outwards hydrophobic phenyl groups, and, due to steric hindrance, protect the perovskite from corrosion, enhancing its stability [308]. Similarly, large conjugated organic cations can be employed as templating agents to produce stable lead halides, with narrow band gaps (2.01–2.35 eV) and, consequently, excellent photocatalytic activity and photochemical stability [317]. Also, the photocatalytic activity can be further improved by doping with additional co-catalysts [310]. In this way, composites of MAPbI3 with reduced graphene oxide (rGO) [318,319], TiO2 [320,321], or Ni3C [322] are able to promote charge carrier transfer and separation and ensure a huge number of reactive centers on the surface. These features boost the efficiency of photocatalytic processes; actually, Ni3C/MAPbI3 composites disclosed hydrogen evolution rate as high as 2362 μmol g−1 h−1, about 55-fold higher than that of bare MAPbI3 (43 μmol g−1 h−1) and even higher than that of Pt/MAPbI3 (534 μmol g−1 h−1) (Figure 19). Table 7 reports a summary of the most promising hybrid perovskites and their photocatalytic features, as considered in this review.
In order to address toxicity issues, environmentally-friendly lead-free perovskites have been designed [309,311,323]. Among those, bismuthate perovskite, (CH3NH3)3Bi2I9, have been rising great interest due to excellent stability in reaction environment, as well as a marked photocatalytic activity under visible light towards H2 production [309] and water contaminants degradation (Figure 20) [323].
Finally, hybrid perovskites have been produced by different synthetic approaches and have shown outstanding photocatalytic performances, principally for H2 production (in the range between 10–7000 μmol·h−1·g−1). Also, some environmental tests have been realized using hybrid perovskites to photodegrade the most common organic molecules, showing productivity values comparable with those obtained by the previously described hybrid nanomaterials.

8. Bioinspired Z-Scheme Photocatalysts

Recently, another feasible strategy for enhancing photocatalytic performance is spreading through the building of the direct Z-scheme photocatalysts, which are conceived based on biomimetic artificial photosynthesis [324,325]. Indeed, these materials owe their name to the fact that their charge transfer mechanism is similar to natural photosynthesis in green plants, in which the charge-carrier transport pathway involves a two-step photoexcitation that looks like the English letter “Z” (Figure 21a). Since 2013, when the concept of a direct Z-scheme photocatalyst was proposed for explaining the high photocatalytic formaldehyde (HCHO) degradation performance of the TiO2/g-C3N4 composite [326], huge activities aimed to design and develop photocatalysts based on the direct Z-scheme have been conducted by several research groups.
The reactions in the natural photosynthesis are driven by the photogenerated electrons (in the CB of PS I), which can maintain a high reducing ability, and the photogenerated holes (in the VB of PS II), which can sustain a high oxidizing ability. Bio-inspired by the natural mechanism, an artificial Z-scheme photocatalyst has huge potential in achieving high photocatalytic performance. Specifically, the structure of a direct Z-scheme photocatalyst (Figure 21b) is similar to that of a type II heterojunction photocatalyst, but its charge carrier migration mechanism is different. An artificial Z-scheme photocatalyst consists of two linked semiconductor photocatalysts. The former is an oxidation photocatalyst, which possesses a low VB position and exhibits a strong oxidation ability; the latter is a reduction photocatalyst, which usually has a high CB position, displaying strong reduction ability. Thanks to the peculiar charge transfer, a Z-scheme photocatalyst with excellent photocatalytic activity possesses several advantages: (i) simultaneous conservation of strong reduction and oxidation capabilities; (ii) spatial separation of reductive and oxidative active sites; (iii) high separation efficiency of photogenerated charge carriers with strong redox power; (iv) extended light-harvesting range, and (v) a wide spectrum of photocatalysts for specific photocatalytic reactions, indicating that coupling two photocatalysts with narrower band gaps meets the requirement of higher redox ability for these reactions [325,327].
Nanostructured hybrid materials, combining inorganic and organic components at the molecular scale, allow intimate interaction at the interface, as well as fine tuning of photoelectronic properties required to achieve ideal bandgap matching; thus, they are ideal candidates to build Z-scheme photocatalysts. Actually, most of the previously described systems can be efficiently combined into a nanocomposite featuring direct Z-scheme heterojunction. Notably, some of them, such as carbon-based materials (graphene oxide, reduced graphene oxide, and especially graphitic carbon nitride) and MOFs, can act as electron mediators required in a solid-state Z-scheme mechanism [219,328,329,330,331].
Z-scheme photocatalysts are, therefore, intensively studied in the last few years and are promising in various applications; the latter include water splitting, carbon dioxide reduction, degradation of pollutants, and biohazard disinfection. The selection of suitable semiconductors with appropriate band structures is a prerequisite for the successful fabrication of direct Z-scheme photocatalyst. However, although they have been widely investigated until now, the obtained photocatalytic efficiency is still moderate and far from the requirement for practical applications. Many studies are still needed to achieve the real application of artificial photosynthesis.

9. Conclusions

The fields of semiconductor-based solar light harvesting for photocatalytic and related applications are rapidly evolving in various directions. The combination of inorganic semiconductors with organic photoactive compounds, polymers, or C-based materials is showing great potential, inducing emerging properties in the resulting hybrid heterostructures, depending on the physicochemical and electronic interaction between the components. Among the advantages offered by hybrid materials, extended light absorption and improved charge separation efficiency and lifetime are frequently obtained, along with the possibility to match the conduction and valence band positions, achieving suitable redox potential for specific reactions. Currently, TiO2 remains the reference photocatalyst, followed by ZnO; however, other semiconductors are revealing promising photochemical properties, which deserve further insight and could be improved through the mentioned approaches.
Dye sensitization is a well-established strategy, offering a huge variety of molecular structures and optoelectronic properties. Much research is focused on seeking effective metal-free dyes, possibly inexpensive or deriving from natural sources. Ligand-to-metal charge transfer sensitization seems promising but relatively less developed and opens the way to progress. Its main advantage is the possibility to use small molecules that are cheaper and easier to synthesize than large dye compounds. The stability and reusability are crucial points for this kind of sensitized photocatalysts and can be faced by carefully choosing the synthesis route and conditions and the ligand structure, in order to obtain stable binding of the sensitizers, possibly through covalent bonds and/or multiple anchoring groups.
Graphene, reduced graphene oxide, C nanotubes, C quantum dots, graphitic C3N4, and other carbon materials provide huge advantages in conjugation with metal-based semiconductors, allowing to design a wide variety of hybrid nanocomposites with tunable properties. They can act as sensitizers and/or electron acceptors, reducing recombination and enhancing electron transport and light conversion efficiency. TiO2, ZnO, CuxO, and Fe oxides are widely explored in this scope. Conductive polymers offer similar benefits and may turn useful in the preparation of stable and recyclable photocatalysts, additionally serving as supports for oxide nanostructures.
In the last years, metal-organic frameworks (MOFs) are emerging as multifunctional materials, with several combinations of metals and linkers offering tailored structures and properties. Many examples of MOFs with outstanding photocatalytic performances are already available, but their potential is still to be fully exploited. Hybrid perovskites, intensely studied in solar cells, are also attracting interest for their radiation harnessing ability.
As regards the applications of hybrid photocatalysts, the removal of water pollutants is widely investigated, though the most used substrates are still dyes and simple model organic molecules, while there are limited focalized studies on the degradation mechanisms of more complex contaminant compounds of emerging concern. An impressive number of works deal with hydrogen generation through water splitting, one of the main future challenges, in which further advances are expected. The production of solar fuels through CO2 reduction is also receiving growing attention but needs strong improvements. On the other hand, an attractive field that seems to have still much to disclose is organic synthesis. Apart from photocatalytic selective oxidations, few organic conversion reactions have been explored in detail on hybrid photocatalytic systems. Limited data are available also on the light-induced antimicrobial activity of hybrid photosensitive materials, which is an application with a strong practical interest. Moreover, the feasibility of the scale-up from the laboratory to a chemical plant scale, which is a fundamental step toward the practical application, is rarely considered.

