Next Article in Journal / Special Issue
An Innovative Technique for Energy Assessment of a Highly Efficient Photovoltaic Module
Previous Article in Journal / Special Issue
Scalable Screen-Printed TiO2 Compact Layers for Fully Printable Carbon-Based Perovskite Solar Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Two-Dimensional Photocatalysts for Energy and Environmental Applications

by
Thaleia Ioannidou
1,
Maria Anagnostopoulou
2 and
Konstantinos C. Christoforidis
1,2,*
1
Department of Environmental Engineering, Democritus University of Thrace, Vasilisis Sofias 12, 67100 Xanthi, Greece
2
Institute of Chemistry for Energy, Environment and Health, Centre National de la Recherche Scientifique (CNRS), University of Strasbourg, 25 rue Becquerel, 67087 Strasbourg, France
*
Author to whom correspondence should be addressed.
Solar 2022, 2(2), 305-320; https://doi.org/10.3390/solar2020017
Submission received: 30 April 2022 / Revised: 27 May 2022 / Accepted: 6 June 2022 / Published: 10 June 2022
(This article belongs to the Special Issue Solar Technologies—A Snapshot of the Editorial Board)

Abstract

:
The depletion of fossil fuels and onset of global warming dictate the achievement of efficient technologies for clean and renewable energy sources. The conversion of solar energy into chemical energy plays a vital role both in energy production and environmental protection. A photocatalytic approach for H2 production and CO2 reduction has been identified as a promising alternative for clean energy production and CO2 conversion. In this process, the most critical parameter that controls efficiency is the development of a photocatalyst. Two-dimensional nanomaterials have gained considerable attention due to the unique properties that arise from their morphology. In this paper, examples on the development of different 2D structures as photocatalysts in H2 production and CO2 reduction are discussed and a perspective on the challenges and required improvements is given.

1. Introduction

Recent advances in nanotechnology and, in particular, precise materials engineering have allowed the development of many different nanomaterials that present unique physiochemical properties. Research in this specific topic has revealed the importance of the precise control of composition, morphology, structure and size in directing their optical, electronic, chemical, mechanical and thermal properties. All these properties play a crucial role in the efficiency of a specific application, especially in the field of clean energy production and environmental depollution. Therefore, the recent advances in the field of materials engineering open up new applications for known materials.
In our way to achieve sustainability, clean energy production and environmental pollution are among the greatest challenges that we are facing. Despite the great efforts of the last decades, there are still many technological barriers that need to be overcome. Many different approaches have been developed and applied in all the above-mentioned issues, including biotic [1,2] and abiotic methodologies [3,4,5]. Among the different technologies, abiotic methodologies, including catalytic processes that are based on the development of novel nanomaterials, have gained importance in recent years.
For all the above-mentioned fields, of particular interest is the development of novel photo-sensitive nanomaterials applied as photocatalysts. As in all technologies that rely on the use of nanomaterials, precise materials engineering has played a key role in the observed progress in these specific fields, allowing the development materials with specific properties via controlling composition, morphology, structure and size in directing their optical, electronic, chemical, mechanical and thermal properties. Since efficiency is closely related to materials properties, advancements in materials engineering are a prerequisite towards improving efficiency.
In addition to composition, morphology at the nanoscale plays an important role in the final properties that regulate efficiency. Among the different nanomaterials developed, of particular importance are two-dimensional materials. This stands especially for applications that are critically dependent on the available surface area together with advantageous charge-handling properties. Therefore, materials possessing a 2D structure have gained considerable attention in recent years. In order for a material to be characterized as 2D, it must contain a sheet-like morphology. The lateral size may vary from nanometers up to micrometers. In addition, the thickness of the layer should range from a single to few atomic thick layers. Their size constrains and the electron confinement in two dimensions offer them unique properties that are difficult to be found in their bulk counterparts [6].
The advancements presented in 2D materials and specially in those applied in light-triggered reactions are mostly linked with their improved physical, electronic and optical properties. This stands mostly for traditional photocatalysts and not for highly porous structures, such as metal–organic frameworks. Two-dimensional materials present a significantly larger specific surface area as compared with the corresponding bulk structures. A large specific surface area is one of the targeted properties in heterogeneous photocatalysis since both applications are related to surface reactions. Enhanced efficiency is linked with more available active sites, superior efficiency in light harvesting and mass transport and interaction with the substrate. In addition, the thickness of the structure may control the electronic properties via a multi-route approach. It may affect light absorption, the actual electronic band structure (a parameter that controls the performance of a specific reaction) and the availability of charges via controlling the abundance and stability of the photogenerated charges (electron and hole pairs, e/h+). Thin 2D structures offer shorter pathways for charges to reach the surface where reactions occur, decreasing the migration distance and thus allowing a faster transfer and higher availability of charges. This is critical considering the required solar energy flux to complete reactions such as water splitting and CO2 reduction [7,8]. An additional benefit of 2D structures is that they can also act as excellent supports for the development of multicomponent materials, allowing a higher interaction between the different parts.
In this review article, we focus on the recent progress on typical photosensitive single-phase 2D structures as well as composites and heterojunctions, applied as photocatalysts for sustainable energy production. Attention is paid to photocatalytic H2 production and CO2 reduction. Before discussing recent contributions in materials development, the fundamental principles and elementary steps in each application are explained.