10. Perspectives

The sunlight harvesting represents an ambitious objective of ongoing research efforts. Hybrid organic-inorganic sunlight-capturing nanomaterials will attract a great deal of attention due to their unique ability to manipulate light and control energy flow at nearly the atomic level. However, there are still many challenges need to be addressed before they are fully used in practical applications with high commercialization. Indeed, to make solar energy economically competitive with other energy sources, it is critical to find ways to reduce the costs and, at the same time, to improve efficiency and performances. A huge variety of novel photocatalytic materials is being proposed; however, a detailed assessment of their electronic and physicochemical properties and their functional behavior is necessary. Particular attention should be devoted to the stability of the photocatalysts in the reaction environment, which is often a limitation to their long-term activity and recycling efficiency, and the reproducibility of both synthetic procedure and catalytic performances.
For this reason, it can be expected that the current knowledge on the various classes of photoactive systems will inspire the researchers’ creativity for the design and development of novel hybrid materials, even by the combination of some of the presented systems, and relying on less known promising semiconductors. The aim will be the optimization of tailored materials for the utilization of solar radiation energy in specific processes in energetic, environmental, and industrial fields, eventually making these photocatalytic processes competitive with traditional technologies. Furthermore, many types of hybrid systems attract interest also in other fields, which request nanomaterials, showing a very broad photo-response or a specific sensitivity to some radiation ranges, e.g., photovoltaic cells and photodynamic therapy. The flexible and tunable optical and electrochemical properties offered by dyes, MOFs, carbon materials, conductive polymers, and their combinations may turn increasingly useful for these applications as well.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhang, T.; Jin, Y.; Shi, Y.; Li, M.; Li, J.; Duan, C. Modulating photoelectronic performance of metal–organic frameworks for premium photocatalysis. Coord. Chem. Rev. 2019, 380, 201–229. [Google Scholar] [CrossRef]
  2. Wang, Z.; Li, C.; Domen, K. Recent developments in heterogeneous photocatalysts for solar-driven overall water splitting. Chem. Soc. Rev. 2019, 48, 2109–2125. [Google Scholar] [CrossRef]
  3. Zhang, P.; Lou, X.W. Design of Heterostructured Hollow Photocatalysts for Solar-to-Chemical Energy Conversion. Adv. Mater. 2019, 31, 1900281. [Google Scholar] [CrossRef]
  4. Gao, C.; Wang, J.; Xu, H.; Xiong, Y. Coordination chemistry in the design of heterogeneous photocatalysts. Chem. Soc. Rev. 2017, 46, 2799–2823. [Google Scholar] [CrossRef]
  5. Christoforidis, K.C.; Fornasiero, P. Photocatalytic Hydrogen Production: A Rift into the Future Energy Supply. ChemCatChem 2017, 9, 1523–1544. [Google Scholar] [CrossRef] [Green Version]
  6. Cao, S.; Piao, L.; Chen, X. Emerging Photocatalysts for Hydrogen Evolution. Trends Chem. 2019, 2, 57–70. [Google Scholar] [CrossRef]
  7. Xu, C.; Anusuyadevi, P.R.; Aymonier, C.; Luque, R.; Marre, S. Nanostructured materials for photocatalysis. Chem. Soc. Rev. 2019, 48, 3868–3902. [Google Scholar] [CrossRef] [PubMed]
  8. Luo, J.; Zhang, S.; Sun, M.; Yang, L.; Luo, S.; Crittenden, J.C. A Critical Review on Energy Conversion and Environmental Remediation of Photocatalysts with Remodeling Crystal Lattice, Surface, and Interface. ACS Nano 2019, 13, 9811–9840. [Google Scholar] [CrossRef] [PubMed]
  9. Truppi, A.; Petronella, F.; Placido, T.; Striccoli, M.; Agostiano, A.; Curri, M.L.; Comparelli, R. Visible-light-active TiO2-based hybrid nanocatalysts for environmental applications. Catalysts 2017, 7, 100. [Google Scholar] [CrossRef]
  10. Zhang, G.; Kim, G.; Choi, W. Visible light driven photocatalysis mediated via ligand-to-metal charge transfer (LMCT): An alternative approach to solar activation of titania. Energy Environ. Sci. 2014, 7, 954–966. [Google Scholar] [CrossRef] [Green Version]
  11. Zhou, N.; López-Puente, V.; Wang, Q.; Polavarapu, L.; Pastoriza-Santos, I.; Xu, Q.-H. Plasmon-enhanced light harvesting: Applications in enhanced photocatalysis, photodynamic therapy and photovoltaics. RSC Adv. 2015, 5, 29076–29097. [Google Scholar] [CrossRef]
  12. Dhakshinamoorthy, A.; Li, Z.; Garcia, H. Catalysis and photocatalysis by metal organic frameworks. Chem. Soc. Rev. 2018, 47, 8134–8172. [Google Scholar] [CrossRef] [PubMed]
  13. Reddy, K.R.; Hassan, M.; Gomes, V.G. Hybrid nanostructures based on titanium dioxide for enhanced photocatalysis. Appl. Catal. A Gen. 2015, 489, 1–16. [Google Scholar] [CrossRef]
  14. Hagfeldt, A.; Boschloo, G.; Sun, L.; Kloo, L.; Pettersson, H. Dye-Sensitized Solar Cells. Chem. Rev. 2010, 110, 6595–6663. [Google Scholar] [CrossRef]
  15. Zhang, X.; Peng, T.; Song, S. Recent advances in dye-sensitized semiconductor systems for photocatalytic hydrogen production. J. Mater. Chem. A 2016, 4, 2365–2402. [Google Scholar] [CrossRef]
  16. Youssef, Z.; Colombeau, L.; Yesmurzayeva, N.; Baros, F.; Vanderesse, R.; Hamieh, T.; Toufaily, J.; Frochot, C.; Roques-Carmes, T. Dye-sensitized nanoparticles for heterogeneous photocatalysis: Cases studies with TiO2, ZnO, fullerene and graphene for water purification. Dyes Pigment. 2018, 159, 49–71. [Google Scholar] [CrossRef]
  17. Diaz-Angulo, J.; Arce-Sarria, A.; Mueses, M.; Hernandez-Ramirez, A.; Machuca-Martinez, F. Analysis of two dye-sensitized methods for improving the sunlight absorption of TiO2 using CPC photoreactor at pilot scale. Mater. Sci. Semicond. Process. 2019, 103, 104640. [Google Scholar] [CrossRef]
  18. Park, H.; Kim, H.; Moon, G.; Choi, W. Photoinduced charge transfer processes in solar photocatalysis based on modified TiO2. Energy Environ. Sci. 2016, 9, 411–433. [Google Scholar] [CrossRef] [Green Version]
  19. Cecconi, B.; Manfredi, N.; Montini, T.; Fornasiero, P.; Abbotto, A. Dye-Sensitized Solar Hydrogen Production: The Emerging Role of Metal-Free Organic Sensitizers. Eur. J. Org. Chem. 2016, 2016, 5194–5215. [Google Scholar] [CrossRef]
  20. Kumar, S.; Wani, M.Y.; Arranja, C.T.; e Silva, J.D.A.; Avula, B.; Sobral, A.J.F.N. Porphyrins as nanoreactors in the carbon dioxide capture and conversion: A review. J. Mater. Chem. A 2015, 3, 19615–19637. [Google Scholar] [CrossRef]
  21. Jaafar, S.N.H.; Minggu, L.J.; Arifin, K.; Kassim, M.B.; Wan, W.R.D. Natural dyes as TiO2 sensitizers with membranes for photoelectrochemical water splitting: An overview. Renew. Sustain. Energy Rev. 2017, 78, 698–709. [Google Scholar] [CrossRef]
  22. Rochkind, M.; Pasternak, S.; Paz, Y. Using dyes for evaluating photocatalytic properties: A critical review. Molecules 2015, 20, 88–110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Barbero, N.; Vione, D. Why Dyes Should Not Be Used to Test the Photocatalytic Activity of Semiconductor Oxides. Environ. Sci. Technol. 2016, 50, 2130–2131. [Google Scholar] [CrossRef]
  24. Clarizia, L.; Vitiello, G.; Pallotti, D.K.; Silvestri, B.; Nadagouda, M.; Lettieri, S.; Luciani, G.; Andreozzi, R.; Maddalena, P.; Marotta, R. Effect of surface properties of copper-modified commercial titanium dioxide photocatalysts on hydrogen production through photoreforming of alcohols. Int. J. Hydrogen Energy 2017, 42, 28349–28362. [Google Scholar] [CrossRef]
  25. Vitiello, G.; Clarizia, L.; Abdelraheem, W.; Esposito, S.; Bonelli, B.; Ditaranto, N.; Vergara, A.; Nadagouda, M.; Dionysiou, D.D.; Andreozzi, R.; et al. Near UV-Irradiation of CuOx-Impregnated TiO2 Providing Active Species for H2 Production Through Methanol Photoreforming. ChemCatChem 2019, 11, 4314–4326. [Google Scholar] [CrossRef]
  26. Ahmed, M.A.; Abou-Gamra, Z.M.; Medien, H.A.A.; Hamza, M.A. Effect of porphyrin on photocatalytic activity of TiO2 nanoparticles toward Rhodamine B photodegradation. J. Photochem. Photobiol. B Biol. 2017, 176, 25–35. [Google Scholar] [CrossRef]
  27. Han, C.; Shao, Q.; Liu, M.; Ge, S.; Liu, Q.; Lei, J. 5,10,15,20-tetrakis(4-chlorophenyl)porphyrin decorated TiO2 nanotube arrays: Composite photoelectrodes for visible photocurrent generation and simultaneous degradation of organic pollutant. Mater. Sci. Semicond. Process. 2016, 56, 166–173. [Google Scholar] [CrossRef]
  28. Zhang, W.; Wang, C.; Liu, X.; Li, J. Enhanced photocatalytic activity in porphyrin-sensitized TiO2 nanorods. J. Mater. Res. 2017, 32, 2773–2780. [Google Scholar] [CrossRef]
  29. Saien, J.; Mesgari, Z. Highly efficient visible-light photocatalyst of nitrogen-doped TiO2 nanoparticles sensitized by hematoporphyrin. J. Mol. Catal. A Chem. 2016, 414, 108–115. [Google Scholar] [CrossRef]
  30. Sułek, A.; Pucelik, B.; Kuncewicz, J.; Dubin, G.; Dąbrowski, J.M. Sensitization of TiO2 by halogenated porphyrin derivatives for visible light biomedical and environmental photocatalysis. Catal. Today 2019, 335, 538–549. [Google Scholar] [CrossRef]
  31. Sułek, A.; Pucelik, B.; Kobielusz, M.; Łabuz, P.; Dubin, G.; Dąbrowski, M.J. Surface Modification of Nanocrystalline TiO2 Materials with Sulfonated Porphyrins for Visible Light Antimicrobial Therapy. Catalysts 2019, 9, 821. [Google Scholar] [CrossRef] [Green Version]
  32. Lovat, G.; Forrer, D.; Abadia, M.; Dominguez, M.; Casarin, M.; Rogero, C.; Vittadini, A.; Floreano, L. On-Surface Synthesis of a Pure and Long-Range-Ordered Titanium(IV)-Porphyrin Contact Layer on Titanium Dioxide. J. Phys. Chem. C 2017, 121, 13738–13746. [Google Scholar] [CrossRef]
  33. Wechsler, D.; Fernández, C.C.; Steinrück, H.-P.; Lytken, O.; Williams, F.J. Covalent Anchoring and Interfacial Reactions of Adsorbed Porphyrins on Rutile TiO2(110). J. Phys. Chem. C 2018, 122, 4480–4487. [Google Scholar] [CrossRef]
  34. Youssef, Z.; Arnoux, P.; Colombeau, L.; Toufaily, J.; Hamieh, T.; Frochot, C.; Roques-Carmes, T. Comparison of two procedures for the design of dye-sensitized nanoparticles targeting photocatalytic water purification under solar and visible light. J. Photochem. Photobiol. A Chem. 2018, 356, 177–192. [Google Scholar] [CrossRef]
  35. Dai, Z.; Li, D.; Chi, L.; Li, Y.; Gao, B.; Qiu, N.; Duan, Q.; Li, Y. Preparation of porphyrin sensitized three layers magnetic nanocomposite Fe3O4@SiO2@TiO2 as an efficient photocatalyst. Mater. Lett. 2019, 241, 239–242. [Google Scholar] [CrossRef]
  36. Watanabe, M.; Sun, S.; Ishihara, T.; Kamimura, T.; Nishimura, M.; Tani, F. Visible Light-Driven Dye-Sensitized Photocatalytic Hydrogen Production by Porphyrin and its Cyclic Dimer and Trimer: Effect of Multi-Pyridyl-Anchoring Groups on Photocatalytic Activity and Stability. ACS Appl. Energy Mater. 2018, 1, 6072–6081. [Google Scholar] [CrossRef]
  37. Bayrak, R.; Albay, C.; Koç, M.; Altın, İ.; Değirmencioğlu, İ.; Sökmen, M. Preparation of phthalocyanine/TiO2 nanocomposites for photocatalytic removal of toxic Cr(VI) ions. Process. Saf. Environ. Prot. 2016, 102, 294–302. [Google Scholar] [CrossRef]
  38. Gorduk, S.; Avciata, O.; Avciata, U. Photocatalytic degradation of methylene blue under visible light irradiation by non-peripherally tetra substituted phthalocyanine-TiO2 nanocomposites. Inorg. Chim. Acta 2018, 471, 137–147. [Google Scholar] [CrossRef]
  39. Altın, İ.; Sökmen, M.; Bıyıklıoğlu, Z. Sol gel synthesis of cobalt doped TiO2 and its dye sensitization for efficient pollutant removal. Mater. Sci. Semicond. Process. 2016, 45, 36–44. [Google Scholar] [CrossRef]
  40. Huang, J.-F.; Liu, J.-M.; Xiao, L.-M.; Zhong, Y.-H.; Liu, L.; Qin, S.; Guo, J.; Su, C.-Y. Facile synthesis of porous hybrid materials based on Calix-3 dye and TiO2 for high photocatalytic water splitting performance with excellent stability. J. Mater. Chem. A 2019, 7, 2993–2999. [Google Scholar] [CrossRef]
  41. Hashim, N.; Thakur, S.; Patang, M.; Crapulli, F.; Ray, A.K. Solar degradation of diclofenac using Eosin-Y-activated TiO2: Cost estimation, process optimization and parameter interaction study. Environ. Technol. 2017, 38, 933–944. [Google Scholar] [CrossRef] [PubMed]
  42. Chowdhury, P.; Gomaa, H.; Ray, A.K. Sacrificial hydrogen generation from aqueous triethanolamine with Eosin Y-sensitized Pt/TiO2 photocatalyst in UV, visible and solar light irradiation. Chemosphere 2015, 121, 54–61. [Google Scholar] [CrossRef] [PubMed]
  43. Yin, M.; Ma, S.; Wu, C.; Fan, Y. A noble-metal-free photocatalytic hydrogen production system based on cobalt(iii) complex and eosin Y-sensitized TiO2. RSC Adv. 2015, 5, 1852–1858. [Google Scholar] [CrossRef]
  44. Peng, S.; Zeng, X.; Li, Y. Titanate nanotube modified with different nickel precursors for enhanced Eosin Y-sensitized photocatalytic hydrogen evolution. Int. J. Hydrogen Energy 2015, 40, 6038–6049. [Google Scholar] [CrossRef]
  45. Malekshoar, G.; Ray, A.K. In-situ grown molybdenum sulfide on TiO2 for dye-sensitized solar photocatalytic hydrogen generation. Chem. Eng. Sci. 2016, 152, 35–44. [Google Scholar] [CrossRef]
  46. Ren, L.; Yang, M.-M.; Tung, C.-H.; Wu, L.-Z.; Cong, H. Visible-Light Photocatalysis Employing Dye-Sensitized Semiconductor: Selective Aerobic Oxidation of Benzyl Ethers. ACS Catal. 2017, 7, 8134–8138. [Google Scholar] [CrossRef]
  47. Zhang, Y.; Wang, Z.; Lang, X. Merging visible light photocatalysis of dye-sensitized TiO2 with TEMPO: The selective aerobic oxidation of alcohols. Catal. Sci. Technol. 2017, 7, 4955–4963. [Google Scholar] [CrossRef]
  48. Lang, X.; Zhao, J.; Chen, X. Visible-Light-Induced Photoredox Catalysis of Dye-Sensitized Titanium Dioxide: Selective Aerobic Oxidation of Organic Sulfides. Angew. Chem. Int. Ed. 2016, 55, 4697–4700. [Google Scholar] [CrossRef]
  49. Hao, H.; Wang, Z.; Shi, J.-L.; Li, X.; Lang, X. Improving the Visible Light Photocatalytic Aerobic Oxidation of Sulfides into Sulfoxides on Dye-Sensitized TiO2. ChemCatChem 2018, 10, 4545–4554. [Google Scholar] [CrossRef]
  50. Wang, Z.; Lang, X. Visible light photocatalysis of dye-sensitized TiO2: The selective aerobic oxidation of amines to imines. Appl. Catal. B Environ. 2018, 224, 404–409. [Google Scholar] [CrossRef]
  51. Li, X.; Shi, J.-L.; Hao, H.; Lang, X. Visible light-induced selective oxidation of alcohols with air by dye-sensitized TiO2 photocatalysis. Appl. Catal. B Environ. 2018, 232, 260–267. [Google Scholar] [CrossRef]
  52. Hao, H.; Li, X.; Lang, X. Anthraquinones as photoredox active ligands of TiO2 for selective aerobic oxidation of organic sulfides. Appl. Catal. B Environ. 2019, 259, 118038. [Google Scholar] [CrossRef]
  53. Hosseini-Sarvari, M.; Koohgard, M.; Firoozi, S.; Mohajeri, A.; Tavakolian, H. Alizarin red S–TiO2-catalyzed cascade C(sp3)–H to C(sp2)–H bond formation/cyclization reactions toward tetrahydroquinoline derivatives under visible light irradiation. New J. Chem. 2018, 42, 6880–6888. [Google Scholar] [CrossRef]
  54. Watanabe, M.; Hagiwara, H.; Ogata, Y.; Staykov, A.; Bishop, S.R.; Perry, N.H.; Chang, Y.J.; Ida, S.; Tanaka, K.; Ishihara, T. Impact of alkoxy chain length on carbazole-based, visible light-driven, dye sensitized photocatalytic hydrogen production. J. Mater. Chem. A 2015, 3, 21713–21721. [Google Scholar] [CrossRef]
  55. Tiwari, A.; Mondal, I.; Pal, U. Visible light induced hydrogen production over thiophenothiazine-based dye sensitized TiO2 photocatalyst in neutral water. RSC Adv. 2015, 5, 31415–31421. [Google Scholar] [CrossRef]
  56. Tiwari, A.; Pal, U. Effect of donor-donor-π-acceptor architecture of triphenylamine-based organic sensitizers over TiO2 photocatalysts for visible-light-driven hydrogen production. Int. J. Hydrogen Energy 2015, 40, 9069–9079. [Google Scholar] [CrossRef]
  57. Dessì, A.; Monai, M.; Bessi, M.; Montini, T.; Calamante, M.; Mordini, A.; Reginato, G.; Trono, C.; Fornasiero, P.; Zani, L. Towards Sustainable H2 Production: Rational Design of Hydrophobic Triphenylamine-based Dyes for Sensitized Ethanol Photoreforming. ChemSusChem 2018, 11, 793–805. [Google Scholar] [CrossRef]
  58. Cecconi, B.; Manfredi, N.; Ruffo, R.; Montini, T.; Romero-Ocaña, I.; Fornasiero, P.; Abbotto, A. Tuning Thiophene-Based Phenothiazines for Stable Photocatalytic Hydrogen Production. ChemSusChem 2015, 8, 4216–4228. [Google Scholar] [CrossRef]
  59. Manfredi, N.; Monai, M.; Montini, T.; Peri, F.; De Angelis, F.; Fornasiero, P.; Abbotto, A. Dye-Sensitized Photocatalytic Hydrogen Generation: Efficiency Enhancement by Organic Photosensitizer–Coadsorbent Intermolecular Interaction. ACS Energy Lett. 2018, 3, 85–91. [Google Scholar] [CrossRef]
  60. Kumar, N.S.; Dhar, A.; Ibrahim, A.A.; Vekariya, R.L.; Bhadja, P. Designing and fabrication of phenothiazine and carbazole based sensitizers for photocatalytic water splitting application. Int. J. Hydrogen Energy 2018, 43, 17057–17063. [Google Scholar] [CrossRef]
  61. Luo, G.G.; Lu, H.; Wang, Y.H.; Dong, J.; Zhao, Y.; Wu, R.B. A D-π-A-π-A metal-free organic dye with improved efficiency for the application of solar energy conversion. Dyes Pigment. 2016, 134, 498–505. [Google Scholar] [CrossRef]
  62. Suryani, O.; Higashino, Y.; Sato, H.; Kubo, Y. Visible-to-Near-Infrared Light-Driven Photocatalytic Hydrogen Production Using Dibenzo-BODIPY and Phenothiazine Conjugate as Organic Photosensitizer. ACS Appl. Energy Mater. 2019, 2, 448–458. [Google Scholar] [CrossRef]
  63. Yu, F.; Cui, S.-C.; Li, X.; Peng, Y.; Yu, Y.; Yun, K.; Zhang, S.-C.; Li, J.; Liu, J.-G.; Hua, J. Effect of anchoring groups on N-annulated perylene-based sensitizers for dye-sensitized solar cells and photocatalytic H2 evolution. Dyes Pigment. 2017, 139, 7–18. [Google Scholar] [CrossRef]
  64. Aslan, E.; Gonce, M.K.; Yigit, M.Z.; Sarilmaz, A.; Stathatos, E.; Ozel, F.; Can, M.; Patir, I.H. Photocatalytic H2 evolution with a Cu2WS4 catalyst on a metal free D-π-A organic dye-sensitized TiO2. Appl. Catal. B Environ. 2017, 210, 320–327. [Google Scholar] [CrossRef]
  65. Sun, Y.; Sun, Y.; Dall’Agnese, C.; Wang, X.-F.; Chen, G.; Kitao, O.; Tamiaki, H.; Sakai, K.; Ikeuchi, T.; Sasaki, S. Dyad Sensitizer of Chlorophyll with Indoline Dye for Panchromatic Photocatalytic Hydrogen Evolution. ACS Appl. Energy Mater. 2018, 1, 2813–2820. [Google Scholar] [CrossRef]
  66. Konieczna, D.D.; Biller, H.; Witte, M.; Schmidt, W.G.; Neuba, A.; Wilhelm, R. New pyridinium based ionic dyes for the hydrogen evolution reaction. Tetrahedron 2018, 74, 142–149. [Google Scholar] [CrossRef]
  67. Tiwari, A.; Duvva, N.; Rao, V.N.; Venkatakrishnan, S.M.; Giribabu, L.; Pal, U. Tetrathiafulvalene Scaffold-Based Sensitizer on Hierarchical Porous TiO2: Efficient Light-Harvesting Material for Hydrogen Production. J. Phys. Chem. C 2019, 123, 70–81. [Google Scholar] [CrossRef]
  68. Jo, M.; Choi, S.; Jo, J.H.; Kim, S.-Y.; Kim, P.S.; Kim, C.H.; Son, H.-J.; Pac, C.; Kang, S.O. Utility of Squaraine Dyes for Dye-Sensitized Photocatalysis on Water or Carbon Dioxide Reduction. ACS Omega 2019, 4, 14272–14283. [Google Scholar] [CrossRef] [Green Version]
  69. Joseph, K.L.V.; Lim, J.; Anthonysamy, A.; Kim, H.; Choi, W.; Kim, J.K. Squaraine-sensitized composite of a reduced graphene oxide/TiO2 photocatalyst: π–π stacking as a new method of dye anchoring. J. Mater. Chem. A 2015, 3, 232–239. [Google Scholar] [CrossRef]
  70. Buddee, S.; Wongnawa, S. Removal of dyes by photocatalytically active curcumin-sensitized amorphous TiO2 under visible light irradiation. J. Sol-Gel Sci. Technol. 2015, 75, 152–163. [Google Scholar] [CrossRef]
  71. Zyoud, A.H.; Saleh, F.; Helal, M.H.; Shawahna, R.; Hilal, H.S. Anthocyanin-Sensitized TiO2 Nanoparticles for Phenazopyridine Photodegradation under Solar Simulated Light. J. Nanomater. 2018, 2018, 2789616. [Google Scholar] [CrossRef] [Green Version]
  72. Dil, M.A.; Haghighatzadeh, A.; Mazinani, B. Photosensitization effect on visible-light-induced photocatalytic performance of TiO2/chlorophyll and flavonoid nanostructures: Kinetic and isotherm studies. Bull. Mater. Sci. 2019, 42, 1–16. [Google Scholar] [CrossRef] [Green Version]
  73. Patiño-Camelo, K.; Diaz-Uribe, C.; Gallego-Cartagena, E.; Vallejo, W.; Martinez, V.; Quiñones, C.; Hurtado, M.; Schott, E. Cyanobacterial biomass pigments as natural sensitizer for TiO2 Thin Films. Int. J. Photoenergy 2019, 2019. [Google Scholar] [CrossRef] [Green Version]
  74. Diaz-Uribe, C.; Vallejo, W.; Camargo, G.; Muñoz-Acevedo, A.; Quiñones, C.; Schott, E.; Zarate, X. Potential use of an anthocyanin-rich extract from berries of Vaccinium meridionale Swartz as sensitizer for TiO2 thin films–An experimental and theoretical study. J. Photochem. Photobiol. A Chem. 2019, 384, 112050. [Google Scholar] [CrossRef]
  75. Fernández, J.A.; Suan, A.; Ramírez, J.C.; Robles, J.; Salcedo, J.C.; Pedroza, A.M.; Daza, C.E. Treatment of real wastewater with TiO2-films sensitized by a natural-dye obtained from Picramnia sellowii. J. Environ. Chem. Eng. 2016, 4, 2848–2856. [Google Scholar] [CrossRef]
  76. Mu, Q.; Sun, Y.; Guo, A.; Yu, X.; Xu, X.; Cai, A.; Wang, X. Bio-templated synthesis of Fe3O4–TiO2 composites derived from Chlorella pyrenoidosa with enhanced visible-light photocatalytic performance. Mater. Res. Express 2019, 6, 0950c3. [Google Scholar] [CrossRef]
  77. Ali, S.S.; Qazi, I.A.; Arshad, M.; Khan, Z.; Voice, T.C.; Mehmood, C.T. Photocatalytic degradation of low density polyethylene (LDPE) films using titania nanotubes. Environ. Nanotechnol. Monit. Manag. 2016, 5, 44–53. [Google Scholar] [CrossRef] [Green Version]
  78. Safaralizadeh, E.; Darzi, S.J.; Mahjoub, A.R.; Abazari, R. Visible light-induced degradation of phenolic compounds by Sudan black dye sensitized TiO2 nanoparticles as an advanced photocatalytic material. Res. Chem. Intermed. 2017, 43, 1197–1209. [Google Scholar] [CrossRef]
  79. Oprea, C.I.; Gîrțu, M.A. Structure and electronic properties of TiO2 nanoclusters and dye–nanocluster systems appropriate to model hybrid photovoltaic or photocatalytic applications. Nanomaterials 2019, 9, 357. [Google Scholar] [CrossRef] [Green Version]
  80. Patwari, J.; Chatterjee, A.; Sardar, S.; Lemmens, P.; Pal, S.K. Ultrafast dynamics in co-sensitized photocatalysts under visible and NIR light irradiation. Phys. Chem. Chem. Phys. 2018, 20, 10418–10429. [Google Scholar] [CrossRef]
  81. Radhika, S.; Thomas, J. Solar light driven photocatalytic degradation of organic pollutants using ZnO nanorods coupled with photosensitive molecules. J. Environ. Chem. Eng. 2017, 5, 4239–4250. [Google Scholar]
  82. Moussawi, R.N.; Patra, D. Nanoparticle Self-Assembled Grain Like Curcumin Conjugated ZnO: Curcumin Conjugation Enhances Removal of Perylene, Fluoranthene and Chrysene by ZnO. Sci. Rep. 2016, 6, 24565. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Zyoud, A.; Dwikat, M.; Al-Shakhshir, S.; Ateeq, S.; Shteiwi, J.; Zu’bi, A.; Helal, M.H.S.; Campet, G.; Park, D.; Kwon, H.; et al. Natural dye-sensitized ZnO nano-particles as photo-catalysts in complete degradation of E. coli bacteria and their organic content. J. Photochem. Photobiol. A Chem. 2016, 328, 207–216. [Google Scholar] [CrossRef]
  84. George, L.; Sappati, S.; Ghosh, P.; Devi, R.N. Surface Site Modulations by Conjugated Organic Molecules To Enhance Visible Light Activity of ZnO Nanostructures in Photocatalytic Water Splitting. J. Phys. Chem. C 2015, 119, 3060–3067. [Google Scholar] [CrossRef]
  85. He, L.; Dou, X.; Li, X.; Qin, L.; Kang, S.-Z. Remarkable enhancement of the photocatalytic activity of ZnO nanorod array by utilizing energy transfer between Eosin Y and Rose Bengal for visible light-driven hydrogen evolution. Int. J. Hydrogen Energy 2018, 43, 15255–15261. [Google Scholar] [CrossRef]
  86. Underwood, T.M.; Robinson, R.S. Enhancing the photo-efficacy of an organic visible-light-activated chromophore (alizarin red S) on zinc oxide with a Ag–Na electrolyte to photo-transform aromatic and aliphatic alcohols. RSC Adv. 2019, 9, 24259–24266. [Google Scholar] [CrossRef] [Green Version]
  87. Yang, L.; Huang, J.; Shi, L.; Cao, L.; Zhou, W.; Chang, K.; Meng, X.; Liu, G.; Jie, Y.; Ye, J. Efficient hydrogen evolution over Sb doped SnO2 photocatalyst sensitized by Eosin Y under visible light irradiation. Nano Energy 2017, 36, 331–340. [Google Scholar] [CrossRef]
  88. He, H.; Luo, Z.; Tang, Z.-Y.; Yu, C. Controllable construction of ZnWO4 nanostructure with enhanced performance for photosensitized Cr(VI) reduction. Appl. Surf. Sci. 2019, 490, 460–468. [Google Scholar] [CrossRef]
  89. Li, B.; Zhang, Y.; Duanmu, F.; Shen, Y.; Shen, Z.; Zhong, S. Eosin Y sensitized BiPO4 nanorods for bi-functionally enhanced visible-light-driven photocatalysis. Photochem. Photobiol. Sci. 2019, 18, 1408–1418. [Google Scholar] [CrossRef]
  90. Zhao, H.; Zhang, Y.; Li, G.; Tian, F.; Tang, H.; Chen, R. Rhodamine B-sensitized BiOCl hierarchical nanostructure for methyl orange photodegradation. RSC Adv. 2016, 6, 7772–7779. [Google Scholar] [CrossRef]
  91. Yadav, M.; Garg, S.; Chandra, A.; Hernadi, K. Quercetin-sensitized BiOF nanostructures: An investigation on photoinduced charge transfer and regeneration process for degradation of organic pollutants. J. Photochem. Photobiol. A Chem. 2019, 383, 112014. [Google Scholar] [CrossRef] [Green Version]
  92. Chianese, G.; Terracciano, M.; Moretta, R.; Cappiello, P.; Vitiello, G.; Aronne, A.; Schiattarella, C.; De Stefano, L.; Rea, I. Synthesis and surface modification of nanostructured F-doped ZnO: Toward a transducer for label-free optical biosensing. Appl. Sci. 2019, 9, 3380. [Google Scholar] [CrossRef] [Green Version]
  93. Bera, A.; Bagchi, D.; Pal, S.K. Improvement of Photostability and NIR Activity of Cyanine Dye through Nanohybrid Formation: Key Information from Ultrafast Dynamical Studies. J. Phys. Chem. A 2019, 123, 7550–7557. [Google Scholar] [CrossRef] [PubMed]
  94. Kumar, P.; Kumar, S.; Ghosh, S.; Pal, S.K. Femtosecond insights into direct electron injection in dye anchored ZnO QDs following charge transfer excitation. Phys. Chem. Chem. Phys. 2016, 18, 20672–20681. [Google Scholar] [CrossRef]
  95. Maity, P.; Debnath, T.; Akbar, A.; Verma, S.; Ghosh, H.N. Ultrafast Electron-Transfer and -Trapping Dynamics in the Inter-Band-Gap States of ZrO2 Nanoparticles Sensitized by Baicalein. J. Phys. Chem. C 2013, 117, 17531–17539. [Google Scholar] [CrossRef]
  96. Rabbani, M.; Rafiee, F.; Ghafuri, H.; Rahimi, R. Synthesis of Fe3O4 nonoparticles via a fast and facile mechanochemicl method: Modification of surface with porphyrin and photocatalytic study. Mater. Lett. 2016, 166, 247–250. [Google Scholar] [CrossRef]
  97. Macyk, W.; Szaciłowski, K.; Stochel, G.; Buchalska, M.; Kuncewicz, J.; Łabuz, P. Titanium(IV) complexes as direct TiO2 photosensitizers. Coord. Chem. Rev. 2010, 254, 2687–2701. [Google Scholar] [CrossRef]
  98. Rajh, T.; Chen, L.X.; Lukas, K.; Liu, T.; Thurnauer, M.C.; Tiede, D.M. Surface Restructuring of Nanoparticles: An Efficient Route for Ligand—Metal Oxide Crosstalk. J. Phys. Chem. B 2002, 106, 10543–10552. [Google Scholar] [CrossRef]
  99. Eda, T.; Fujisawa, J.; Hanaya, M. Inorganic Control of Interfacial Charge-Transfer Transitions in Catechol-Functionalized Titanium Oxides Using SrTiO3, BaTiO3, and TiO2. J. Phys. Chem. C 2018, 122, 16216–16220. [Google Scholar] [CrossRef]
  100. Kondratyeva, I.; Orzeł, Ł.; Kobasa, I.; Doroshenko, A.; Macyk, W. Photosensitization of titanium dioxide with 4′-dimethylaminoflavonol. Mater. Sci. Semicond. Process. 2016, 42, 62–65. [Google Scholar] [CrossRef]
  101. Varaganti, S.; Ramakrishna, G. Dynamics of Interfacial Charge Transfer Emission in Small-molecule Sensitized TiO2 Nanoparticles: Is it Localized or Delocalized? J. Phys. Chem. C 2010, 114, 13917–13925. [Google Scholar] [CrossRef]
  102. Qi, K.; Zasada, F.; Piskorz, W.; Indyka, P.; Gryboś, J.; Trochowski, M.; Buchalska, M.; Kobielusz, M.; Macyk, W.; Sojka, Z. Self-Sensitized Photocatalytic Degradation of Colorless Organic Pollutants Attached to Rutile Nanorods: Experimental and Theoretical DFT+D Studies. J. Phys. Chem. C 2016, 120, 5442–5456. [Google Scholar] [CrossRef]
  103. Kim, G.; Lee, S.H.; Choi, W. Glucose-TiO2 charge transfer complex-mediated photocatalysis under visible light. Appl. Catal. B Environ. 2015, 162, 463–469. [Google Scholar] [CrossRef]
  104. Finkelstein-Shapiro, D.; Davidowski, S.K.; Lee, P.B.; Guo, C.; Holland, G.P.; Rajh, T.; Gray, K.A.; Yarger, J.L.; Calatayud, M. Direct Evidence of Chelated Geometry of Catechol on TiO2 by a Combined Solid-State NMR and DFT Study. J. Phys. Chem. C 2016, 120, 23625–23630. [Google Scholar] [CrossRef]
  105. Mowbray, D.J.; Migani, A. Optical Absorption Spectra and Excitons of Dye-Substrate Interfaces: Catechol on TiO2(110). J. Chem. Theory Comput. 2016, 12, 2843–2852. [Google Scholar] [CrossRef] [Green Version]
  106. Ooyama, Y.; Yamaji, K.; Ohshita, J. Photovoltaic performances of type-II dye-sensitized solar cells based on catechol dye sensitizers: Retardation of back-electron transfer by PET (photo-induced electron transfer). Mater. Chem. Front. 2017, 1, 2243–2255. [Google Scholar] [CrossRef]
  107. Shi, J.L.; Hao, H.; Li, X.; Lang, X. Merging the catechol-TiO2 complex photocatalyst with TEMPO for selective aerobic oxidation of amines into imines. Catal. Sci. Technol. 2018, 8, 3910–3917. [Google Scholar] [CrossRef]
  108. Barbieriková, Z.; Dvoranová, D.; Brezová, V.; Džunuzović, E.; Sredojević, D.N.; Lazić, V.; Nedeljković, J.M. Visible-light-responsive surface-modified TiO2 powder with 4-chlorophenol: A combined experimental and DFT study. Opt. Mater. 2019, 89, 237–242. [Google Scholar] [CrossRef]
  109. Higashimoto, S.; Nishi, T.; Yasukawa, M.; Azuma, M.; Sakata, Y.; Kobayashi, H. Photocatalysis of titanium dioxide modified by catechol-type interfacial surface complexes (ISC) with different substituted groups. J. Catal. 2015, 329, 286–290. [Google Scholar] [CrossRef]
  110. Nishikiori, H.; Todoroki, K.; Setiawan, R.A.; Teshima, K.; Fujii, T.; Satozono, H. Titanium Complex Formation of Organic Ligands in Titania Gels. Langmuir 2015, 31, 964–969. [Google Scholar] [CrossRef]
  111. Fukui, M.; Tanaka, A.; Hashimoto, K.; Kominami, H. Visible light-induced heterogeneous Meerwein–Ponndorf–Verley-type reduction of an aldehyde group over an organically modified titanium dioxide photocatalyst. Chem. Commun. 2017, 53, 4215–4218. [Google Scholar] [CrossRef] [PubMed]
  112. Yamamoto, Y.; Fukui, M.; Tanaka, A.; Hashimoto, K.; Kominami, H. Hydrogen- and noble metal-free conversion of nitro aromatics to amino aromatics having reducible groups over an organically modified TiO2 photocatalyst under visible light irradiation. Catal. Sci. Technol. 2019, 9, 966–973. [Google Scholar] [CrossRef]
  113. Kuncewicz, J.; Koroński, K.; Majewska, P.; Adamowicz, W.; Macyk, W. Visible light active titanates photosensitized by Ti(IV) surface complexes. Appl. Surf. Sci. 2019, 473, 1066–1073. [Google Scholar] [CrossRef]
  114. Zhang, X.; Peng, B.; Peng, T.; Yu, L.; Li, R.; Zhang, J. A new route for visible/near-infrared-light-driven H2 production over titania: Co-sensitization of surface charge transfer complex and zinc phthalocyanine. J. Power Sources 2015, 298, 30–37. [Google Scholar] [CrossRef]
  115. Li, X.; Xu, H.; Shi, J.-L.; Hao, H.; Yuan, H.; Lang, X. Salicylic acid complexed with TiO2 for visible light-driven selective oxidation of amines into imines with air. Appl. Catal. B Environ. 2019, 244, 758–766. [Google Scholar] [CrossRef]
  116. Božanić, D.K.; Garcia, G.A.; Nahon, L.; Sredojević, D.; Lazić, V.; Vukoje, I.; Ahrenkiel, S.P.; Djoković, V.; Šljivančanin, Ž.; Nedeljković, J.M. Interfacial Charge Transfer Transitions in Colloidal TiO2 Nanoparticles Functionalized with Salicylic acid and 5-Aminosalicylic acid: A Comparative Photoelectron Spectroscopy and DFT Study. J. Phys. Chem. C 2019, 123, 29057–29066. [Google Scholar] [CrossRef]
  117. Savić, T.D.; Čomor, M.I.; Abazović, N.D.; Šaponjić, Z.V.; Marinović-Cincović, M.T.; Veljković, D.Ž.; Zarić, S.D.; Janković, I.A. Anatase nanoparticles surface modified with fused ring salicylate-type ligands (1-hydroxy-2-naphthoic acids): A combined DFT and experimental study. J. Alloys Compd. 2015, 630, 226–235. [Google Scholar] [CrossRef]
  118. Karthik, P.; Neppolian, B. TiO2-glucose LMCT complex/reduced graphene oxide sheet (rGO): An efficient visible light photocatalyst for Cr(VI) reduction. J. Environ. Chem. Eng. 2018, 6, 3664–3672. [Google Scholar] [CrossRef]
  119. Sredojević, D.N.; Kovač, T.; Džunuzović, E.; Ðorđević, V.; Grgur, B.N.; Nedeljković, J.M. Surface-modified TiO2 powders with phenol derivatives: A comparative DFT and experimental study. Chem. Phys. Lett. 2017, 686, 167–172. [Google Scholar] [CrossRef]
  120. Qu, L.; Huang, D.; Shi, H.; Gu, M.; Li, J.; Dong, F.; Luo, Z. TiO2/carboxylate-rich porous carbon: A highly efficient visible-light-driven photocatalyst based on the ligand-to-metal charge transfer (LMCT) process. J. Phys. Chem. Solids 2015, 85, 173–179. [Google Scholar] [CrossRef]
  121. Tian, Q.; Wang, C.; Chen, C.; Li, X.; Weng, S.; Zheng, Y.; Liu, P.; Zhuang, J. Synergetic effects of the interfacial dyadic structure on the interfacial charge transfer between surface-complex and TiO2. Appl. Surf. Sci. 2019, 496, 143711. [Google Scholar] [CrossRef]
  122. Karthik, P.; Vinoth, R.; Babu, S.G.; Wen, M.; Kamegawa, T.; Yamashita, H.; Neppolian, B. Synthesis of highly visible light active TiO2-2-naphthol surface complex and its application in photocatalytic chromium(vi) reduction. RSC Adv. 2015, 5, 39752–39759. [Google Scholar] [CrossRef]
  123. Mao, W.X.; Lin, X.J.; Zhang, W.; Chi, Z.X.; Lyu, R.W.; Cao, A.M.; Wan, L.J. Core-shell structured TiO2@polydopamine for highly active visible-light photocatalysis. Chem. Commun. 2016, 52, 7122–7125. [Google Scholar] [CrossRef] [PubMed]
  124. Kim, S.; Moon, G.H.; Kim, G.; Kang, U.; Park, H.; Choi, W. TiO2 complexed with dopamine-derived polymers and the visible light photocatalytic activities for water pollutants. J. Catal. 2017, 346, 92–100. [Google Scholar] [CrossRef]
  125. Karthik, P.; Vinoth, R.; Selvam, P.; Balaraman, E.; Navaneethan, M.; Hayakawa, Y.; Neppolian, B. A visible-light active catechol–metal oxide carbonaceous polymeric material for enhanced photocatalytic activity. J. Mater. Chem. A 2017, 5, 384–396. [Google Scholar] [CrossRef]
  126. Sadowski, R.; Strus, M.; Buchalska, M.; Heczko, P.B.; Macyk, W. Visible light induced photocatalytic inactivation of bacteria by modified titanium dioxide films on organic polymers. Photochem. Photobiol. Sci. 2015, 14, 514–519. [Google Scholar] [CrossRef] [Green Version]
  127. Zhang, L.; Yang, J.; Zhao, X.; Xiao, X.; Sun, F.; Zuo, X.; Nan, J. Small-molecule surface-modified bismuth-based semiconductors as a new class of visible-light-driven photocatalytic materials: Structure-dependent photocatalytic properties and photosensitization mechanism. Chem. Eng. J. 2020, 380, 122546. [Google Scholar] [CrossRef]
  128. Park, J.; Moon, G.; Shin, K.-O.; Kim, J. Oxalate-TiO2 complex-mediated oxidation of pharmaceutical pollutants through ligand-to-metal charge transfer under visible light. Chem. Eng. J. 2018, 343, 689–698. [Google Scholar] [CrossRef]
  129. Lim, J.; Kwak, D.; Sieland, F.; Kim, C.; Bahnemann, D.W.; Choi, W. Visible light-induced catalytic activation of peroxymonosulfate using heterogeneous surface complexes of amino acids on TiO2. Appl. Catal. B Environ. 2018, 225, 406–414. [Google Scholar] [CrossRef]
  130. Naniwa, S.; Tyagi, A.; Yamamoto, A.; Yoshida, H. Visible-light photoexcitation of pyridine surface complex, leading to selective dehydrogenative cross-coupling with cyclohexane. Phys. Chem. Chem. Phys. 2018, 20, 28375–28381. [Google Scholar] [CrossRef] [Green Version]
  131. Nishikiori, H.; Natori, D.; Ebara, H.; Teshima, K.; Fujii, T. Zinc complex formation of organic ligands on zinc oxide and titanium dioxide. J. Photochem. Photobiol. A Chem. 2016, 327, 51–57. [Google Scholar] [CrossRef]
  132. Nauth, A.M.; Schechtel, E.; Dören, R.; Tremel, W.; Opatz, T. TiO2 Nanoparticles Functionalized with Non-innocent Ligands Allow Oxidative Photocyanation of Amines with Visible/Near-Infrared Photons. J. Am. Chem. Soc. 2018, 140, 14169–14177. [Google Scholar] [CrossRef] [PubMed]
  133. Xiao, X.; Zhu, W.-W.; Lei, Y.-B.; Liu, Q.-Y.; Li, Q.; Li, W.-W. Zwitterionic buffer-induced visible light excitation of TiO2 for efficient pollutant photodegradation. RSC Adv. 2016, 6, 35449–35454. [Google Scholar] [CrossRef]
  134. Jiang, Z.; Liu, Y.; Jing, T.; Huang, B.; Wang, Z.; Zhang, X.; Qin, X.; Dai, Y. Enhancing visible light photocatalytic activity of TiO2 using a colorless molecule (2-methoxyethanol) due to hydrogen bond effect. Appl. Catal. B Environ. 2017, 200, 230–236. [Google Scholar] [CrossRef]
  135. Yamamoto, A.; Ohara, T.; Yoshida, H. Visible-light-induced photocatalytic benzene/cyclohexane cross-coupling utilizing a ligand-to-metal charge transfer benzene complex adsorbed on titanium oxides. Catal. Sci. Technol. 2018, 8, 2046–2050. [Google Scholar] [CrossRef]
  136. Peng, Y.; Lu, B.; Wu, F.; Zhang, F.; Lu, J.E.; Kang, X.; Ping, Y.; Chen, S. Point of Anchor: Impacts on Interfacial Charge Transfer of Metal Oxide Nanoparticles. J. Am. Chem. Soc. 2018, 140, 15290–15299. [Google Scholar] [CrossRef]
  137. Vukoje, I.; Kovač, T.; Džunuzović, J.; Džunuzović, E.; Lončarević, D.; Ahrenkiel, S.P.; Nedeljković, J.M. Photocatalytic Ability of Visible-Light-Responsive TiO2 Nanoparticles. J. Phys. Chem. C 2016, 120, 18560–18569. [Google Scholar] [CrossRef]
  138. Chae, A.; Jo, S.; Choi, Y.; Min, B.; Park, S.Y.; In, I. Visible-light-driven photocatalysis with dopamine-derivatized titanium dioxide/N-doped carbon core/shell nanoparticles. J. Mater. Sci. 2017, 52, 5582–5588. [Google Scholar] [CrossRef]
  139. Zhu, M.; Muhammad, Y.; Hu, P.; Wang, B.; Wu, Y.; Sun, X.; Tong, Z.; Zhao, Z. Enhanced interfacial contact of dopamine bridged melamine-graphene/TiO2 nano-capsules for efficient photocatalytic degradation of gaseous formaldehyde. Appl. Catal. B Environ. 2018, 232, 182–193. [Google Scholar] [CrossRef]