2. Photocatalysis Based on 2D Materials

2.1. Fundamental Steps in Photocatalysis

Heterogeneous photocatalysis mimics the mechanism of natural photosynthetic centers. Under light irradiation, redox equivalents are formed and perform a chemical reaction. Lately, the term artificial photosynthesis has been used for any reaction that is initiated by light, ideally solar light, which stores energy (i.e., solar) in the form of chemical energy. As a consequence, the term solar fuel is used to characterize fuels produced using sunlight energy.
The advantage of the photocatalytic process is mostly reflected on the potential to use a free and endless energy, e.g., solar energy. The surface of the Earth annually receives solar energy (e.g., 1.3 × 105 TW) that is, by far, greater than the estimated future global energy demands (28–35 TW per year global energy demand by 2050) [9]. In addition, in contrast to other chemical processes, it may be applied under ambient conditions (pressure and temperature). Furthermore, the process is truly sustainable when suitable substrates are used, i.e., water. Considering the different systems developed, the use of semiconductors (SCs) in the heterogeneous photocatalytic process offers the advantages of the easy separation and reuse of the catalyst compared with homogeneous photocatalysts.
Electron and hole pairs (e/h+) are formed in a SC when it is exposed to light greater than the energy difference between the bottom level of the conduction band (CB) and the top level of the valence band (VB). Ideally, the energy source must be solar irradiation in order to totally explore the sustainable character of the process. This mechanism adds the first restriction for the formation of the redox equivalents, that is, the development of photocatalysts activated by solar energy. Considering that a major part of solar energy is in the visible region, photocatalysts that are activated under visible light irradiation are preferred. However, this should not be pursuit in the expense of the targeted energy of the charges. e and h+ pairs may undertake several paths upon their formation. Charge recombination either in the bulk or on the surface of the catalyst must be avoided. Therefore, strategies in this direction have been studied, including compositional modification and size constrains, but shape control as well. Figure 1 presents the elemental steps and fundamental reactions that take place on a photocatalyst focusing on H2 production and CO2 reduction. An important parameter that must always be taken into account is the energy of the photogenerated charges, which depends solely on the actual band structure of the SC.
Solar fuel production using photocatalysis has been explored focusing on the production of H2 and mainly C1 and C2 products from CO2 reduction. H2 may be produced via the photocatalytic route from the so-called water splitting process that converts water into H2 and O2. Water oxidation is a highly demanding step since it is a four e process. Alternatively, H2 may be produced by the photoreforming of oxygenated organic substrates. The reactions are given in the following formulas:
Overall water splitting process:
H 2 O H 2 + 1 / 2 O 2
Photoreforming process:
C x H y O z + ( 2 x z ) H 2 O ( 2 x z + y / 2 ) H 2 + xCO 2
In the latter, instead of water, an organic substrate is oxidized. This offers two key advantages. Firstly, the organic substrate may be oxidized easier than water, eliminating the rate limiting step of the water splitting process. Secondly, the oxidation of the organic substrate may provide protons for H2 production. However, the proper choice of organic substate defines the sustainability of the process. The importance of H2 as a fuel is expressed by its zero greenhouse gas emissions when combined with O2 in the fuel cell technology. Therefore, green hydrogen is produced via the photocatalytic process using solar energy and water as a substrate. On the other hand, CO2 may be reduced into a variety of valuable chemicals used as fuels (i.e., CH4) or in other important aspects in the chemical industry, i.e., production of high-added-value chemicals. Depending on the targeted CO2 reduction product, the process may be characterized as carbon neutral or carbon negative. The latter takes the carbon-neutral concept a step further since it removes more CO2 than it emits.
Both photocatalytic CO2 reduction and H2 productions share similarities and were demonstrated over the same period [11,12]. However, H2 evolution reaction has progressed significantly more compared to CO2. This is mostly attributed to the challenges of the latter reaction. The multi-electron transfer process, C–C coupling and hydrogenation reactions required in the CO2 reduction reaction highlight the more challenging thermodynamic and kinetic barriers. Furthermore, the photocatalytic conversion of CO2 may result in different products (Table 1), a fact that may require additional complex separation processes or the development of highly selective photocatalysts. The low solubility of CO2 in water is another practical issue. High pressures or the use of chemicals with a high CO2 affinity (i.e., solvents) may be applied to avoid this issue, increasing the process complexity, but also reducing the environmental character of the overall process. Finally, we should also highlight that the two reactions are competing reactions since both require the use of photogenerated e. Considering that H2O is the ideal electron donor in a CO2 photocatalytic reduction reaction, the competition between H2O and CO2 for e consumption highlights the importance of selectivity towards accomplishing a specific reaction.
Overall, based on the above analysis, an efficient photocatalyst must contain several characteristics. Among others, the catalyst must have a suitable band structure in order to utilize solar light and generate charges of a specific energy, generate sufficient charges avoiding recombination phenomena, present high specific surface area and exposed active sites and also be selective towards a specific reaction, especially in the case of CO2 reduction. By the definition given above, 2D structures present a high specific surface area and, therefore, increased surface active sites as well as favorable charge-handling properties compared with their bulk counterparts. These characteristics make 2D structures as promising materials in photocatalysis and sparked intensive research over the last years.
Given that the photocatalytic activity greatly depends on the physicochemical properties of the photocatalyst as well as on the specific conditions used in the reaction and especially the intensity and the wavelength of the incident radiation, it is rather difficult to compare the performance of the different systems tested. Production rates are usually reported in μmol h−1 g−1; however, the normalization of the activity to the specific surface area of the catalyst may be also applied. Quantum yields (QY) or apparent quantum yields (AQY) have been introduced as a way to compare the efficiency of photocatalysts tested in different setups. In this process, the number of the reacted electrons (based on the product, i.e., 2e per H2) over the number of absorbed/incident photons is reported. However, instead of the “apparent,” the so-called “true” quantum efficiency should be reported. The latter process offers the benefit of providing the accurate values of the efficiency since, for the optical properties of the catalyst, its behavior in the media of the reaction is taken into account [13].