  140. Vitiello, G.; Zanfardino, A.; Tammaro, O.; Di Napoli, M.; Caso, M.F.; Pezzella, A.; Varcamonti, M.; Silvestri, B.; D’Errico, G.; Costantini, A.; et al. Bioinspired hybrid eumelanin-TiO2 antimicrobial nanostructures: The key role of organo-inorganic frameworks in tuning eumelanin’s biocide action mechanism through membrane interaction. RSC Adv. 2018, 8, 28275–28283. [Google Scholar] [CrossRef] [Green Version]
  141. Vitiello, G.; Pezzella, A.; Calcagno, V.; Silvestri, B.; Raiola, L.; D’Errico, G.; Costantini, A.; Branda, F.; Luciani, G. 5,6-Dihydroxyindole-2-carboxylic Acid-TiO2 Charge Transfer Complexes in the Radical Polymerization of Melanogenic Precursor(s). J. Phys. Chem. C 2016, 120, 6262–6268. [Google Scholar] [CrossRef]
  142. Vitiello, G.; Pezzella, A.; Zanfardino, A.; Varcamonti, M.; Silvestri, B.; Costantini, A.; Branda, F.; Luciani, G. Titania as a driving agent for DHICA polymerization: A novel strategy for the design of bioinspired antimicrobial nanomaterials. J. Mater. Chem. B 2015, 3, 2808–2815. [Google Scholar] [CrossRef]
  143. Vitiello, G.; Melone, P.; Brigida, S.; Pezzella, A.; Di Donato, P.; D’Errico, G.; Di Napoli, M.; Zanfardino, A.; Varcamonti, M.; Luciani, G. Titanium based complexes with melanin precursors as a tool for directing melanogenic pathways. Pure Appl. Chem. 2019, 91, 1605–1616. [Google Scholar] [CrossRef]
  144. Sannino, F.; Pernice, P.; Imparato, C.; Aronne, A.; D’Errico, G.; Minieri, L.; Perfetti, M.; Pirozzi, D. Hybrid TiO2-acetylacetonate amorphous gel-derived material with stably adsorbed superoxide radical active in oxidative degradation of organic pollutants. RSC Adv. 2015, 5, 93831–93839. [Google Scholar] [CrossRef]
  145. Muñoz-García, A.B.; Sannino, F.; Vitiello, G.; Pirozzi, D.; Minieri, L.; Aronne, A.; Pernice, P.; Pavone, M.; D’Errico, G. Origin and Electronic Features of Reactive Oxygen Species at Hybrid Zirconia-Acetylacetonate Interfaces. ACS Appl. Mater. Interfaces 2015, 7, 21662–21667. [Google Scholar] [CrossRef] [PubMed]
  146. Sannino, F.; Pernice, P.; Minieri, L.; Camandona, G.A.; Aronne, A.; Pirozzi, D. Oxidative Degradation of Different Chlorinated Phenoxyalkanoic Acid Herbicides by a Hybrid ZrO2 Gel-Derived Catalyst without Light Irradiation. ACS Appl. Mater. Interfaces 2015, 7, 256–263. [Google Scholar] [CrossRef]
  147. Pirozzi, D.; Imparato, C.; D’Errico, G.; Vitiello, G.; Aronne, A.; Sannino, F. Three-year lifetime and regeneration of superoxide radicals on the surface of hybrid TiO2 materials exposed to air. J. Hazard. Mater. 2019, 121716. [Google Scholar] [CrossRef]
  148. Tachan, Z.; Hod, I.; Zaban, A. The TiO2-catechol complex: Coupling type II sensitization with efficient catalysis of water oxidation. Adv. Energy Mater. 2014, 4, 1–7. [Google Scholar] [CrossRef]
  149. Addonizio, M.L.; Aronne, A.; Imparato, C. Amorphous hybrid TiO2 thin films: The role of organic ligands and UV irradiation. Appl. Surf. Sci. 2020, 502, 144095. [Google Scholar] [CrossRef]
  150. Alam, M.; Chakraborty, I.; Mandal, K. Microwave synthesis of surface functionalized ErFeO3 nanoparticles for photoluminescence and excellent photocatalytic activity. J. Lumin. 2018, 196, 387–391. [Google Scholar] [CrossRef]
  151. Tamai, K.; Murakami, K.; Hosokawa, S.; Asakura, H.; Teramura, K.; Tanaka, T. Visible-Light Selective Photooxidation of Aromatic Hydrocarbons via Ligand-to-Metal Charge Transfer Transition on Nb2O5. J. Phys. Chem. C 2017, 121, 22854–22861. [Google Scholar] [CrossRef]
  152. Schindler, T.; Lin, W.; Schmutzler, T.; Lindner, P.; Peukert, W.; Segets, D.; Unruh, T. Evolution of the Ligand Shell Around Small ZnO Nanoparticles During the Exchange of Acetate by Catechol: A Small Angle Scattering Study. ChemNanoMat 2019, 5, 116–123. [Google Scholar] [CrossRef]
  153. Burger, A.; Kunzmann, A.; Costa, R.D.; Srikantharajah, R.; Peukert, W.; Guldi, D.M.; Hirsch, A. Synergy of Catechol-Functionalized Zinc Oxide Nanorods and Porphyrins in Layer-by-Layer Assemblies. Chem. Eur. J. 2018, 24, 7896–7905. [Google Scholar] [CrossRef]
  154. Yang, J.; Hao, J.; Xu, S.; Dai, J.; Wang, Y.; Pang, X. Visible-light-driven photocatalytic degradation of 4-CP and the synergistic reduction of Cr(VI) on one-pot synthesized amorphous Nb2O5 nanorods/graphene heterostructured composites. Chem. Eng. J. 2018, 353, 100–114. [Google Scholar] [CrossRef]
  155. Liu, X.; Ma, R.; Wang, X.; Ma, Y.; Yang, Y.; Zhuang, L.; Zhang, S.; Jehan, R.; Chen, J.; Wang, X. Graphene oxide-based materials for efficient removal of heavy metal ions from aqueous solution: A review. Environ. Pollut. 2019, 252, 62–73. [Google Scholar] [CrossRef]
  156. Albero, J.; Mateo, D.; García, H. Graphene-based materials as efficient photocatalysts for water splitting. Molecules 2019, 24, 906. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Zhang, K.; Wang, L.; Kim, J.K.; Ma, M.; Veerappan, G.; Lee, C.L.; Kong, K.J.; Lee, H.; Park, J.H. An order/disorder/water junction system for highly efficient co-catalyst-free photocatalytic hydrogen generation. Energy Environ. Sci. 2016, 9, 499–503. [Google Scholar] [CrossRef]
  158. Colmenares, J.C.; Varma, R.S.; Lisowski, P. Sustainable hybrid photocatalysts: Titania immobilized on carbon materials derived from renewable and biodegradable resources. Green Chem. 2016, 18, 5736–5750. [Google Scholar] [CrossRef] [Green Version]
  159. Lisowski, P.; Colmenares, J.C.; Mašek, O.; Lisowski, W.; Lisovytskiy, D.; Kamińska, A.; Łomot, D. Dual Functionality of TiO2/Biochar Hybrid Materials: Photocatalytic Phenol Degradation in the Liquid Phase and Selective Oxidation of Methanol in the Gas Phase. ACS Sustain. Chem. Eng. 2017, 5, 6274–6287. [Google Scholar] [CrossRef] [Green Version]
  160. Cao, S.; Yu, J. Carbon-based H2-production photocatalytic materials. J. Photochem. Photobiol. C Photochem. Rev. 2016, 27, 72–99. [Google Scholar] [CrossRef]
  161. Sharma, S.; Dutta, V.; Singh, P.; Raizada, P.; Rahmani-Sani, A.; Hosseini-Bandegharaei, A.; Thakur, V.K. Carbon quantum dot supported semiconductor photocatalysts for efficient degradation of organic pollutants in water: A review. J. Clean. Prod. 2019, 228, 755–769. [Google Scholar] [CrossRef]
  162. Ge, J.; Zhang, Y.; Park, S.J. Recent advances in carbonaceous photocatalysts with enhanced photocatalytic performances: A mini review. Materials 2019, 12, 1916. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Ng, Y.H.; Ikeda, S.; Matsumura, M.; Amal, R. A perspective on fabricating carbon-based nanomaterials by photocatalysis and their applications. Energy Environ. Sci. 2012, 5, 9307–9318. [Google Scholar] [CrossRef]
  164. Cheng, G.; Xu, F.; Xiong, J.; Tian, F.; Ding, J.; Stadler, F.J.; Chen, R. Enhanced adsorption and photocatalysis capability of generally synthesized TiO2-carbon materials hybrids. Adv. Powder Technol. 2016, 27, 1949–1962. [Google Scholar] [CrossRef]
  165. Tian, H.; Shen, K.; Hu, X.; Qiao, L.; Zheng, W. N, S co-doped graphene quantum dots-graphene-TiO2nanotubes composite with enhanced photocatalytic activity. J. Alloys Compd. 2017, 691, 369–377. [Google Scholar] [CrossRef]
  166. Razzaq, A.; Grimes, C.A.; In, S. Il Facile fabrication of a noble metal-free photocatalyst: TiO2 nanotube arrays covered with reduced graphene oxide. Carbon 2016, 98, 537–544. [Google Scholar] [CrossRef]
  167. Lettieri, S.; Gargiulo, V.; Pallotti, D.K.; Vitiello, G.; Maddalena, P.; Alfè, M.; Marotta, R. Evidencing opposite charge-transfer processes at TiO2/graphene-related materials interface through combined EPR, photoluminescence and photocatalysis assessment. Catal. Today 2018, 315, 19–30. [Google Scholar] [CrossRef]
  168. Andreozzi, M.; Álvarez, M.G.; Contreras, S.; Medina, F.; Clarizia, L.; Vitiello, G.; Llorca, J.; Marotta, R. Treatment of saline produced water through photocatalysis using rGO-TiO2 nanocomposites. Catal. Today 2018, 315, 194–204. [Google Scholar] [CrossRef] [Green Version]
  169. Khalid, N.R.; Majid, A.; Tahir, M.B.; Niaz, N.A.; Khalid, S. Carbonaceous-TiO2 nanomaterials for photocatalytic degradation of pollutants: A review. Ceram. Int. 2017, 43, 14552–14571. [Google Scholar] [CrossRef]
  170. Yang, Y.; Yao, Y.; He, L.; Zhong, Y.; Ma, Y.; Yao, J. Nonaqueous synthesis of TiO2-carbon hybrid nanomaterials with enhanced stable photocatalytic hydrogen production activity. J. Mater. Chem. A 2015, 3, 10060–10068. [Google Scholar] [CrossRef]
  171. Markad, G.B.; Kapoor, S.; Haram, S.K.; Thakur, P. Metal free, carbon-TiO2 based composites for the visible light photocatalysis. Sol. Energy 2017, 144, 127–133. [Google Scholar] [CrossRef]
  172. Rangel-Mendez, J.R.; Matos, J.; Cházaro-Ruiz, L.F.; González-Castillo, A.C.; Barrios-Yáñez, G. Microwave-assisted synthesis of C-doped TiO2 and ZnO hybrid nanostructured materials as quantum-dots sensitized solar cells. Appl. Surf. Sci. 2018, 434, 744–755. [Google Scholar] [CrossRef]
  173. Qi, K.; Selvaraj, R.; Al Fahdi, T.; Al-Kindy, S.; Kim, Y.; Wang, G.C.; Tai, C.W.; Sillanpää, M. Enhanced photocatalytic activity of anatase-TiO2 nanoparticles by fullerene modification: A theoretical and experimental study. Appl. Surf. Sci. 2016, 387, 750–758. [Google Scholar] [CrossRef]
  174. Kandjani, A.E.; Sabri, Y.M.; Field, M.R.; Coyle, V.E.; Smith, R.; Bhargava, S.K. Candle-Soot Derived Photoactive and Superamphiphobic Fractal Titania Electrode. Chem. Mater. 2016, 28, 7919–7927. [Google Scholar] [CrossRef]
  175. Krýsa, J.; Baudys, M.; Vislocka, X.; Neumann-Spallart, M. Composite photocatalysts based on TiO2–carbon for air pollutant removal: Aspects of adsorption. Catal. Today 2020, 340, 34–39. [Google Scholar] [CrossRef]
  176. Magalhães, P.; Ângelo, J.; Sousa, V.M.; Nunes, O.C.; Andrade, L.; Mendes, A. Synthesis and assessment of a graphene-based composite photocatalyst. Biochem. Eng. J. 2015, 104, 20–26. [Google Scholar] [CrossRef] [Green Version]
  177. Kuvarega, A.T.; Mamba, B.B. TiO2-based Photocatalysis: Toward Visible Light-Responsive Photocatalysts Through Doping and Fabrication of Carbon-Based Nanocomposites. Crit. Rev. Solid State Mater. Sci. 2017, 42, 295–346. [Google Scholar] [CrossRef]
  178. Jiang, Z.; Jiang, D.; Yan, Z.; Liu, D.; Qian, K.; Xie, J. A new visible light active multifunctional ternary composite based on TiO2-In2O3 nanocrystals heterojunction decorated porous graphitic carbon nitride for photocatalytic treatment of hazardous pollutant and H2 evolution. Appl. Catal. B Environ. 2015, 170–171, 195–205. [Google Scholar] [CrossRef]
  179. Mondal, K.; Gupta, A. Recent Advances in Carbon–Semiconductor Nanocomposites for Water Remediation. In Water Remediation; Bhattacharya, S., Gupta, A.B., Gupta, A., Pandey, A., Eds.; Springer: Singapore, 2018; pp. 45–74. ISBN 978-981-10-7551-3. [Google Scholar]
  180. Jin, S.; Li, C.; Shrestha, L.K.; Yamauchi, Y.; Ariga, K.; Hill, J.P. Simple Fabrication of Titanium Dioxide/N-Doped Carbon Hybrid Material as Non-Precious Metal Electrocatalyst for the Oxygen Reduction Reaction. ACS Appl. Mater. Interfaces 2017, 9, 18782–18789. [Google Scholar] [CrossRef]
  181. Nagaraju, G.; Manjunath, K.; Sarkar, S.; Gunter, E.; Teixeira, S.R.; Dupont, J. TiO2-RGO hybrid nanomaterials for enhanced water splitting reaction Dedicated to 108th Birthday (born April 1, 1907) of Dr. Sree Sree Sree Shivakumara Mahaswamiji, Siddaganga Matta, Tumakuru, Karnataka, India. Int. J. Hydrogen Energy 2015, 40, 12209–12216. [Google Scholar] [CrossRef]
  182. Reddy, C.V.; Reddy, K.R.; Shetti, N.P.; Shim, J.; Aminabhavi, T.M.; Dionysiou, D.D. Hetero-nanostructured metal oxide-based hybrid photocatalysts for enhanced photoelectrochemical water splitting—A review. Int. J. Hydrogen Energy 2019. [Google Scholar] [CrossRef]
  183. Ramesh Reddy, N.; Bhargav, U.; Mamatha Kumari, M.; Cheralathan, K.K.; Sakar, M. Review on the interface engineering in the carbonaceous titania for the improved photocatalytic hydrogen production. Int. J. Hydrogen Energy 2019. [Google Scholar] [CrossRef]
  184. Shi, R.; Li, Z.; Yu, H.; Shang, L.; Zhou, C.; Waterhouse, G.I.N.; Wu, L.Z.; Zhang, T. Effect of Nitrogen Doping Level on the Performance of N-Doped Carbon Quantum Dot/TiO2 Composites for Photocatalytic Hydrogen Evolution. ChemSusChem 2017, 10, 4650–4656. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Wang, M.; Han, J.; Xiong, H.; Guo, R. Yolk@Shell Nanoarchitecture of Au@r-GO/TiO2 Hybrids as Powerful Visible Light Photocatalysts. Langmuir 2015, 31, 6220–6228. [Google Scholar] [CrossRef]
  186. Theerthagiri, J.; Salla, S.; Senthil, R.A.; Nithyadharseni, P.; Madankumar, A.; Arunachalam, P.; Maiyalagan, T.; Kim, H.S. A review on ZnO nanostructured materials: Energy, environmental and biological applications. Nanotechnology 2019, 30. [Google Scholar] [CrossRef]
  187. Samadi, M.; Zirak, M.; Naseri, A.; Kheirabadi, M.; Ebrahimi, M.; Moshfegh, A.Z. Design and Tailoring of One-Dimensional ZnO Nanomaterials for Photocatalytic Degradation of Organic Dyes: A Review. In Research on Chemical Intermediates; Springer: Heidelberg, Germany, 2019; Volume 45, ISBN 0123456789. [Google Scholar]
  188. Raizada, P.; Sudhaik, A.; Singh, P. Photocatalytic water decontamination using graphene and ZnO coupled photocatalysts: A review. Mater. Sci. Energy Technol. 2019, 2, 509–525. [Google Scholar] [CrossRef]
  189. Dong, S.; Feng, J.; Fan, M.; Pi, Y.; Hu, L.; Han, X.; Liu, M.; Sun, J.; Sun, J. Recent developments in heterogeneous photocatalytic water treatment using visible light-responsive photocatalysts: A review. RSC Adv. 2015, 5, 14610–14630. [Google Scholar] [CrossRef]
  190. Sampaio, M.J.; Bacsa, R.R.; Benyounes, A.; Axet, R.; Serp, P.; Silva, C.G.; Silva, A.M.T.; Faria, J.L. Synergistic effect between carbon nanomaterials and ZnO for photocatalytic water decontamination. J. Catal. 2015, 331, 172–180. [Google Scholar] [CrossRef]
  191. Zhang, N.; Xu, Y.J. The endeavour to advance graphene-semiconductor composite-based photocatalysis. CrystEngComm 2015, 18, 24–37. [Google Scholar] [CrossRef]
  192. Thangavel, S.; Krishnamoorthy, K.; Krishnaswamy, V.; Raju, N.; Kim, S.J.; Venugopal, G. Graphdiyne-ZnO Nanohybrids as an Advanced Photocatalytic Material. J. Phys. Chem. C 2015, 119, 22057–22065. [Google Scholar] [CrossRef]
  193. Tie, W.; Zheng, Z.; Xu, C.; Zheng, Z.; Bhattacharyya, S.S.; He, W.; Lee, S.H. Facile synthesis of carbon nanotubes covalently modified with ZnO nanorods for enhanced photodecomposition of dyes. J. Colloid Interface Sci. 2019, 537, 652–660. [Google Scholar] [CrossRef]
  194. Tie, W.; Jin, J.; Bhattacharyya, S.S.; Yue, H.; Lei, Y.; Zheng, Z.; He, W.; Lee, S.H. Covalent bonding of ZnO nanostructures with dispersible carbon nanotubes for self-assembly photocatalytic heterostructures. Appl. Surf. Sci. 2019, 492, 219–227. [Google Scholar] [CrossRef]
  195. Wang, H.; Qiu, X.; Liu, W.; Yang, D. Facile preparation of well-combined lignin-based carbon/ZnO hybrid composite with excellent photocatalytic activity. Appl. Surf. Sci. 2017, 426, 206–216. [Google Scholar] [CrossRef]
  196. Wang, H.; Qiu, X.; Zhong, R.; Fu, F.; Qian, Y.; Yang, D. One-pot in-situ preparation of a lignin-based carbon/ZnO nanocomposite with excellent photocatalytic performance. Mater. Chem. Phys. 2017, 199, 193–202. [Google Scholar] [CrossRef]
  197. Atchudan, R.; Edison, T.N.J.I.; Perumal, S.; Karthik, N.; Karthikeyan, D.; Shanmugam, M.; Lee, Y.R. Concurrent synthesis of nitrogen-doped carbon dots for cell imaging and ZnO@nitrogen-doped carbon sheets for photocatalytic degradation of methylene blue. J. Photochem. Photobiol. A Chem. 2018, 350, 75–85. [Google Scholar] [CrossRef]
  198. Bozetine, H.; Wang, Q.; Barras, A.; Li, M.; Hadjersi, T.; Szunerits, S.; Boukherroub, R. Green chemistry approach for the synthesis of ZnO-carbon dots nanocomposites with good photocatalytic properties under visible light. J. Colloid Interface Sci. 2016, 465, 286–294. [Google Scholar] [CrossRef]
  199. Braghiroli, F.L.; Fierro, V.; Szczurek, A.; Gadonneix, P.; Ghanbaja, J.; Parmentier, J.; Medjahdi, G.; Celzard, A. Hydrothermal treatment of Tannin: A route to porous metal oxides and metal/carbon hybrid materials. Inorganics 2017, 5, 7. [Google Scholar] [CrossRef] [Green Version]
  200. Ashouri, R.; Ghasemipoor, P.; Rasekh, B.; Yazdian, F.; Mofradnia, S.R.; Fattahi, M. The effect of ZnO-based carbonaceous materials for degradation of benzoic pollutants: A review. Int. J. Environ. Sci. Technol. 2019, 16, 1729–1740. [Google Scholar] [CrossRef]
  201. Moradlou, O.; Rabiei, Z.; Banazadeh, A.; Warzywoda, J.; Zirak, M. Carbon quantum dots as nano-scaffolds for A-Fe2O3 growth: Preparation of Ti/CQD@A-Fe2O3 photoanode for water splitting under visible light irradiation. Appl. Catal. B Environ. 2018, 227, 178–189. [Google Scholar] [CrossRef] [Green Version]
  202. Araújo, T.C.; Oliveira, H.D.S.; Teles, J.J.S.; Fabris, J.D.; Oliveira, L.C.A.; de Mesquita, J.P. Hybrid heterostructures based on hematite and highly hydrophilic carbon dots with photocatalytic activity. Appl. Catal. B Environ. 2016, 182, 204–212. [Google Scholar] [CrossRef]
  203. Christoforidis, K.C.; Montini, T.; Bontempi, E.; Zafeiratos, S.; Jaén, J.J.D.; Fornasiero, P. Synthesis and photocatalytic application of visible-light active β-Fe2O3/g-C3N4 hybrid nanocomposites. Appl. Catal. B Environ. 2016, 187, 171–180. [Google Scholar] [CrossRef]
  204. Kumar, R.; Singh, R.K.; Vaz, A.R.; Savu, R.; Moshkalev, S.A. Self-Assembled and One-Step Synthesis of Interconnected 3D Network of Fe3O4/Reduced Graphene Oxide Nanosheets Hybrid for High-Performance Supercapacitor Electrode. ACS Appl. Mater. Interfaces 2017, 9, 8880–8890. [Google Scholar] [CrossRef] [PubMed]
  205. Arshad, A.; Iqbal, J.; Ahmad, I.; Israr, M. Graphene/Fe3O4 nanocomposite: Interplay between photo-Fenton type reaction, and carbon purity for the removal of methyl orange. Ceram. Int. 2018, 44, 2643–2648. [Google Scholar] [CrossRef]
  206. Vinodhkumar, G.; Wilson, J.; Inbanathan, S.S.R.; Potheher, I.V.; Ashokkumar, M.; Peter, A.C. Solvothermal synthesis of magnetically separable reduced graphene oxide/Fe3O4 hybrid nanocomposites with enhanced photocatalytic properties. Phys. B Condens. Matter 2019, 580, 411752. [Google Scholar] [CrossRef]
  207. Gusain, R.; Kumar, P.; Sharma, O.P.; Jain, S.L.; Khatri, O.P. Reduced graphene oxide-CuO nanocomposites for photocatalytic conversion of CO2 into methanol under visible light irradiation. Appl. Catal. B Environ. 2016, 181, 352–362. [Google Scholar] [CrossRef]
  208. Li, H.; Zhang, X.; MacFarlane, D.R. Carbon Quantum Dots/Cu2O Heterostructures for Solar-Light-Driven Conversion of CO2 to Methanol. Adv. Energy Mater. 2015, 5, 1401077. [Google Scholar] [CrossRef]
  209. Yu, Y.; Zhu, Z.; Liu, Z.; Dong, H.; Liu, Y.; Wei, M.; Huo, P.; Li, C.; Yan, Y. Construction of the biomass carbon quantum dots modified heterojunction Bi2WO6/Cu2O photocatalysis for enhancing light utilization and mechanism insight. J. Taiwan Inst. Chem. Eng. 2019, 102, 197–201. [Google Scholar] [CrossRef]
  210. Song, B.; Wang, T.; Sun, H.; Shao, Q.; Zhao, J.; Song, K.; Hao, L.; Wang, L.; Guo, Z. Two-step hydrothermally synthesized carbon nanodots/WO3 photocatalysts with enhanced photocatalytic performance. Dalton Trans. 2017, 46, 15769–15777. [Google Scholar] [CrossRef]
  211. Pan, Y.X.; You, Y.; Xin, S.; Li, Y.; Fu, G.; Cui, Z.; Men, Y.L.; Cao, F.F.; Yu, S.H.; Goodenough, J.B. Photocatalytic CO2 Reduction by Carbon-Coated Indium-Oxide Nanobelts. J. Am. Chem. Soc. 2017, 139, 4123–4129. [Google Scholar] [CrossRef]
  212. Yan, Y.; Sun, S.; Song, Y.; Yan, X.; Guan, W.; Liu, X.; Shi, W. Microwave-assisted in situ synthesis of reduced graphene oxide-BiVO4 composite photocatalysts and their enhanced photocatalytic performance for the degradation of ciprofloxacin. J. Hazard. Mater. 2013, 250–251, 106–114. [Google Scholar] [CrossRef]
  213. Bai, J.; Xu, H.; Chen, G.; Lv, W.; Ni, Z.; Wang, Z.; Yang, J.; Qin, H.; Zheng, Z.; Li, X. Facile fabrication of α-Fe2O3/porous g-C3N4 heterojunction hybrids with enhanced visible-light photocatalytic activity. Mater. Chem. Phys. 2019, 234, 75–80. [Google Scholar] [CrossRef]
  214. Liu, C.; Wang, W.; Liu, B.; Qiao, J.; Lv, L.; Gao, X.; Zhang, X.; Xu, D.; Liu, W.; Liu, J.; et al. Recent Advances in MOF-based Nanocatalysts for Photo-Promoted CO2 Reduction Applications. Catalysts 2019, 9, 658. [Google Scholar] [CrossRef] [Green Version]
  215. Liu, Y.; Liu, Z.; Huang, D.; Cheng, M.; Zeng, G.; Lai, C.; Zhang, C.; Zhou, C.; Wang, W.; Jiang, D.; et al. Metal or metal-containing nanoparticle@MOF nanocomposites as a promising type of photocatalyst. Coord. Chem. Rev. 2019, 388, 63–78. [Google Scholar] [CrossRef]
  216. Aleksandrzak, M.; Baranowska, D.; Kedzierski, T.; Sielicki, K.; Zhang, S.; Biegun, M.; Mijowska, E. Superior synergy of g-C3N4/Cd compounds and Al-MOF-derived nanoporous carbon for photocatalytic hydrogen evolution. Appl. Catal. B Environ. 2019, 257, 117906. [Google Scholar] [CrossRef]
  217. Tian, P.; He, X.; Zhao, L.; Li, W.; Fang, W.; Chen, H.; Zhang, F.; Huang, Z.; Wang, H. Ti3C2 nanosheets modified Zr-MOFs with Schottky junction for boosting photocatalytic HER performance. Sol. Energy 2019, 188, 750–759. [Google Scholar] [CrossRef]
  218. Zuo, Q.; Liu, T.; Chen, C.; Ji, Y.; Gong, X.; Mai, Y.; Zhou, Y. Ultrathin Metal–Organic Framework Nanosheets with Ultrahigh Loading of Single Pt Atoms for Efficient Visible-Light-Driven Photocatalytic H2 Evolution. Angew. Chem. 2019, 131, 10304–10309. [Google Scholar] [CrossRef]
  219. Wang, L.; Jin, P.; Huang, J.; She, H.; Wang, Q. Integration of Copper(II)-Porphyrin Zirconium Metal–Organic Framework and Titanium Dioxide to Construct Z-Scheme System for Highly Improved Photocatalytic CO2 Reduction. ACS Sustain. Chem. Eng. 2019, 7, 15660–15670. [Google Scholar] [CrossRef]
  220. Kim, J.; Kwon, E.E. Photoconversion of carbon dioxide into fuels using semiconductors. J. CO2 Util. 2019, 33, 72–82. [Google Scholar] [CrossRef]
  221. Zhang, L.; Zhang, J. Metal-organic frameworks for CO2 photoreduction. Front. Energy 2019, 13, 221–250. [Google Scholar] [CrossRef]
  222. Zhuang, S.; Liu, Y.; Wang, J. Covalent organic frameworks as efficient adsorbent for sulfamerazine removal from aqueous solution. J. Hazard. Mater. 2020, 383, 121126. [Google Scholar] [CrossRef]
  223. Zhang, S.; Wang, Y.; Cao, Z.; Xu, J.; Hu, J.; Huang, Y.; Cui, C.; Liu, H.; Wang, H. Simultaneous enhancements of light-harvesting and charge transfer in UiO-67/CdS/rGO composites toward ofloxacin photo-degradation. Chem. Eng. J. 2020, 381, 122771. [Google Scholar] [CrossRef]
  224. Wu, T.; Liu, X.; Liu, Y.; Cheng, M.; Liu, Z.; Zeng, G.; Shao, B.; Liang, Q.; Zhang, W.; He, Q.; et al. Application of QD-MOF composites for photocatalysis: Energy production and environmental remediation. Coord. Chem. Rev. 2020, 403, 213097. [Google Scholar] [CrossRef]
  225. Lu, G.; Huang, X.; Li, Y.; Zhao, G.; Pang, G.; Wang, G. Covalently integrated core-shell MOF@COF hybrids as efficient visible-light-driven photocatalysts for selective oxidation of alcohols. J. Energy Chem. 2020, 43, 8–15. [Google Scholar] [CrossRef] [Green Version]
  226. Han, D.; Han, Y.; Li, J.; Liu, X.; Yeung, K.W.K.; Zheng, Y.; Cui, Z.; Yang, X.; Liang, Y.; Li, Z.; et al. Enhanced photocatalytic activity and photothermal effects of cu-doped metal-organic frameworks for rapid treatment of bacteria-infected wounds. Appl. Catal. B Environ. 2020, 261, 118248. [Google Scholar] [CrossRef]
  227. Gao, J.; Miao, J.; Li, P.-Z.; Teng, W.Y.; Yang, L.; Zhao, Y.; Liu, B.; Zhang, Q. A p-type Ti(iv)-based metal–organic framework with visible-light photo-response. Chem. Commun. 2014, 50, 3786–3788. [Google Scholar] [CrossRef]
  228. Doan, T.L.H.; Nguyen, H.L.; Pham, H.Q.; Pham-Tran, N.-N.; Le, T.N.; Cordova, K.E. Tailoring the Optical Absorption of Water-Stable ZrIV- and HfIV-Based Metal–Organic Framework Photocatalysts. Chem. Asian J. 2015, 10, 2660–2668. [Google Scholar] [CrossRef]
  229. Yuan, S.; Liu, T.-F.; Feng, D.; Tian, J.; Wang, K.; Qin, J.; Zhang, Q.; Chen, Y.-P.; Bosch, M.; Zou, L.; et al. A single crystalline porphyrinic titanium metal–organic framework. Chem. Sci. 2015, 6, 3926–3930. [Google Scholar] [CrossRef] [Green Version]
  230. Xu, H.-Q.; Hu, J.; Wang, D.; Li, Z.; Zhang, Q.; Luo, Y.; Yu, S.-H.; Jiang, H.-L. Visible-Light Photoreduction of CO2 in a Metal–Organic Framework: Boosting Electron–Hole Separation via Electron Trap States. J. Am. Chem. Soc. 2015, 137, 13440–13443. [Google Scholar] [CrossRef]
  231. Shi, J.; Zhang, J.; Liang, T.; Tan, D.; Tan, X.; Wan, Q.; Cheng, X.; Zhang, B.; Han, B.; Liu, L.; et al. Bipyridyl-Containing Cadmium-Organic Frameworks for Efficient Photocatalytic Oxidation of Benzylamine. ACS Appl. Mater. Interfaces 2019, 11, 30953–30958. [Google Scholar] [CrossRef]
  232. Li, P.; Li, J.; Feng, X.; Li, J.; Hao, Y.; Zhang, J.; Wang, H.; Yin, A.; Zhou, J.; Ma, X.; et al. Metal-organic frameworks with photocatalytic bactericidal activity for integrated air cleaning. Nat. Commun. 2019, 10, 1–10. [Google Scholar] [CrossRef]
  233. Toyao, T.; Saito, M.; Dohshi, S.; Mochizuki, K.; Iwata, M.; Higashimura, H.; Horiuchi, Y.; Matsuoka, M. Development of a Ru complex-incorporated MOF photocatalyst for hydrogen production under visible-light irradiation. Chem. Commun. 2014, 50, 6779–6781. [Google Scholar] [CrossRef] [PubMed]
  234. Wang, X.S.; Chen, C.H.; Ichihara, F.; Oshikiri, M.; Liang, J.; Li, L.; Li, Y.; Song, H.; Wang, S.; Zhang, T.; et al. Integration of adsorption and photosensitivity capabilities into a cationic multivariate metal-organic framework for enhanced visible-light photoreduction reaction. Appl. Catal. B Environ. 2019, 253, 323–330. [Google Scholar] [CrossRef]
  235. Hu, S.; Liu, M.; Li, K.; Song, C.; Zhang, G.; Guo, X. Surfactant-assisted synthesis of hierarchical NH2-MIL-125 for the removal of organic dyes. RSC Adv. 2017, 7, 581–587. [Google Scholar] [CrossRef] [Green Version]
  236. Guo, F.; Guo, J.-H.; Wang, P.; Kang, Y.-S.; Liu, Y.; Zhao, J.; Sun, W.-Y. Facet-dependent photocatalytic hydrogen production of metal–organic framework NH2-MIL-125(Ti). Chem. Sci. 2019, 10, 4834–4838. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  237. Chambers, M.B.; Wang, X.; Ellezam, L.; Ersen, O.; Fontecave, M.; Sanchez, C.; Rozes, L.; Mellot-Draznieks, C. Maximizing the Photocatalytic Activity of Metal–Organic Frameworks with Aminated-Functionalized Linkers: Substoichiometric Effects in MIL-125-NH2. J. Am. Chem. Soc. 2017, 139, 8222–8228. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  238. Sun, D.; Ye, L.; Li, Z. Visible-light-assisted aerobic photocatalytic oxidation of amines to imines over NH2-MIL-125(Ti). Appl. Catal. B Environ. 2015, 164, 428–432. [Google Scholar] [CrossRef]
  239. Shi, L.; Wang, T.; Zhang, H.; Chang, K.; Meng, X.; Liu, H.; Ye, J. An Amine-Functionalized Iron(III) Metal–Organic Framework as Efficient Visible-Light Photocatalyst for Cr(VI) Reduction. Adv. Sci. 2015, 2, 1500006. [Google Scholar] [CrossRef]
  240. Karmakar, A.; Pombeiro, A.J.L. Recent advances in amide functionalized metal organic frameworks for heterogeneous catalytic applications. Coord. Chem. Rev. 2019, 395, 86–129. [Google Scholar] [CrossRef]
  241. Liu, D.-C.; Ouyang, T.; Xiao, R.; Liu, W.-J.; Zhong, D.-C.; Xu, Z.; Lu, T.-B. Anchoring CoII Ions into a Thiol-Laced Metal–Organic Framework for Efficient Visible-Light-Driven Conversion of CO2 into CO. ChemSusChem 2019, 12, 2166–2170. [Google Scholar] [CrossRef]
  242. Li, F.; Wang, D.; Xing, Q.-J.; Zhou, G.; Liu, S.-S.; Li, Y.; Zheng, L.-L.; Ye, P.; Zou, J.-P. Design and syntheses of MOF/COF hybrid materials via postsynthetic covalent modification: An efficient strategy to boost the visible-light-driven photocatalytic performance. Appl. Catal. B Environ. 2019, 243, 621–628. [Google Scholar] [CrossRef]
  243. Yu, M.; Xie, Y.; Wang, X.; Li, Y.; Li, G. Highly Water-Stable Dye@Ln-MOFs for Sensitive and Selective Detection toward Antibiotics in Water. ACS Appl. Mater. Interfaces 2019, 11, 21201–21210. [Google Scholar] [CrossRef] [PubMed]
  244. Yang, X.; Huang, T.; Gao, S.; Cao, R. Boosting photocatalytic oxidative coupling of amines by a Ru-complex-sensitized metal-organic framework. J. Catal. 2019, 378, 248–255. [Google Scholar] [CrossRef]
  245. Yuan, Y.-P.; Yin, L.-S.; Cao, S.-W.; Xu, G.-S.; Li, C.-H.; Xue, C. Improving photocatalytic hydrogen production of metal–organic framework UiO-66 octahedrons by dye-sensitization. Appl. Catal. B Environ. 2015, 168–169, 572–576. [Google Scholar] [CrossRef]
  246. Gómez-Avilés, A.; Peñas-Garzón, M.; Bedia, J.; Dionysiou, D.D.; Rodríguez, J.J.; Belver, C. Mixed Ti-Zr metal-organic-frameworks for the photodegradation of acetaminophen under solar irradiation. Appl. Catal. B Environ. 2019, 253, 253–262. [Google Scholar] [CrossRef]
  247. Wen, M.; Kuwahara, Y.; Mori, K.; Zhang, D.; Li, H.; Yamashita, H. Synthesis of Ce ions doped metal–organic framework for promoting catalytic H2 production from ammonia borane under visible light irradiation. J. Mater. Chem. A 2015, 3, 14134–14141. [Google Scholar] [CrossRef]
  248. Peng, S.; Li, M.; Yang, X.; Li, P.; Liu, H.; Xiong, W.; Peng, X. Atomic layer deposition of Pt nanoparticles on ZrO2 based metal-organic frameworks for increased photocatalytic activity. Ceram. Int. 2019, 45, 18128–18134. [Google Scholar] [CrossRef]
  249. Luo, S.; Zeng, Z.; Zeng, G.; Liu, Z.; Xiao, R.; Chen, M.; Tang, L.; Tang, W.; Lai, C.; Cheng, M.; et al. Metal Organic Frameworks as Robust Host of Palladium Nanoparticles in Heterogeneous Catalysis: Synthesis, Application, and Prospect. ACS Appl. Mater. Interfaces 2019, 11, 32579–32598. [Google Scholar] [CrossRef]
  250. Chen, Y.-F.; Tan, L.-L.; Liu, J.-M.; Qin, S.; Xie, Z.-Q.; Huang, J.-F.; Xu, Y.-W.; Xiao, L.-M.; Su, C.-Y. Calix[4]arene based dye-sensitized Pt@UiO-66-NH2 metal-organic framework for efficient visible-light photocatalytic hydrogen production. Appl. Catal. B Environ. 2017, 206, 426–433. [Google Scholar] [CrossRef]
  251. Sun, D.; Liu, W.; Fu, Y.; Fang, Z.; Sun, F.; Fu, X.; Zhang, Y.; Li, Z. Noble metals can have different effects on photocatalysis over metal-organic frameworks (MOFs): A case study on M/NH2-MIL-125(Ti) (M=Pt and Au). Chem. Eur. J. 2014, 20, 4780–4788. [Google Scholar] [CrossRef]
  252. Xiao, L.; Zhang, Q.; Chen, P.; Chen, L.; Ding, F.; Tang, J.; Li, Y.J.; Au, C.T.; Yin, S.F. Copper-mediated metal-organic framework as efficient photocatalyst for the partial oxidation of aromatic alcohols under visible-light irradiation: Synergism of plasmonic effect and schottky junction. Appl. Catal. B Environ. 2019, 248, 380–387. [Google Scholar] [CrossRef]
  253. Yan, Z.H.; Ma, B.; Li, S.R.; Liu, J.; Chen, R.; Du, M.H.; Jin, S.; Zhuang, G.L.; Long, L.S.; Kong, X.J.; et al. Encapsulating a Ni(II) molecular catalyst in photoactive metal–organic framework for highly efficient photoreduction of CO2. Sci. Bull. 2019, 64, 976–985. [Google Scholar] [CrossRef] [Green Version]
  254. Choi, K.M.; Kim, D.; Rungtaweevoranit, B.; Trickett, C.A.; Barmanbek, J.T.D.; Alshammari, A.S.; Yang, P.; Yaghi, O.M. Plasmon-enhanced photocatalytic CO2 conversion within metal-organic frameworks under visible light. J. Am. Chem. Soc. 2017, 139, 356–362. [Google Scholar] [CrossRef] [PubMed]
  255. Wang, C.C.; Yi, X.H.; Wang, P. Powerful combination of MOFs and C3N4 for enhanced photocatalytic performance. Appl. Catal. B Environ. 2019, 247, 24–48. [Google Scholar] [CrossRef]
  256. Zhang, M.; Shang, Q.; Wan, Y.; Cheng, Q.; Liao, G.; Pan, Z. Self-template synthesis of double-shell TiO2@ZIF-8 hollow nanospheres via sonocrystallization with enhanced photocatalytic activities in hydrogen generation. Appl. Catal. B Environ. 2019, 241, 149–158. [Google Scholar] [CrossRef]
  257. Crake, A.; Christoforidis, K.C.; Gregg, A.; Moss, B.; Kafizas, A.; Petit, C. The Effect of Materials Architecture in TiO 2 /MOF Composites on CO 2 Photoreduction and Charge Transfer. Small 2019, 15, 1–12. [Google Scholar] [CrossRef]
  258. Wang, X.; Liu, W.; Fu, H.; Yi, X.H.; Wang, P.; Zhao, C.; Wang, C.C.; Zheng, W. Simultaneous Cr(VI) reduction and Cr(III) removal of bifunctional MOF/Titanate nanotube composites. Environ. Pollut. 2019, 249, 502–511. [Google Scholar] [CrossRef] [PubMed]
  259. Yuan, R.; Yue, C.; Qiu, J.; Liu, F.; Li, A. Highly efficient sunlight-driven reduction of Cr(VI) by TiO2@NH2-MIL-88B(Fe) heterostructures under neutral conditions. Appl. Catal. B Environ. 2019, 251, 229–239. [Google Scholar] [CrossRef]
  260. Hu, Q.; Chen, Y.; Li, M.; Zhang, Y.; Wang, B.; Zhao, Y.; Xia, J.; Yin, S.; Li, H. Construction of NH2-UiO-66/BiOBr composites with boosted photocatalytic activity for the removal of contaminants. Colloids Surf. A Physicochem. Eng. Asp. 2019, 579, 123625. [Google Scholar] [CrossRef]
  261. Yang, Z.; Tong, X.; Feng, J.; He, S.; Fu, M.; Niu, X.; Zhang, T.; Liang, H.; Ding, A.; Feng, X. Flower-like BiOBr/UiO-66-NH2 nanosphere with improved photocatalytic property for norfloxacin removal. Chemosphere 2019, 220, 98–106. [Google Scholar] [CrossRef]
  262. Emam, H.E.; Ahmed, H.B.; Gomaa, E.; Helal, M.H.; Abdelhameed, R.M. Doping of silver vanadate and silver tungstate nanoparticles for enhancement the photocatalytic activity of MIL-125-NH2 in dye degradation. J. Photochem. Photobiol. A Chem. 2019, 383, 111986. [Google Scholar] [CrossRef]
  263. Liu, Y.; Deng, L.; Sheng, J.; Tang, F.; Zeng, K.; Wang, L.; Liang, K.; Hu, H.; Liu, Y.N. Photostable core-shell CdS/ZIF-8 composite for enhanced photocatalytic reduction of CO2. Appl. Surf. Sci. 2019, 498, 143899. [Google Scholar] [CrossRef]
  264. Chen, W.Q.; Li, L.Y.; Li, L.; Qiu, W.H.; Tang, L.; Xu, L.; Xu, K.J.; Wu, M.H. MoS2/ZIF-8 Hybrid Materials for Environmental Catalysis: Solar-Driven Antibiotic-Degradation Engineering. Engineering 2019, 5, 755–767. [Google Scholar] [CrossRef]
  265. Wu, L.Y.; Mu, Y.F.; Guo, X.X.; Zhang, W.; Zhang, Z.M.; Zhang, M.; Lu, T.B. Encapsulating Perovskite Quantum Dots in Iron-Based Metal–Organic Frameworks (MOFs) for Efficient Photocatalytic CO2 Reduction. Angew. Chem. Int. Ed. 2019, 58, 9491–9495. [Google Scholar] [CrossRef] [PubMed]
  266. Mahmoodi, N.M.; Abdi, J.; Taghizadeh, M.; Taghizadeh, A.; Hayati, B.; Shekarchi, A.A.; Vossoughi, M. Activated carbon/metal-organic framework nanocomposite: Preparation and photocatalytic dye degradation mathematical modeling from wastewater by least squares support vector machine. J. Environ. Manag. 2019, 233, 660–672. [Google Scholar] [CrossRef]
  267. Shen, L.; Huang, L.; Liang, S.; Liang, R.; Qin, N.; Wu, L. Electrostatically derived self-assembly of NH2-mediated zirconium MOFs with graphene for photocatalytic reduction of Cr(vi). RSC Adv. 2014, 4, 2546–2549. [Google Scholar] [CrossRef]
  268. Liang, R.; Jing, F.; Shen, L.; Qin, N.; Wu, L. MIL-53(Fe) as a highly efficient bifunctional photocatalyst for the simultaneous reduction of Cr(VI) and oxidation of dyes. J. Hazard. Mater. 2015, 287, 364–372. [Google Scholar] [CrossRef]
  269. Wang, C.; Liu, X.-M.; Zhang, M.; Geng, Y.; Zhao, L.; Li, Y.-G.; Su, Z.-M. Two-Dimensional Cobaltporphyrin-based Cobalt–Organic Framework as an Efficient Photocatalyst for CO2 Reduction Reaction: A Computational Study. ACS Sustain. Chem. Eng. 2019, 7, 14102–14110. [Google Scholar] [CrossRef]
  270. Gao, Y.; Xia, J.; Liu, D.; Kang, R.; Yu, G.; Deng, S. Synthesis of mixed-linker Zr-MOFs for emerging contaminant adsorption and photodegradation under visible light. Chem. Eng. J. 2019, 378, 122118. [Google Scholar] [CrossRef]
  271. Gascon, J.; Hernández-Alonso, M.D.; Almeida, A.R.; van Klink, G.P.M.; Kapteijn, F.; Mul, G. Isoreticular MOFs as Efficient Photocatalysts with Tunable Band Gap: An Operando FTIR Study of the Photoinduced Oxidation of Propylene. ChemSusChem 2008, 1, 981–983. [Google Scholar] [CrossRef]
  272. Li, S.; Luo, P.; Wu, H.; Wei, C.; Hu, Y.; Qiu, G. Strategies for Improving the Performance and Application of MOFs Photocatalysts. ChemCatChem 2019, 11, 2978–2993. [Google Scholar] [CrossRef]
  273. Santaclara, J.G.; Nasalevich, M.A.; Castellanos, S.; Evers, W.H.; Spoor, F.C.M.; Rock, K.; Siebbeles, L.D.A.; Kapteijn, F.; Grozema, F.; Houtepen, A.; et al. Organic Linker Defines the Excited-State Decay of Photocatalytic MIL-125(Ti)-Type Materials. ChemSusChem 2016, 9, 388–395. [Google Scholar] [CrossRef] [PubMed]
  274. 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] [PubMed] [Green Version]
  275. Hu, Q.; Di, J.; Wang, B.; Ji, M.; Chen, Y.; Xia, J.; Li, H.; Zhao, Y. In-situ preparation of NH2-MIL-125(Ti)/BiOCl composite with accelerating charge carriers for boosting visible light photocatalytic activity. Appl. Surf. Sci. 2019, 466, 525–534. [Google Scholar] [CrossRef]
  276. Wang, Y.; Vogel, A.; Sachs, M.; Sprick, R.S.; Wilbraham, L.; Moniz, S.J.A.; Godin, R.; Zwijnenburg, M.A.; Durrant, J.R.; Cooper, A.I.; et al. Current understanding and challenges of solar-driven hydrogen generation using polymeric photocatalysts. Nat. Energy 2019, 4, 746–760. [Google Scholar] [CrossRef]
  277. Lee, S.L.; Chang, C.J. Recent developments about conductive polymer based composite photocatalysts. Polymers 2019, 11, 206. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  278. Liras, M.; Barawi, M.; de la Peña O’Shea, V.A. Hybrid materials based on conjugated polymers and inorganic semiconductors as photocatalysts: From environmental to energy applications. Chem. Soc. Rev. 2019, 48, 5454–5487. [Google Scholar] [CrossRef]
  279. Colmenares, J.C.; Kuna, E. Photoactive hybrid catalysts based on natural and synthetic polymers: A comparative overview. Molecules 2017, 22, 790. [Google Scholar] [CrossRef] [Green Version]
  280. Eskandari, E.; Kosari, M.; Davood Abadi Farahani, M.H.; Khiavi, N.D.; Saeedikhani, M.; Katal, R.; Zarinejad, M. A review on polyaniline-based materials applications in heavy metals removal and catalytic processes. Sep. Purif. Technol. 2020, 231, 115901. [Google Scholar] [CrossRef]
  281. Zarrintaj, P.; Khalili, R.; Vahabi, H.; Saeb, M.R.; Ganjali, M.R.; Mozafari, M. Polyaniline/Metal Oxides Nanocomposites. In Fundamentals and Emerging Applications of Polyaniline; Narendra, M.M., Chauhan, P.S., Eds.; Elsevier: Melbourne, Australia, 2019; pp. 131–141. ISBN 9780128179154. [Google Scholar]
  282. Zheng, L.; Yu, P.; Hu, K.; Teng, F.; Chen, H.; Fang, X. Scalable-Production, Self-Powered TiO2 Nanowell–Organic Hybrid UV Photodetectors with Tunable Performances. ACS Appl. Mater. Interfaces 2016, 8, 33924–33932. [Google Scholar] [CrossRef]
  283. Guo, N.; Liang, Y.; Lan, S.; Liu, L.; Zhang, J.; Ji, G.; Gan, S. Microscale hierarchical three-dimensional flowerlike TiO2/PANI composite: Synthesis, characterization, and its remarkable photocatalytic activity on organic dyes under UV-light and sunlight irradiation. J. Phys. Chem. C 2014, 118, 18343–18355. [Google Scholar] [CrossRef]
  284. Sahu, K.; Rahamn, K.H.; Kar, A.K. Synergic effect of polyaniline and ZnO to enhance the photocatalytic activity of their nanocomposite. Mater. Res. Express 2019, 6. [Google Scholar] [CrossRef]
  285. Zengxia, P.; Luyao, D.; Meiliang, L.; Zihan, F.; Sunxian, W.; Jun, H.; Ping, L. Synergistic Effect in Polyaniline-Hybrid Defective ZnO with Enhanced Photocatalytic Activity and Stability. J. Phys. Chem. C 2014, 118, 9570–9577. [Google Scholar]
  286. Sharma, S.; Singh, S.; Khare, N. Enhanced photosensitization of zinc oxide nanorods using polyaniline for efficient photocatalytic and photoelectrochemical water splitting. Int. J. Hydrogen Energy 2016, 41, 21088–21098. [Google Scholar] [CrossRef]
  287. Carević, M.V.; Abazović, N.D.; Mitrić, M.N.; Ćirić-Marjanović, G.; Mojović, M.D.; Ahrenkiel, S.P.; Čomor, M.I. Properties of Zirconia/Polyaniline hybrid nanocomposites and their application as photocatalysts for degradation of model pollutants. Mater. Chem. Phys. 2018, 205, 130–137. [Google Scholar] [CrossRef]
  288. Kim, K.N.; Jung, H.R.; Lee, W.J. Hollow cobalt ferrite-polyaniline nanofibers as magnetically separable visible-light photocatalyst for photodegradation of methyl orange. J. Photochem. Photobiol. A Chem. 2016, 321, 257–265. [Google Scholar] [CrossRef]
  289. Li, J.; Peng, T.; Zhang, Y.; Zhou, C.; Zhu, A. Polyaniline modified SnO2 nanoparticles for efficient photocatalytic reduction of aqueous Cr(VI) under visible light. Sep. Purif. Technol. 2018, 201, 120–129. [Google Scholar] [CrossRef]
  290. Silvestri, S.; Burgo, T.A.L.; Dias-Ferreira, C.; Labrincha, J.A.; Tobaldi, D.M. Polypyrrole-TiO2 composite for removal of 4-chlorophenol and diclofenac. React. Funct. Polym. 2019, 146, 104401. [Google Scholar] [CrossRef]
  291. Li, X.; Wang, P.; Huang, B.; Qin, X.; Zhang, X.; Zhang, Q.; Zhu, X.; Dai, Y. Precisely locate Pd-Polypyrrole on TiO2 for enhanced hydrogen production. Int. J. Hydrogen Energy 2017, 42, 25195–25202. [Google Scholar] [CrossRef]
  292. Yan, B.; Wang, Y.; Jiang, X.; Liu, K.; Guo, L. Flexible Photocatalytic Composite Film of ZnO-Microrods/Polypyrrole. ACS Appl. Mater. Interfaces 2017, 9, 29113–29119. [Google Scholar] [CrossRef]
  293. Zhang, M.; Chang, L.; Zhao, Y.; Yu, Z. Fabrication of Zinc Oxide/Polypyrrole Nanocomposites for Brilliant Green Removal from Aqueous Phase. Arab. J. Sci. Eng. 2019, 44, 111–121. [Google Scholar] [CrossRef]
  294. Silvestri, S.; Ferreira, C.D.; Oliveira, V.; Varejão, J.M.T.B.; Labrincha, J.A.; Tobaldi, D.M. Synthesis of PPy-ZnO composite used as photocatalyst for the degradation of diclofenac under simulated solar irradiation. J. Photochem. Photobiol. A Chem. 2019, 375, 261–269. [Google Scholar] [CrossRef]
  295. Brijesh, K.; Bindu, K.; Shanbhag, D.; Nagaraja, H.S. Chemically prepared Polypyrrole/ZnWO4 nanocomposite electrodes for electrocatalytic water splitting. Int. J. Hydrogen Energy 2019, 44, 757–767. [Google Scholar] [CrossRef]
  296. Wang, C.; Yang, M.; Liu, L.; Xu, Y.; Zhang, X.; Cheng, X.; Gao, S.; Gao, Y.; Huo, L. One-step synthesis of polypyrrole/Fe2O3 nanocomposite and the enhanced response of NO2 at low temperature. J. Colloid Interface Sci. 2020, 560, 312–320. [Google Scholar] [CrossRef] [PubMed]
  297. Torki, F.; Faghihian, H. Sunlight-assisted decomposition of cephalexin by novel synthesized NiS-PPY-Fe3O4 nanophotocatalyst. J. Photochem. Photobiol. A Chem. 2017, 338, 49–59. [Google Scholar] [CrossRef]
  298. Zhang, K.; Zhang, X.; Zhao, X.; Gai, X.; An, W.; Fang, G.; Zhang, A.; Chen, X. Polypyrrole coated Fe3O4 nanoparticles decorated carbon nanotubes nanocomposites and the microwave absorption properties. J. Mater. Sci. Mater. Electron. 2019, 30, 17333–17341. [Google Scholar] [CrossRef]
  299. Torki, F.; Faghihian, H. Visible Light Degradation of Naproxen by Enhanced Photocatalytic Activity of NiO and NiS, Scavenger Study and Focus on Catalyst Support and Magnetization. Photochem. Photobiol. 2018, 94, 491–502. [Google Scholar] [CrossRef]
  300. Toshima, N.; Oshima, K.; Anno, H.; Nishinaka, T.; Ichikawa, S.; Iwata, A.; Shiraishi, Y. Novel Hybrid Organic Thermoelectric Materials:Three-Component Hybrid Films Consisting of a Nanoparticle Polymer Complex, Carbon Nanotubes, and Vinyl Polymer. Adv. Mater. 2015, 27, 2246–2251. [Google Scholar] [CrossRef]
  301. Singh, S.K.; Crispin, X.; Zozoulenko, I. V Oxygen Reduction Reaction in Conducting Polymer PEDOT: Density Functional Theory Study. J. Phys. Chem. C 2017, 121, 12270–12277. [Google Scholar] [CrossRef] [Green Version]
  302. Mitraka, E.; Gryszel, M.; Vagin, M.; Jafari, M.J.; Singh, A.; Warczak, M.; Mitrakas, M.; Berggren, M.; Ederth, T.; Zozoulenko, I.; et al. Electrocatalytic Production of Hydrogen Peroxide with Poly(3,4-ethylenedioxythiophene) Electrodes. Adv. Sustain. Syst. 2019, 3, 1800110. [Google Scholar] [CrossRef]
  303. Wadnerkar, N.S.; Berggren, M.; Zozoulenko, I. Exploring Hydrogen Storage in PEDOT: A Computational Study. J. Phys. Chem. C 2019, 123, 2066–2074. [Google Scholar] [CrossRef]
  304. Abdelnasser, S.; Park, G.; Han, H.; Toth, R.; Yoon, H. Enhanced photocatalytic performance of poly(3,4-ethylenedioxythiophene)-coated TiO2 nanotube electrodes. Synth. Met. 2019, 251, 120–126. [Google Scholar] [CrossRef]
  305. Khan, S.; Narula, A.K. Ternary photocatalyst based on conducting polymer doped functionalized multiwall carbon nanotubes decorated with nanorods of metal oxide. Mater. Sci. Eng. B 2019, 243, 86–95. [Google Scholar] [CrossRef]
  306. Ramakrishnan, R.; Devaki, S.J.; Aashish, A.; Thomas, S.; Varma, M.R.; Kpp, N. Nanostructured Semiconducting PEDOT-TiO2/ZnO Hybrid Composites for Nanodevice Applications. J. Phys. Chem. C 2016, 120, 4199–4210. [Google Scholar] [CrossRef]
  307. Ansari, M.O.; Khan, M.M.; Ansari, S.A.; Cho, M.H. Polythiophene nanocomposites for photodegradation applications: Past, present and future. J. Saudi Chem. Soc. 2015, 19, 494–504. [Google Scholar] [CrossRef] [Green Version]
  308. Wang, Y.; Luo, L.; Chen, L.; Ng, P.F.; Lee, K.I.; Fei, B. Enhancement of Stability and Photocatalytic Performance on Hybrid Perovskite with Aniline. ChemNanoMat 2018, 4, 1054–1058. [Google Scholar] [CrossRef]
  309. Guo, Y.; Liu, G.; Li, Z.; Lou, Y.; Chen, J.; Zhao, Y. Stable Lead-Free (CH3NH3)3Bi2I9 Perovskite for Photocatalytic Hydrogen Generation. ACS Sustain. Chem. Eng. 2019, 7, 15080–15085. [Google Scholar] [CrossRef]
  310. Stolarczyk, J.K.; Bhattacharyya, S.; Polavarapu, L.; Feldmann, J. Challenges and Prospects in Solar Water Splitting and CO2 Reduction with Inorganic and Hybrid Nanostructures. ACS Catal. 2018, 8, 3602–3635. [Google Scholar] [CrossRef]
  311. Rokesh, K.; Sakar, M.; Do, T.O. 2-(Aminomethyl pyridine)SbI 5: An emerging visible-light driven organic–inorganic hybrid perovskite for photoelectrochemical and photocatalytic applications. Mater. Lett. 2019, 242, 99–102. [Google Scholar] [CrossRef]
  312. Zhu, X.; Lin, Y.; Sun, Y.; Beard, M.C.; Yan, Y. Lead-Halide Perovskites for Photocatalytic α-Alkylation of Aldehydes. J. Am. Chem. Soc. 2019, 141, 733–738. [Google Scholar] [CrossRef]
  313. Lyu, M.; Yun, J.H.; Chen, P.; Hao, M.; Wang, L. Addressing Toxicity of Lead: Progress and Applications of Low-Toxic Metal Halide Perovskites and Their Derivatives. Adv. Energy Mater. 2017, 7, 1602512. [Google Scholar]
  314. Zhao, Z.; Wu, J.; Zheng, Y.Z.; Li, N.; Li, X.; Ye, Z.; Lu, S.; Tao, X.; Chen, C. Stable hybrid perovskite MAPb(I1−xBrx)3 for photocatalytic hydrogen evolution. Appl. Catal. B Environ. 2019, 253, 41–48. [Google Scholar] [CrossRef]
  315. Wu, Y.; Wang, P.; Guan, Z.; Liu, J.; Wang, Z.; Zheng, Z.; Jin, S.; Dai, Y.; Whangbo, M.H.; Huang, B. Enhancing the Photocatalytic Hydrogen Evolution Activity of Mixed-Halide Perovskite CH3NH3PbBr3- xIx Achieved by Bandgap Funneling of Charge Carriers. ACS Catal. 2018, 8, 10349–10357. [Google Scholar] [CrossRef]
  316. Park, S.; Chang, W.J.; Lee, C.W.; Park, S.; Ahn, H.Y.; Nam, K.T. Photocatalytic hydrogen generation from hydriodic acid using methylammonium lead iodide in dynamic equilibrium with aqueous solution. Nat. Energy 2017, 2, 1–8. [Google Scholar] [CrossRef]
  317. Yue, C.Y.; Zhao, H.F.; Jiang, H.; Guo, Y.H.; Che, H.X.; Li, J.Z.; Chu, W.X.; Yuan, Y.; Jing, Z.H.; Lei, X.W. Large Conjugated Organic Cations Sensitized Hybrid Lead Halides as Visible Light Driven Photocatalysts. Cryst. Growth Des. 2019, 19, 4564–4570. [Google Scholar] [CrossRef]
  318. Wang, Q.; Tao, L.; Jiang, X.; Wang, M.; Shen, Y. Graphene oxide wrapped CH 3 NH 3 PbBr 3 perovskite quantum dots hybrid for photoelectrochemical CO2 reduction in organic solvents. Appl. Surf. Sci. 2019, 465, 607–613. [Google Scholar] [CrossRef]
  319. Wu, Y.; Wang, P.; Zhu, X.; Zhang, Q.; Wang, Z.; Liu, Y.; Zou, G.; Dai, Y.; Whangbo, M.H.; Huang, B. Composite of CH3NH3PbI3 with Reduced Graphene Oxide as a Highly Efficient and Stable Visible-Light Photocatalyst for Hydrogen Evolution in Aqueous HI Solution. Adv. Mater. 2018, 30, 2–7. [Google Scholar]
  320. Wang, X.; Wang, H.; Zhang, H.; Yu, W.; Wang, X.; Zhao, Y.; Zong, X.; Li, C. Dynamic Interaction between Methylammonium Lead Iodide and TiO2 Nanocrystals Leads to Enhanced Photocatalytic H2 Evolution from HI Splitting. ACS Energy Lett. 2018, 3, 1159–1164. [Google Scholar] [CrossRef]
  321. Luo, J.; Yang, H.; Liu, Z.; Li, F.; Liu, S.; Ma, J.; Liu, B. Organic-inorganic hybrid perovskite—TiO2 nanorod arrays for efficient and stable photoelectrochemical hydrogen evolution from HI splitting. Mater. Today Chem. 2019, 12, 1–6. [Google Scholar] [CrossRef]
  322. Zhao, Z.; Wu, J.; Zheng, Y.Z.; Li, N.; Li, X.; Tao, X. Ni3C-Decorated MAPbI3 as Visible-Light Photocatalyst for H2 Evolution from HI Splitting. ACS Catal. 2019, 9, 8144–8152. [Google Scholar] [CrossRef]
  323. Bresolin, B.M.; Hammouda, S.B.; Sillanpää, M. Methylammonium iodo bismuthate perovskite (CH3NH3)3Bi2I9 as new effective visible light-responsive photocatalyst for degradation of environment pollutants. J. Photochem. Photobiol. A Chem. 2019, 376, 116–126. [Google Scholar] [CrossRef]
  324. Qi, K.; Cheng, B.; Yu, J.; Ho, W. A review on TiO2-based Z-scheme photocatalysts. Chin. J. Catal. 2017, 38, 1936–1955. [Google Scholar] [CrossRef]
  325. Xu, Q.; Zhang, L.; Yu, J.; Wageh, S.; Al-Ghamdi, A.A.; Jaroniec, M. Direct Z-scheme photocatalysts: Principles, synthesis, and applications. Mater. Today 2018, 21, 1042–1063. [Google Scholar] [CrossRef]
  326. Yu, J.; Wang, S.; Low, J.; Xiao, W. Enhanced photocatalytic performance of direct Z-scheme g-C3N4–TiO2 photocatalysts for the decomposition of formaldehyde in air. Phys. Chem. Chem. Phys. 2013, 15, 16883–16890. [Google Scholar] [CrossRef] [PubMed]
  327. Liu, Y.; Ji, G.; Dastageer, M.A.; Zhu, L.; Wang, J.; Zhang, B.; Chang, X.; Gondal, M.A. Highly-active direct Z-scheme Si/TiO2 photocatalyst for boosted CO2 reduction into value-added methanol. RSC Adv. 2014, 4, 56961–56969. [Google Scholar] [CrossRef]
  328. He, S.; Rong, Q.; Niu, H.; Cai, Y. Platform for molecular-material dual regulation: A direct Z-scheme MOF/COF heterojunction with enhanced visible-light photocatalytic activity. Appl. Catal. B Environ. 2019, 247, 49–56. [Google Scholar] [CrossRef]
  329. Ju, B.; Yang, F.; Huang, K.; Wang, Y. Fabrication, characterization and photocatalytic mechanism of a novel Z-scheme BiOBr/Ag3PO4@rGO composite for enhanced visible light photocatalytic degradation. J. Alloys Compd. 2020, 815, 151886. [Google Scholar] [CrossRef]
  330. Chen, Q.; Li, J.; Cheng, L.; Liu, H. Construction of CdLa2S4/MIL-88A(Fe) heterojunctions for enhanced photocatalytic H2-evolution activity via a direct Z-scheme electron transfer. Chem. Eng. J. 2020, 379, 122389. [Google Scholar] [CrossRef]
  331. Zhou, T.; Zhang, G.; Zhang, H.; Yang, H.; Ma, P.; Li, X.; Qiu, X.; Liu, G. Highly efficient visible-light-driven photocatalytic degradation of rhodamine B by a novel Z-scheme Ag3PO4/MIL-101/NiFe2O4 composite. Catal. Sci. Technol. 2018, 8, 2402–2416. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of the elementary steps of a photocatalytic process involving an inorganic semiconductor photocatalyst (a) or an organic molecule (c). Also, a schematization of the combination between the fundamental units of semiconductor-like nodes and dye moiety containing bridging ligands in photocatalytic hybrid materials (b). Reprinted with permission from [1], Copyright (2019) Elsevier.
Figure 1. Schematic illustration of the elementary steps of a photocatalytic process involving an inorganic semiconductor photocatalyst (a) or an organic molecule (c). Also, a schematization of the combination between the fundamental units of semiconductor-like nodes and dye moiety containing bridging ligands in photocatalytic hybrid materials (b). Reprinted with permission from [1], Copyright (2019) Elsevier.
Catalysts 10 00103 g001
Scheme 1. Bandgap energies, conduction, and valence band positions (referred to the Normal Hydrogen Electrode (NHE) potential at pH 0) of different semiconductors in relation to the redox potentials of water splitting. Adapted and reprinted with permission from [5], Copyright (2017) Royal Society of Chemistry (RSC).
Scheme 1. Bandgap energies, conduction, and valence band positions (referred to the Normal Hydrogen Electrode (NHE) potential at pH 0) of different semiconductors in relation to the redox potentials of water splitting. Adapted and reprinted with permission from [5], Copyright (2017) Royal Society of Chemistry (RSC).
Catalysts 10 00103 sch001
Scheme 2. Schematic illustration of the mechanisms describing (a) dye sensitization and (b) ligand-to-metal charge transfer (LMCT) sensitization for TiO2 photocatalyst. A and D indicate electron acceptor and donor species, respectively. Adapted and reprinted with permission from [10], Copyright (2014) Royal Society of Chemistry (RSC).
Scheme 2. Schematic illustration of the mechanisms describing (a) dye sensitization and (b) ligand-to-metal charge transfer (LMCT) sensitization for TiO2 photocatalyst. A and D indicate electron acceptor and donor species, respectively. Adapted and reprinted with permission from [10], Copyright (2014) Royal Society of Chemistry (RSC).
Catalysts 10 00103 sch002
Figure 2. Molecular structures of different dyes commonly employed as sensitizers in photocatalysis. Adapted and reprinted with permission from [15], Copyright (2016) Elsevier.
Figure 2. Molecular structures of different dyes commonly employed as sensitizers in photocatalysis. Adapted and reprinted with permission from [15], Copyright (2016) Elsevier.
Catalysts 10 00103 g002
Figure 3. MCTPP (5-monocarboxyphenyl-10,15,20-triphenylporphyrin) and its possible self-metalation product after adsorption on rutile TiO2(110) surface. Adapted and reprinted with permission from [33], Copyright (2018) American Chemical Society (ACS).
Figure 3. MCTPP (5-monocarboxyphenyl-10,15,20-triphenylporphyrin) and its possible self-metalation product after adsorption on rutile TiO2(110) surface. Adapted and reprinted with permission from [33], Copyright (2018) American Chemical Society (ACS).
Catalysts 10 00103 g003
Figure 4. Proposed mechanism for the visible-light-induced selective oxidation of sulfide into sulfoxide with air on alizarin red S (ARS)-sensitized TiO2 with TEMPO ((2,2,6,6-tetramethylpiperidin-1-yl)oxyl) as a redox mediator. Adapted and reprinted with permission from [49], Copyright (2018) Wiley.
Figure 4. Proposed mechanism for the visible-light-induced selective oxidation of sulfide into sulfoxide with air on alizarin red S (ARS)-sensitized TiO2 with TEMPO ((2,2,6,6-tetramethylpiperidin-1-yl)oxyl) as a redox mediator. Adapted and reprinted with permission from [49], Copyright (2018) Wiley.
Catalysts 10 00103 g004
Figure 5. Structure of donor-acceptor dyes with phenothiazine as electron donor, functionalized with an alkyl chain (a) or a glucose derivative (b), thiophene spacers, and cyanoacrylic acid-based acceptor-anchoring groups. Adapted and reprinted with permissions from [58], Copyright (2015) Wiley and [59], Copyright (2018) American Chemical Society (ACS).
Figure 5. Structure of donor-acceptor dyes with phenothiazine as electron donor, functionalized with an alkyl chain (a) or a glucose derivative (b), thiophene spacers, and cyanoacrylic acid-based acceptor-anchoring groups. Adapted and reprinted with permissions from [58], Copyright (2015) Wiley and [59], Copyright (2018) American Chemical Society (ACS).
Catalysts 10 00103 g005
Figure 6. Geometries of Ti-catechol complexes at TiO2 anatase (101) surface: (a) molecular adsorption, (b) monodentate, (c) bidentate bridging, (d) bidentate chelating coordination; (e) UV-visible absorption spectra of TiO2 nanoparticles sensitized by acetylacetone (Acac), catechol (Cat), dopamine (DOPA), salicylic acid (Sal), and hydroxamic acid (Hydr). Adapted and reprinted with permissions from [104] and [101], Copyright (2012 and 2016) American Chemical Society (ACS).
Figure 6. Geometries of Ti-catechol complexes at TiO2 anatase (101) surface: (a) molecular adsorption, (b) monodentate, (c) bidentate bridging, (d) bidentate chelating coordination; (e) UV-visible absorption spectra of TiO2 nanoparticles sensitized by acetylacetone (Acac), catechol (Cat), dopamine (DOPA), salicylic acid (Sal), and hydroxamic acid (Hydr). Adapted and reprinted with permissions from [104] and [101], Copyright (2012 and 2016) American Chemical Society (ACS).
Catalysts 10 00103 g006
Scheme 3. Illustrations of the electron transfer mechanisms driving visible light photocatalytic activity in (a) TiO2-dopamine (DA) and TiO2-polydopamine (pDA) systems (LMCT and dye-sensitization, respectively); (b) bismuth-based semiconductors surface-modified by CT complexes with dihydroxyphenyl molecules. Adapted and reprinted with permission from [124,127], Copyright (2017 and 2020) Elsevier.
Scheme 3. Illustrations of the electron transfer mechanisms driving visible light photocatalytic activity in (a) TiO2-dopamine (DA) and TiO2-polydopamine (pDA) systems (LMCT and dye-sensitization, respectively); (b) bismuth-based semiconductors surface-modified by CT complexes with dihydroxyphenyl molecules. Adapted and reprinted with permission from [124,127], Copyright (2017 and 2020) Elsevier.
Catalysts 10 00103 sch003
Figure 7. Structure of the most common carbon-based nanostructured materials.
Figure 7. Structure of the most common carbon-based nanostructured materials.
Catalysts 10 00103 g007
Figure 8. Main roles of carbon-based materials (a) and fundamental strategies for tailoring their photocatalytic properties (b). Adapted and reprinted with permission from [160], Copyright (2016) Elsevier.
Figure 8. Main roles of carbon-based materials (a) and fundamental strategies for tailoring their photocatalytic properties (b). Adapted and reprinted with permission from [160], Copyright (2016) Elsevier.
Catalysts 10 00103 g008
Figure 9. Schematic illustration of the preparation method to obtain carbon-based/TiO2 nanohybrids. Adapted and reprinted with permission from [164], Copyright (2016) Elsevier.
Figure 9. Schematic illustration of the preparation method to obtain carbon-based/TiO2 nanohybrids. Adapted and reprinted with permission from [164], Copyright (2016) Elsevier.
Catalysts 10 00103 g009
Figure 10. Schemes of photoinduced charge transfer processes occurring in photocatalytic metal-organic frameworks (MOFs): (a) light excitation of a node; (b) light excitation of a photo-active ligand (dye) in semiconductor-type MOF (c) Charge transfer and (d) Energy transfer processes in dye active MOFs; (e) dual excitation phenomena in the presence of semiconductor nodes and photoactive dye molecules into the same MOF [1], Copyright (2019) Elsevier.
Figure 10. Schemes of photoinduced charge transfer processes occurring in photocatalytic metal-organic frameworks (MOFs): (a) light excitation of a node; (b) light excitation of a photo-active ligand (dye) in semiconductor-type MOF (c) Charge transfer and (d) Energy transfer processes in dye active MOFs; (e) dual excitation phenomena in the presence of semiconductor nodes and photoactive dye molecules into the same MOF [1], Copyright (2019) Elsevier.
Catalysts 10 00103 g010
Figure 11. The activity of porphyrin-based MOF towards CO2 capture and reduction. Adapted and reprinted with permission from [230], Copyright (2015) American Chemical Society.
Figure 11. The activity of porphyrin-based MOF towards CO2 capture and reduction. Adapted and reprinted with permission from [230], Copyright (2015) American Chemical Society.
Catalysts 10 00103 g011
Figure 12. Schematic illustration of metal-organic framework (MOF)-based filter (MOFilter) for integrated air cleaning. Adapted and reprinted with permission from [232], Copyright (2019) Springer Nature.
Figure 12. Schematic illustration of metal-organic framework (MOF)-based filter (MOFilter) for integrated air cleaning. Adapted and reprinted with permission from [232], Copyright (2019) Springer Nature.
Catalysts 10 00103 g012
Figure 13. Scheme of bonding features and electronic properties of MOF/COF (covalent organic framework) hybrid materials and their involvement in the photooxidation of aromatic alcohols. Adapted and reprinted with permission from [225], Copyright (2020) Elsevier.
Figure 13. Scheme of bonding features and electronic properties of MOF/COF (covalent organic framework) hybrid materials and their involvement in the photooxidation of aromatic alcohols. Adapted and reprinted with permission from [225], Copyright (2020) Elsevier.
Catalysts 10 00103 g013
Figure 14. Scheme of the synthesis approach of Cu@UiO-66 composite and reaction mechanism for partial oxidation of benzyl alcohol under visible light irradiation. Adapted and reprinted with permission from [252], Copyright (2019) Elsevier.
Figure 14. Scheme of the synthesis approach of Cu@UiO-66 composite and reaction mechanism for partial oxidation of benzyl alcohol under visible light irradiation. Adapted and reprinted with permission from [252], Copyright (2019) Elsevier.
Catalysts 10 00103 g014
Figure 15. Molecular structures of the most used conducting polymers combined with the semiconducting metal oxides.
Figure 15. Molecular structures of the most used conducting polymers combined with the semiconducting metal oxides.
Catalysts 10 00103 g015
Figure 16. Schematic illustration of the mechanism of action of hybrid materials obtained by the combination of conductive polymers to semiconductor metal oxides.
Figure 16. Schematic illustration of the mechanism of action of hybrid materials obtained by the combination of conductive polymers to semiconductor metal oxides.
Catalysts 10 00103 g016
Figure 17. Proposed photocatalytic mechanism according to the band alignment (a). Atomic schematic diagram, illustrating the synergistic effect between defects and polyaniline (PANI) (b). Adapted and reprinted with permission from [285], Copyright (2014) American Chemical Society.
Figure 17. Proposed photocatalytic mechanism according to the band alignment (a). Atomic schematic diagram, illustrating the synergistic effect between defects and polyaniline (PANI) (b). Adapted and reprinted with permission from [285], Copyright (2014) American Chemical Society.
Catalysts 10 00103 g017
Figure 18. Scheme of α-alkylations reactions catalyzed by hybrid MAPbI3, perovskites. Reprinted with permission from [312], Copyright (2019) American Chemical Society.
Figure 18. Scheme of α-alkylations reactions catalyzed by hybrid MAPbI3, perovskites. Reprinted with permission from [312], Copyright (2019) American Chemical Society.
Catalysts 10 00103 g018
Figure 19. Comparison of H2 evolution in aqueous HI solutions containing MAPbI3 and Ni3C decorated MAPbI3. Reprinted with permission from [322], Copyright (2019) American Chemical Society.
Figure 19. Comparison of H2 evolution in aqueous HI solutions containing MAPbI3 and Ni3C decorated MAPbI3. Reprinted with permission from [322], Copyright (2019) American Chemical Society.
Catalysts 10 00103 g019
Figure 20. UV-Vis spectra (a) and schematic representation (b) of the MA3Bi2I9 powder for H2 evolution. Reprinted with permission from [309], Copyright (2019) American Chemical Society.
Figure 20. UV-Vis spectra (a) and schematic representation (b) of the MA3Bi2I9 powder for H2 evolution. Reprinted with permission from [309], Copyright (2019) American Chemical Society.
Catalysts 10 00103 g020
Figure 21. Panel (a) The mechanism of natural photosynthesis, in which P680 indicates the pigment (chlorophyll) that can absorb 680 nm light in photosystem II (PS II), and P680* is the excited state of P680. P700 indicates the pigment (chlorophyll) that can absorb 700 nm light in photosystem I (PS I), while P700* is the excited state of P700 (Z-scheme. Available online: http://commons.wikimedia.org/wiki/File:Z-scheme.png, accessed on 9 January 2020). Panel (b) Schematic illustration of direct Z-scheme heterojunction, in which CB, VB, PCI, and PCII indicate the conduction band, valence band, photocatalyst I, and photocatalyst II, respectively. Reprinted with permission [325]. Copyright Elsevier 2018.
Figure 21. Panel (a) The mechanism of natural photosynthesis, in which P680 indicates the pigment (chlorophyll) that can absorb 680 nm light in photosystem II (PS II), and P680* is the excited state of P680. P700 indicates the pigment (chlorophyll) that can absorb 700 nm light in photosystem I (PS I), while P700* is the excited state of P700 (Z-scheme. Available online: http://commons.wikimedia.org/wiki/File:Z-scheme.png, accessed on 9 January 2020). Panel (b) Schematic illustration of direct Z-scheme heterojunction, in which CB, VB, PCI, and PCII indicate the conduction band, valence band, photocatalyst I, and photocatalyst II, respectively. Reprinted with permission [325]. Copyright Elsevier 2018.
Catalysts 10 00103 g021
Table 1. Principal groups of organic dyes used in photosensitization, some relevant members, relative radiation absorption range, and the semiconductors coupled to them reported in this review.
Table 1. Principal groups of organic dyes used in photosensitization, some relevant members, relative radiation absorption range, and the semiconductors coupled to them reported in this review.
Dye GroupRelevant DyesAbsorption RangeSemiconductors
PorphyrinsTCPP400–450, 500–700 nmTiO2, ZnO, Fe3O4
Phthalocyanines 600–800 nmTiO2
Xanthene basedEosin Y, erythrosin, rhodamine B, fluorescein, rose Bengal 400–600 nmTiO2, ZnO, SnO2, BiPO4, BiOCl
Anthraquinone basedAlizarin, alizarin red S400–600 nmTiO2, ZnO
D-π-A (donor-bridge-acceptor)Triarylamines, phenothiazines, carbazoles, coumarins (donor groups) variableTiO2
Cationic dyesMethylene blue, thionine, Nile blue A600–700 nmTiO2, ZnWO4
FlavonoidsAnthocyanines, quercetin 580–700 nmTiO2, ZnO, BiOF
Table 2. Dye-sensitized systems and their photocatalytic activity data under visible irradiation or solar light irradiation. The best results achieved in each work are reported.
Table 2. Dye-sensitized systems and their photocatalytic activity data under visible irradiation or solar light irradiation. The best results achieved in each work are reported.
SemiconductorSensitizerPreparationApplicationProductivityRef.
TiO2 TCPP porphyrin (0.05–1 wt%) Wet impregnation Rhodamine B (RhB, 10−5 M) degradation 93% removal after 10 h (90 W)[26]
TiO2 nanotubes arrayH2TClPP porphyrinAdsorption RhB (5 mg/L) degradation 93% removal after 4 h (300 W)[27]
TiO2 nanorodsTwo substituted porphyrinsAdsorption 4-nitrophenol (10−4 M) degradation 100% removal in ca.
1 h (400 W)
[28]
TiO2 and N-doped TiO2 HematoporphyrinAdsorption Methyl orange (MO, 5 mg/L) degradation 33% on TiO2, 88.5% on N-TiO2 after 3 h[29]
TiO2-SiO2 core-shell NPs Carboxy-TCPP, Chorin e6Covalent graftingMO (10 mg/L) degradation 91% removal with Chlorin e6 after 2 h (200 W)[34]
Fe3O4@SiO2@TiO2AminoporphyrinCovalent graftingRhB (10−5 M) degradation 96% removal after 1 h (with H2O2, 300 W)[35]
Pt/TiO2 (anatase) Pyridyl-porphyrin dimer, trimer (30 µmol/g ca.) Adsorption H2 production (ascorbic acid solution)1 mmol/g/h (300 W) [36]
TiO2 (anatase) Phthalocyanines (H2Pc)Adsorption Cr(VI) (10 mg/L) reduction 83%–100% removal in 2.5 h (UV light)[37]
TiO2 Phthalocyanines Hydrothermal methodMethylene blue (MB, 3 mg/L) degradation 100% removal in ca.
2 h (250 W)
[38]
Co-doped TiO2 Phthalocyanine derivative (1 wt%)Sol-gel MO (10 mg/L) degradation, Cr(VI) (10 mg/L) reduction 60% MO removal, 71% Cr(VI) removal in 2.5 h (UV)[39]
Pt/TiO2Calixarene-based dyeSol-gel H2 production (TEOA solution)12.2 mmol/g/h (300 W)[40]
TiO2 (P25) on fibreglass Eosin Y In situ addition (2 mg/L)Diclofenac (30 mg/L) degradation 49% removal in 5 min (solar simulator)[41]
Pt/TiO2 (P25)Eosin Y (7.5–10 wt%)Adsorption H2 production (TEOA solution)2.7 mmol/g/h (visible and solar)[42]
Co complex/TiO2 (P25) Eosin Y In situ addition (4×10−5 M)H2 production (TEA solution)0.45 mmol/g/h ca. (300 W) (TON = 90)[43]
Ni/TiO2 nanotubesEosin Y In situ addition (10−4 M)H2 production (TEOA solution)1.2 mmol/g/h ca. (400 W)[44]
TiO2-MoSx Eosin Y In situ addition (10−4–10−3 M)H2 production (TEOA solution)1.4 mmol/g/h ca. (solar simulator)[45]
TiO2 (P25)Eosin Y (0.1 mmol/g)Adsorption Benzyl ethers oxidation 100% conv., 71% yield (12 h, visible LED) [46]
TiO2 (anatase)Eosin Y, others Adsorption Benzyl alcohols oxidation 80% yield ca. (1 h, with TEMPO, visible LED) [47]
TiO2 (anatase)Alizarin red S (0.20 or 0.025 mmol/g) Adsorption Sulfides oxidation 80%–90% conv. and sel. (varying time, TEMPO, 300 W or LED) [48,49]
TiO2 (anatase)Alizarin red S (0.20 mmol/g) Adsorption Amines oxidation 90% ca. conv., sel. > 90% (varying time, TEMPO, 3 W) [50]
TiO2 (anatase, P25), ZnO Alizarin red S (0.025 mmol/g), Flavin, Thionine Adsorption Alcohols oxidation 60%–90% conv., 98% ca. sel. (0.5 h, 3 W LEDs) [51]
TiO2 (anatase)Anthraquinones In situ addition (0.020 mmol/g) Sulfides oxidation 80%–90% conv. and sel. (varying time, TEMPO, 3 W LED)[52]
TiO2 (P25)Alizarin red S, othersAdsorptionDimethylanilines and maleimides cyclization91% yield (20 h, 12 W LED)[53]
Pt/TiO2 (anatase) Substituted carbazole dyesAdsorptionH2 production (TEA solution)42 mmol/g ca. after 24 h (300 W) [54]
Pt/TiO2 Substituted phenothiazine or triphenylamine dyes (0.1 mmol/g)AdsorptionH2 production (TEOA solution)70–90 mmol/g ca. after 6 h (400 W) [55,56]
Pt/mesoporous TiO2 Tetrathiafulvalene dyes (0.1 mmol/g)AdsorptionH2 production (TEOA or glycerol solution)24.5 mmol/g/h (glycerol) (450 W) [67]
Pt/TiO2 (P25) Triphenylamine dyes (10 μmol/g)AdsorptionH2 production (TEOA solution)0.2 mmol/g/h ca. (150 W) [57]
Pt/TiO2 (P25) Phenothiazine dyes (30 or 60 μmol/g)AdsorptionH2 production (TEOA solution)2.4 or 1.4 mmol/g ca. after 20 h (150 W) [58,59]
Pt/TiO2 film (anatase) Phenothiazine and carbazole dyes (0.12 mmol/g)AdsorptionH2 production (TEOA solution)0.14 mmol/g ca. after 24 h (200 W) [60]
Pt/porous TiO2Phenothiazine dye (30 μmol/g)AdsorptionH2 production (ascorbic acid solution)90 mmol/g ca. after 10 h (300 W) [62]
Pt/TiO2Perylene dye (30 μmol/g)AdsorptionH2 production (TEOA solution)1 mmol/h/g ca.
(300 W)
[63]
Cu2Ws4/TiO2 (P25)Triphenylamine dyesAdsorptionH2 production (TEOA solution)1.4 mmol/g/h (300 W) [64]
Pt/TiO2 (P25)Chlorophyll-indoline dyad (8 μmol/g each)AdsorptionH2 production (ascorbic acid solution)4.2 mmol/g ca. after 6 h (350 W) [65]
Pt/rGO-TiO2 (P25)Squaraine dye In situ addition (50 μM)H2 production (EDTA solution)0.040 mmol/h ca.
(300 W)
[69]
Pt/TiO2; Re(I) complex/TiO2Squaraine dyes (20–480 μmol/g)AdsorptionH2 production (ascorbic acid, TEOA or EDTA solution); CO2 reduction 4.2 mmol H2/g ca. after 2 h; 0.3 mmol CO/g ca. after 10 h (60 W LED) [68]
Sol-gel TiO2 Curcumin (1–10 wt%)ImpregnationMB (10−4 M) and Orange II (2.5 × 10−5 M) degradation 90% removal ca. after 8 h (18 W)[70]
TiO2/activated carbonAnthocyanin (5 μmol/g)AdsorptionPhenoazopyridine degradation 88% removal in 1.5 h (solar simulator)[71]
TiO2 (anatase) Chlorophyll, flavonoids Adsorption MB (5–15 mg/L) degradation 90% removal ca. after 3 h (150 W)[72]
TiO2 (P25) Sudan black B Adsorption Phenol (5 mg/L) degradation 96% removal in 100 min (400 W)[78]
TiO2 nanotubes Brilliant green Adsorption LDPE films degradation 48% weight loss in 45 days (18 W)[77]
ZnO Protoporphyrin, squaraine dyeAdsorptionAcridine orange (5 × 10−5 M) degradation 70% removal ca. in
1 h
[80]
ZnO nanorods Fluorescein, protoporphyrin, PTCDA Adsorption MB, RhB, Eosin Y, resorcinol (30 or 50 mg/L) degradation 100% ca. removal after 1 h (dyes) or 5 h (resorcinol) (solar simulator)[81]
ZnO Curcumin Precipitation, adsorption Perylene, fluoranthene, crysene (12 mg/L) degradation90% removal ca. (varying time, sunlight)[82]
ZnO Anthocyanin (0.5 µmol/g ca.)Adsorption Escherichia coli (5 × 105 cfu/L) disinfection; paracetamol, MB degradation 92% bacteria loss, 70% paracetamol, 80% MB removal (1.5 h, solar sim.)[83]
ZnO 4ABBN (diazo dye)Precipitation, sol-gelH2 production (MeOH solution)102 mmol/g after 4 h (400 W)[84]
Pt/ZnO nanorods Eosin Y and rose BengalIn situ additionH2 production (TEOA solution)0.52 L/m2/h after 8 h (300 W)[85]
ZnOAlizarin red S Adsorption Alcohols oxidation 70%–99% ca. conv. after 7 h (TEMPO, 70 W) [86]
Sb-doped SnO2Eosin Y In situ addition (0.1 g/g)H2 production (TEOA solution)0.25 mmol/h/g (300 W)[87]
ZnWO4 MB In situ addition (5–15 mg/L)Cr(VI) (40 mg/L) reduction 97% removal in 45 min (UV-vis, 400 W) [88]
BiPO4 nanorods Eosin Y (5–30 wt%)Adsorption MB, 2,4-dichorophenol (10 mg/L) degradation75% MB removal in 80 min, 33% 2,4-DCP removal in 3 h (300 W)[89]
BiOCl RhB In situ additionRhB, methyl orange degradation 100% RhB removal in 15 min, 90% ca. MO in 2 h (500 W)[90]
BiOF Quercetin PrecipitationMethyl orange, benzotriazole (20 mg/L) degradation81% MO removal in 1.5 h, 72% BT removal in 9 h
(4 × 28 W)
[91]
Table 3. Systems sensitized by ligand-to-metal charge transfer (LMCT) complexes and their photocatalytic activity data under visible irradiation or solar light irradiation. The best results achieved in each work are reported.
Table 3. Systems sensitized by ligand-to-metal charge transfer (LMCT) complexes and their photocatalytic activity data under visible irradiation or solar light irradiation. The best results achieved in each work are reported.
SemiconductorSensitizerPreparationApplicationProductivityRef.
Pt/TiO2 (anatase)Catechol and derivatives
(80 µmol/g)
Adsorption H2 production (TEOA solution)0.60 mmol/g after 5 h (100 W)[109]
TiO2 Catechol and substituted catechols In situ addition (0.8 mol%)Amines oxidation 50%–60% conv., >90% sel. in 45 min (with TEMPO, 3 W)[107]
TiO2 Salicylic acid and derivativesAdsorption, in situ addition (48 µmol/g)Amines oxidation 90%–98% conv., 98% ca. sel. (varying time, TEMPO, 3 W)[115]
TiO2 (rutile) nanorods Catechol, salicylic, phthalic, terephthalic acid In situ addition (25 mg/g)Catechol, salicylic, phthalic, terephthalic acid degradation (self-sensitized)15%–20% TOC decrease after 30 min [102]
TiO2 film Catechol Impregnation E. coli, S. aureus (105 CFU/mL) inactivation 1000-fold bacteria CFU decrease (10 mW)[126]
Mesoporous TiO2Catechol (polymerized)
(0.5–3.0 wt%)
Adsorption Cr(VI) (30 mg/L) reduction; H2 production (TEOA solution)100% Cr(VI) removal in 20 min (150 W); 10.9 mmol/h/g H2 (solar light)[125]
TiO2 (anatase)2,3-dihydroxy-naphthalene
(0.1–2.5 wt%)
ImpregnationAldehydes or nitroaromatics reduction 80%–99% conv. (varying time, blue LED)[111,112]
TiO2 2-naphthol
(0.5–2.0 wt%)
Adsorption Cr(VI) (20–50 mg/L) reduction 100% removal in 1 h (150 W)[122]
TiO2 (anatase)2-naphthol (0.5–2.0 wt%) Adsorption RhB (2 × 10−5 M) degradation; As(III) (2 mg/L) removal 100% RhB removal in 2 h, 82% As(III) in 1 h (300 W)[121]
Pt/TiO2 (P25) Ascorbic acid, Zn phthalocyanine Impregnation, in situ addition H2 production (ascorbic acid solution) 16 mmol/h/g ca. (300 W)[114]
TiO2 (P25) Glucose, other saccharidesIn situ addition (10−4–1 M) Cr(VI) (50–100 µM) reduction; O2 reduction to H2O2 100% Cr(VI) removal in 40 min; 30 µmol/g H2O2 in 2.5 h (300 W)[103]
TiO2-rGOGlucose In situ addition (20 wt%) Cr(VI) (30 mg/L) reduction 100% Cr(VI) removal in 1 h (150 W)[118]
TiO2 (P25), Pt/TiO2 (P25)Dopamine and derivatives (polymerized)Self-polymerization Cr(VI) reduction; As(III) oxidation; O2 reduction to H2O2; degradation of CCl4, dyes, and 4-chlorophenol100% Cr(VI) removal in 1.5 h; 80% As(III) in 3 h; ca. 200 µmol/g H2O2 in 2 h; low dyes and 4-CP removal (300 W)[124]
TiO2 (anatase) Dopamine, polydopamine Adsorption, self-polymerization RhB (10 mg/L) degradation 100% RhB removal in 1.5 h (50 W)[123]
TiO2 Oxalate, other carboxylates, glucose In situ addition (500 µM) Ranitidine (100 µM) and other pharmaceuticals degradation 90% ranitidine removal after 6 h (oxalate, 300 W) [128]
TiO2 (P25) Alanine, glycine, histidine, phenylalanine, serine In situ addition (100 µM)Amino acids self-sensitized degradation Partial mineralization to NH4+ (with peroxymonosulfate, 300 W)[129]
Pt, Pd, or Rh/TiO2 (anatase, rutile, mixed phase)Pyridine In situ addition Pyridine-cyclohexane cross-coupling 3% conv., 99% sel. after 48 h [130]
TiO2 (anatase) 6,7-dihydroxy-2-methylisoquinoliniumImpregnation Amines photocyanation96% yield after 3 h (100 W blue LED), 90% after 62 h (51 W IR)[132]
TiO2 (anatase) Zwitterionic (Good’s) buffersIn situ addition (50 mM) Methyl orange (10 mg/L) degradation100% removal after 1.5 h (60 W)[133]
TiO22-methoxyethanolSolvothermal processRhB (20 mg/L) degradation100% removal after 2.5 h (300 W)[134]
TiO2EthynylpyreneHydrothermal processMB (20 mg/L) degradation100% removal after 15 min (UV)[136]
TiO2, CaTiO3, BaTiO3, SrTiO3Catechol, salicylic acid, 2,3-dihydroxy-naphthalene Adsorption Terephthalic acid (6 mM) oxidation Increased •OH photogeneration by titanates (150 W)[113]
ErFeO3Tartrate Adsorption MB degradation100% removal in 1.5 h (UV) [150]
Nb2O5Toluene In situ addition Toluene oxidation (self-sensitized)18 µmol benzaldehyde after 3 h (200 W)[151]
BiOIO3, BiOCl, BiOBr, Bi2WO6, Bi2MoO6, BiPO4, Bi2O2CO3Hydroxybenzoic acid and derivatives (25 mg/g)Adsorption Hydroxybenzoic acids (20 mg/L) self-sensitized degradation 90% ca. removal after 2 h on BiOIO3 (1000 W)[127]
Table 4. Carbon-based hybrid nanomaterials and their photocatalytic activity data under visible irradiation or solar light irradiation. The best results achieved in each work are reported.
Table 4. Carbon-based hybrid nanomaterials and their photocatalytic activity data under visible irradiation or solar light irradiation. The best results achieved in each work are reported.
Hybrid Carbon-Based MaterialPreparationApplicationProductivityRef.
biochar/TiO2Ultrasound-assisted, surfactant-free synthesisPhenol (50 ppm) degradation;
Selective methanol (0.9 vol% in air) photooxidation
64.1% (UV light);
33.6% (visible light);
~80%–90% methyl formate production
[159]
CNTs/TiO2 nanostructuresSolvothermal methodRhodamine B (10−5 mol/L) removal and degradation
(under simulated sunlight)
34.71 mg/g;
~95% after 90 min
[164]
rGO/TiO2 nanostructuresHydrothermal methodRhodamine B (10−5 mol/L) removal and degradation
(under simulated sunlight)
14.07 mg/g;
100% after 90 min
[164]
Carbon/TiO2 nanostructuresHydrothermal methodRhodamine B (10−5 mol/L) degradation (under simulated sunlight)74.85 mg/g;
~95% after 90 min
[164]
rGO sheet/N, S-CQD/TiO2-P25 nanotubesAlkaline hydrothermal and physical stirring Methyl orange (10 ppm) degradation (under visible light)92% after 4 h[165]
rGO/TiO2 nanotubesElectrophoretic depositionCO2 conversion to methane (under simulated sunlight)~6 ppm/cm2×h [166]
rGO/TiO2 nanoparticlesHydrothermal methodBio-recalcitrant photo-removal 5.093 × 103 ppm TOC/kWh [168]
Graphene/CQDs/TiO2 nanoparticlesHydrothermal method and physical stirring Methyl blue (3 ppm) degradation
(under visible light)
100% after 90 min[171]
CQDs/ZnO nanostructured spheresMicrowave-assisted solvothermal synthesisPhotoelectrodes for quantum-dots sensitized solar cells−1.3 × 10−4–8.7 × 10−5 A/cm2 of current density[172]
Soot/TiO2 nanostructuresChemical vapor depositionRhodamine B degradation (under UV light)80%[174]
Carbon/TiO2-P25 nanostructuresPhysical mixing under stirringAir pollutant (5 ppm) removal (under UV light)47% toluene conversion;
≥60% acetaldehyde degradation
[175]
Graphene nano-platelets/P25-TiO2 nanoparticlesOne-step hydrothermal methodMethylene blue (0.01 g/L) degradation, NO conversion (under UV light);~95% after 120 min
~80% after 150 h
[176]
TiO2–In2O3/g–C3N4 nanostructuresSolvothermal methodRhodamine B degradation (under visible light);
H2 production
~99%
8.6 μmol/h
[178]
N-doped Carbon/TiO2Mixing reaction and thermolysis methodOxygen reduction reaction4 μmol NOx removal[180]
Exfoliated rGO/TiO2 nanostructuresIonothermal treatmentH2 production by water splitting in water–ethanol system3 mmol/g[181]
N-doped CQDs/TiO2-P25 nanoparticlesHydrothermal treatment and physical stirring processH2 production by water splitting using methanol (25 vol%) as sacrificial agent 21.2–58.6 nmol h−1 (under visible light);
2.70–9.80 nmol h−1 (under full-light)
[184]
Au/rGO/TiO2 yolk@shell hybrid nanostructuresPrecursors reacting in aqueous/organic mixtureH2 production by water splitting using methanol as a sacrificial agent;
Rhodamine B (2 × 10−5 mol/L) degradation
(under simulated sunlight)
309 mmol h−1 g−1 (visible light);
462 mmol h−1 g−1 (simulated sunlight);
~99.6% after 100 min
[185]
N-doped CNTs/ZnO nanostructuresHomogeneous chemical vapor depositionPhenol (20 mg/L) degradation
(under simulated sunlight)
100% after 20 min[190]
CNTs/ZnO nanorodssolvothermal synthesisRhodamine B (2 × 10−2 mg/mL) degradation
(under simulated sunlight)
90% after 2 h[193]
CNTs/ZnO nanostructuresHydrothermal treatmentRhodamine B (10 mg/mL) degradation
(under simulated sunlight)
100% after 2 h[194]
Lignin-based carbon/ZnO nanoparticlesSolvothermal synthesisMethyl orange and rhodamine B (15 mg/L) degradation (under simulated sunlight)99.9% after 30 min;
79.2% after 50 min
[195]
N-doped C-sheet/ZnO NPsHydrothermal treatmentMethylene blue (50 mL) degradation
(under UV light)
>95% after 60 min [197]
ZnO/CQDs nanocompositesHydrothermal treatmentRhodamine B (5 × 10−6 M) degradation
(under visible light)
83% within 105 min
(annealed at 80 °C);
94% within 105 min
(annealed at 200 °C)
[198]
CQDs/α-Fe2O3 nanostructuresHydrothermal methodElectrochemical performance as photoanodes2.1 mA cm−2 of photocurrent density[201]
CQDs/α-Fe2O3 nanostructuresCo-precipitationIndigo carmin (25 mg/L) degradation
(under visible light)
100% after 60 min[202]
Porous g-C3N4/α-Fe2O3 nanocompositeCalcination methodRhodamine B (10 mg/L) degradation (under visible light)91.1% after 15 min[213]
rGO/Fe3O4 nanohybridsOne-pot microwave approachElectrode material for high-performance supercapacitors.91.4% cycling performance after 190 cycles[204]
Graphene/Fe3O4 nanocompositeHydrothermal methodMethyl orange (20 mg/L) degradation (under UV)99.24%[205]
rGO/Fe3O4 hybrid nanocompositesSolvothermal synthesisMethylene blue (12.5 mg/L) degradation (under visible light)74% after 60 min[206]
rGO/CuxO nanorodsSolvothermal methodCO2 conversion to CH3OH (under visible light) 862 mmol g−1
with rGO/Cu2O;
1228 mmol g−1 with rGO/CuO
[207]
CQDs/Cu2O heterostructuresOne-step ultrasonic methodCO2 conversion to CH3OH (under solar-light)55.7 μmol g−1 h−1[208]
biomass CQDs/Bi2WO6/Cu2O heterostructuresHydrothermal methodTetracycline (20 mg/L)
degradation (under visible light)
80% in 90 min[209]
Carbon layer/In2O3 nanobelts Hydrothermal synthesisCO2 reduction to CO and CH4126.6 μmol h–1;
27.9 μmol h–1
[211]
Table 5. Strategies to improve metal-organic frameworks (MOFs) photocatalytic activity and performance data of obtained systems under visible irradiation or solar light irradiation. The best results achieved in each work are reported.
Table 5. Strategies to improve metal-organic frameworks (MOFs) photocatalytic activity and performance data of obtained systems under visible irradiation or solar light irradiation. The best results achieved in each work are reported.
StrategySystem/SampleApplicationProductivityRef.
High Absorption Photoactive Ligands
Metalloporphyrins ligandsCu2+-PCN224 MOFTreatment of infected wounds99.71% against Staphyloccoccus aureus after 20 min[226]
Ligands containing specific functional groups (–NH2, –NO2, –SH, and –OH) Ti(IV)-based MOF (NTU-9) Rhodamine B (RhB) and methylene blue (MB) degradation 100% after 40 min[227]
Highly π-conjugated ligands1,4-bis(2-[4-carboxyphenyl]ethynyl)benzene
Zr6O4(OH)4(CO2)12
Methylene blue (MB) and methyl orange (MO) degradationMB 100% MO 83%
after 3 h (higher activity than TiO2)
[228]
Tetrakis(4-carboxyphenyl)-porphyrin (TCPP) ligand(TCPP)—benzoic acid—titanium MOF (PCN-22)Benzyl alcohol oxidation to benzaldehydeConversion 28%;
selectivity 100%
after 2 h
[229]
Tetrakis(4-carboxyphenyl)-porphyrin (H2TCPP) ligandzirconium–porphyrin MOF, PCN-222 (also known as MOF-545 or MMPF-6)CO2 reduction30 μmol in 10 h[230]
Bipyridyl ligandscadmium–organic frameworks Cd(dcbpy) (dcbpy = 2,2′-bipyridine-5,5′-dicarboxylate), Cd(bdc)(bpy) (bdc = 1,4-benzenedicarboxylate; bpy = 2,2′-bipyridyl), and Cd(bdc)(2Me-bpy) (2Me-bpy = 4,4′-dimethyl-2,2′-bipyridyl)Oxidation of benzylamine 2748 μmol·g−1·h−1;
1217 μmol·g–1·h−1;
555 μmol·g−1·h–1;
Selectivity to immine >99%
in 7 h
[231]
Imidazolate-Zn(II) charge transfer complexZinc-imidazolate MOF (ZIF-8)Antibacterial >99.9999% inactivation efficiency[232]
Metalloporphyrins ligandsPtII tetrakis(4-carboxyphenyl)-
porphyrin (PtTCPP)—Cu2(COO)4 2D-MOF
H2 production 11,320 μmol g−1 h−1 [218]
Metal pyridine complexes Ti-MOF-Ru(tpy)2
Ru(tpy): bis(4′-(4-carboxyphenyl)-terpyridine)Ru(II) complex as an organic linker
H2 production(photoreforming)183.3 μmol·h−1·g−1
(triethanolamine as a sacrificial agent, Pt as co-catalyst)
[233]
Mixed ligandH2TCPP⊂(I-)Meim-UiO-66
(H2TCPP: tetrakis (4-carboxyphenyl) porphyrin, I−:iodide ion, Me: methyl, im: imidazolium)
Cr(VI) reduction13.3 mgCr(VI)·min−1·g−1[234]
Ligand Functionalization
Amine functionalizationHierarchical NH2-MIL-125(Ti)Rhodamine B removal83% in 2 h[235]
Amine functionalizationNH2-MIL-125(Ti)H2 production (photoreforming)60 μmol·h−1·g−1
TEOA as a sacrificial agent
[236]
Amine functionalization TiO2/1,4-benzenedicarboxylate (bdc) MOF (MIL-125)Oxidation of benzyl alcohol to benzaldehyde200 μmol·h−1·g−1[237]
Amine functionalization NH2-MIL-125(Ti) Oxidation of amines to imines73% conversion, selectivity towards N-benzylidene benzylamine 86% (in acetonitrile) in 12 h[238]
Amine functionalization NH2-MIL-88B (Fe)
NH2–MIL-101 (Fe)
Cr(VI) reduction100% in 1 h[239]
Amide functionalization [M(L4)0.5(phen)(H2O)]n.2nH2O [M = Zn; Cd]
phen = 1,10-phenantroline
Rhodamine B decomposition79%–90% in 8 h[240]
Co–thiolate moietiesZr-DMBD-Co (DMBD = 2,5-dimercapto-1,4-benzenedicarboxylate)CO2 reduction to CO 3000 μmol·h−1·g−1[241]
Covalent organic framework (COF)Benzoic acid-modified covalent triazine-based frameworks (B-CTF-1) covalently bonded with NH2-MIL-125(Ti) or NH2-UiO-66(Zr) MOFs H2 production 360 μmol·h−1·g−1[242]
Rhodamine B (RhB) dye[Me2NH2][Tb3(dcpcpt)3(HCOO)]·DMF·15H2O (Tb-dcpcpt = 3-(3,5-dicarboxylphenyl)-5-(4-carboxylphenl)-1H-1,2,4-triazole,
DMF = N,N’-dimethylformamide)
Nitrofuran and quinolone;
Antibiotic detection
Nitrofuran (99 ppb);
Quinolone 1 mM
[243]
Ru(II)-polypyridyl complexes Ru(bpy)3@MIL-125Selective oxidative coupling of benzylamines75% conversion in 3 h[244]
Erythrosin B dyeErythrosin B dye sensitized UiO-66 Zr-MOFH2 production (photoreforming)400 μmol·h−1·g−1
methanol as sacrificial reagent,
Pt co-catalyst
[245]
Engineering of Inorganic Building Units
Replacement of Zr(IV) to Ti(IV)NH2-MIL-125(Ti)Degradation of emerging pollutants (acetaminophen)100% in 2 h[246]
Ce(III) dopingPd/CeMIL-101H2 production from ammonia borane 50,000 μmol·h−1·g−1[247]
Encapsulation of Photocatalytic Active Species (MOF Composites Photocatalysts)
PtZrO2-UIO66-MOFRhB degradation100% in 40 min[248]
PdPd@Cu(II)-MOFBenzyl alcohol oxidation reactionConversion 93%–99%;
Selectivity >99% (to benzaldehydes)
[249]
PtCalixarene-sensitized Pt@UiO-66-NH2H2 production (photoreforming)1528 μmol·h−1·g−1
methanol as sacrificial reagent
[250]
Pt/AuM-doped NH2-MIL-125(Ti)
(M = Pt and Au)
H2 and HCOO productionPt-MOF
H2: 562 μmol·h−1·g−1
HCOO: 32.5 μmol·h−1·g−1;
Au-MOF
H2: 100 μmol·h−1·g−1
HCOO: 8 μmol·h−1·g−1
[251]
Cu deposited and encapsulatedCu/Cu@UiO-66Oxidation of benzyl alcohol49.50%[252]
NiNiII(bpet)(H2O)2]@Ru-UiO-67 (Ni@Ru-UiO-67)CO2 reduction to CO196.8 μmol·h−1·g−1[253]
ReI(CO)3(BPYDC)Cl, BPYDC = 2,2′-bipyridine-5,5′- dicarboxylate + AgAg⊂Re3-MOFReI(CO)3(BPYDC)Cl, BPYDC = 2,2′-bipyridine-5,5′-dicarboxylate linked to a zirconium MOF, UiO-67 (Ren-MOF)CO2 reductionTurnover number (TON) 1.4 × 10−2 h−1[254]
C3N4g-C3N4/ZIF-8CO2 reduction;
H2 production
CH3OH production
0.75 μmol h−1 g−1,
309.5 μmol·L−1·h−1·g−1;
TEOA as a sacrificial reagent
[255]
TiO2 hollow nanospheres (HNSP)double-shell TiO2@ZIF-8 hollow nanospheres H2 production254.2  μmol h−1 g−1
(TEOA, methanol as sacrificial reagent)
[256]
TiO2 nanofibersTiO2 nanofibers with NH2–UiO-66CO2 reduction 0.8 μmol h−1 g−1[257]
Titanate nanotube (TNTs)BUC-21/TNTs
BUC21: bis-1,3-dibenzyl-2-oxo-4,5-imidazolidinedicarboxylic acid—Zn2+ 2D MOF
Cr(VI) reduction and adsorptive Cr(III) removal99% in 40 min[258]
TiO2TiO2@NH2-MIL-88B(Fe)Cr(VI) reduction100% removal in 25 min[259]
BiOBr nanosheetNH2-UiO-66/BiOBr
NH2-MIL-125(Ti)/BOBr
Tetracycline hydrochloride (TC) degradation and Cr(VI) reduction; TC decomposition and Cr(VI) removal: 75% in 120 min;[260]
Norfloxacin (NOR) degradationNOR removal efficiency 93.6%[261]
Ag3VO4, Ag2WO4IL-125-NH2/ Ag3VO4 Ag2WO4Methylene blue (MB) and rhodamine B (RhB) degradation 90% MB removal;
100% RhB removal in 60 min
[262]
CdSCdS/ZIF-8 CO2 reduction to COCO evolution 93.3 μmol h−1 g−1[263]
MoS2MoS2/ZIF-8Ciprofloxacin (CIP) and tetracycline hydrochloride (TC) degradationCIP 95%;
TC 80% in 180 min
[264]
CH3NH3PbI3 (MAPbI3) perovskite QDsMAPbI3@PCN-221(Fe0.2)CO2 reduction 1559 μmol g−1;
CO2 converted to CO (34%) and CH4 (66%)
[265]
Activated carbon (AC)AC/MIL-88B (Fe)Reactive Red198 (RR198)
degradation
99% in 30 min[266]
rGOrGO-Zr-based UiO-66(NH2)Cr(VI) reduction100% in 100 min[267]
rGOMIL-53(Fe)Cr(VI) reduction;
RhB degradation
Cr(VI) 100%;
RhB 83% in 80 min
[268]
Table 6. Main properties of semiconducting polymer/metal oxides hybrid nanomaterials and their photocatalytic activity data under visible irradiation or solar light irradiation. The best results achieved in each work are reported.
Table 6. Main properties of semiconducting polymer/metal oxides hybrid nanomaterials and their photocatalytic activity data under visible irradiation or solar light irradiation. The best results achieved in each work are reported.
SemiconductorConductive PolymerPreparationApplicationProductivityRef.
n-type TiO2 nanowellsPANIIn situ polymerization in ice-water bath PhotodetectorsResponsivity of 37 μA/W–3.6 mA/W[282]
3D-flowerlike TiO2 NPsPANIIn situ oxidative polymerizationCongo red and methyl orange degradation (under UV and sunlight)95% and 98% (UV);
90% and 96% (sunlight) after 2 h
[283]
Defective ZnO hybridsPANIChemisorption methodMetyl Orange and 4-chlorophenol degradation (under UV light)97% after 100 min
82% after 2 h
[285]
ZnO nanorods PANIChemisorption methodMethylene blue degradation (under visible light)98% after 60 min[284]
ZrO2 nanoparticlesPANIIn situ chemical oxidative polymerizationTrichlorophenol degradation (under simulated sunlight)~80% after 240 min [287]
Hollow CoFe2O4 nanofibersPANIElectrospinning, in-situ oxidative polymerizationMethyl orange degradation (under visible light)85% after 120 min[288]
SnO2 nanoparticlesPANIIn situ chemical polymerizationCr(VI) reduction in aqueous medium (under visible light)60%–100% after 210 min as function of concentration[289]
TiO2 nanoparticlesPPyIn situ chemical polymerizationDiclofenac and 4-chlorophenol degradation (under simulated sunlight)90% and 40%
after 60 min
[290]
Pd/TiO2 nanoparticlesPPySimultaneous photodeposition methodH2 production (under UV-visible light)1400 μmol after 2 h (UV light);
3.5 μmol after 2 h (visible light)
[291]
ZnO microrod arraysPPyElectropolymerizationMethylene blue degradation (under sunlight)85.95% after 3 h[292]
ZnO nanoparticlesPPyIn situ chemical polymerizationDiclofenac degradation (under simulated sunlight)81% after 60 min[294]
ZnWO4 nanoparticlesPPyIn situ chemical polymerizationPhotoelectrodesTafel slopes of 76 and 84 mV dec−1 for HER and OER [295]
Fe2O3 nanostructuresPPyHydrothermal methodGas sensing0.1–10 ppm of NO2[296]
NiS/core-shell Fe3O4 nanoparticlesPPyIn situ chemical oxidative polymerizationCephalexin degradation (under sunlight) 80% after 30 min[297]
NiS,NiO/core-shell Fe3O4 nanoparticlesPPyIn situ chemical oxidative polymerizationNaproxen degradation (under visible light)87% after 30 min[299]
TiO2 nanotubesPEDOTElectrochemical polymerization and spin-coatingPhotoelectrodes12.9–20 mA/cm2 of photocurrent density[304]
CNTs/ZnO nanotubesPEDOTIn situ oxidative polymerizationMethyl orange degradation (under visible light)75% after 200 min[305]
TiO2/ZnO nanostructuresPEDOTIn situ chemical polymerizationHybrid nanodevices52 Scm−1 of electrical conductivity; 0.83 cm2 V–1 S–1 of charge carrier mobility [306]
Table 7. Hybrid perovskites and data on their photocatalytic activity under visible irradiation or solar light irradiation. The best results achieved in each work are reported.
Table 7. Hybrid perovskites and data on their photocatalytic activity under visible irradiation or solar light irradiation. The best results achieved in each work are reported.
System/SamplePreparationApplicationProductivityRef.
Aniline-capped MAPbBr3 (MA = methylammonium ion, CH3NH3 +) One-pot crystallizationMalachite green (MaG) degradation90% in 70 min[308]
APbBr3 (A = Cs or methylammonium)One-pot crystallizationα-alkylation of aldehydes99%[312]
MAPb(I1−xBrx)3One-pot crystallizationH2 production (HI splitting)1471 μmol·h−1·g−1 for 62 cycles, 4 h each run[314]
MAPbBr3−xIx Light-assisted halide-exchange methodH2 production (HI splitting)1021 μmol·h−1·g−1;
with Pt (co-catalyst)
2604 μmol·h−1·g−1;
For 6 cycles, 5 h each
[315]
MAPbBr3HydrothermalH2 production (HI splitting)34.4 μmol·h−1·g−1;
with Pt (co-catalyst)
60 μmol·h−1·g−1
[316]
H[(Me)3-TIB]2Pb5I17 (TIB = 1,3,5-tris(1-imidazolyl)benzene)SolvothermalMO, RhB, and MB degradation 100% in 60 min (MO, RhB) or 30 min (MB)[317]
GO/MAPbBr3 PrecipitationCO2 reduction to CO1.05 μmol cm−2 h−1[318]
rGO/MAPbI3HydrothermalH2 production (HI splitting)14 μmol·h−1·g−1 for 200 cycles (10 h each cycle)[319]
Pt/TiO2 nanoparticles - MAPbI3HydrothermalH2 production (HI splitting)7277 μmol·h−1·g−1[320]
MAPbI3-TiO2 nanorod array (TNAs)Spin coating + heating at 70 °CH2 production (HI splitting)33.3 μmol cm-2 h-1[321]
Ni3C/MAPbI3SolvothermalH2 production (HI splitting)2362 μmol·h−1·g−1[322]
2-(aminomethyl pyridine)SbI5 Sol-GelH2 production106.7 μmol·h−1·g−1
with GO (co-catalyst) 185.8 μmol·h−1·g−1
[311]
(CH3NH3)3Bi2I9HydrothermalH2 production in HI aqueous solution12.19 μmol·h−1·g−1
with Pt 169.21 μmol·h−1·g−1
[309]
(CH3NH3)3Bi2I9PrecipitationRhB, MBMB 70%;
RB 100%
in 130 min
[323]

Share and Cite

MDPI and ACS Style

Luciani, G.; Imparato, C.; Vitiello, G. Photosensitive Hybrid Nanostructured Materials: The Big Challenges for Sunlight Capture. Catalysts 2020, 10, 103. https://0-doi-org.brum.beds.ac.uk/10.3390/catal10010103

AMA Style

Luciani G, Imparato C, Vitiello G. Photosensitive Hybrid Nanostructured Materials: The Big Challenges for Sunlight Capture. Catalysts. 2020; 10(1):103. https://0-doi-org.brum.beds.ac.uk/10.3390/catal10010103

Chicago/Turabian Style

Luciani, Giuseppina, Claudio Imparato, and Giuseppe Vitiello. 2020. "Photosensitive Hybrid Nanostructured Materials: The Big Challenges for Sunlight Capture" Catalysts 10, no. 1: 103. https://0-doi-org.brum.beds.ac.uk/10.3390/catal10010103

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

Article Metrics

Back to TopTop