2.2. Heterojunctions

An efficient approach to improve several important aspects of a photocatalyst is the coupling with another SC, i.e., the formation of a heterojunction. Visible light response may be given to a wide Eg SC through its coupling with a narrow Eg SCs. This may offer a better control of the light absorption properties compared with other competing approaches, such as the widely studied doping process. Furthermore, many examples in the literature have demonstrated that efficient charge separation may be also achieved in heterojunctions [14]. However, special attention must be given on the actual band structure of the coupled SCs in order to realize such improvements. Charge separation may be accomplished via two mechanisms. In both cases, the fundamental condition that must be fulfilled is that the CB of the one SC must lie within the band gap of the second. The formation of type-II heterojunctions (Figure 2) is the most common case. In such systems, the difference in the band energy level controls the migration of the charges. h+ are directed in one SC, while e follow the opposite direction. Therefore, efficient charge separation occurs. The major issue in type-II heterojunctions is that e and h+ lose part of their energy since they are spontaneously transferred from bands of high energy to those of lower. This shortcoming is bypassed when charge migration occurs via the Z-scheme mechanism (Figure 2), where e on the low energy CB recombines with h+ on the low energy VB. Usually, in such systems, a charge mediator may be required, allowing the recombination of e on the low energy CB with h+ on the less positive VB [15].
Charge separation in heterojunctions requires the formation of a tight interface between the different parts. Therefore, the large lateral size of 2D structures offers the benefit of providing a large contact area between the coupled systems [15]. Therefore, 2D structures may be used as excellent platform to be coupled with materials of other morphologies, but charge separation via interface is manifested when forming 2D/2D heterostructures [14]. Exciton dissociation is favored in 2D/2D heterojunctions even in the presence of a lattice mismatch compared with heterojunctions of different morphology [16].

3. Solar Fuel Production Using 2D Materials

Many different photocatalysts have been developed and applied in H2 production and CO2 reduction, of which TiO2 nanomaterials occupy the vast majority. Among the different TiO2 photocatalysts, 2D TiO2 structures are of particular importance. Facets with the lowest surface energy define the morphology of the crystal. Surfactants are mostly used to control the shape of the TiO2 crystal. In anatase TiO2, {101} surfaces are predominately exposed, resulting in a bipyramid structure. The formation of 2D anatase single crystal structures with preferentially exposed {001} facets has been achieved using fluorine as a capping agent under hydrothermal conditions [17,18]. Most importantly, the exposed surfaces may control the activity of the catalyst. For many reactions, it has been shown that {001} facets offer higher activity, a fact that has been attributed to the chemistry of the different facets and, in particular, the coordination number of the terminated Ti centers [19]. However, controversial results have been reported regarding the activity of anatase crystals with different exposed surfaces. The actual conditions of the synthesis affect important parameters of the catalyst. This could be the case for the different activity reported, a fact that highlights the importance of more studies being required in this field, including a complete physicochemical characterization to identify key material parameters. For example, Liu et al. reported that the exposed crystal facets in 2D anatase TiO2 together with the primary particle size and the proper removal of the capping agent control H2 production [20]. Structures with a smaller particle size after the complete removal of fluorine were significantly more active. A synergy was suggested for the enhanced activity, including an improved surface area and the shift of the CB minimum to more negative levels. Surface heterojunctions may be also formed by controlling the ratio of the co-exposed facets. This has been suggested using theoretical calculations on anatase TiO2 nanocrystals with co-exposed {001} and {101} facets [21]. The {001}/{101} ratio of the facets was controlled using different amounts of the capping agent and the spatial separation of the charges (Figure 3) was suggested as being the main contributor in the improved CO2 reduction reaction.
As a way to improve the main inherent drawback of anatase, i.e., being inactive under visible light irradiation, a doping method was applied to improve the electronic properties linked with light absorption. The nitrogen doping of anatase TiO2 sheets expanded light absorption to the visible region. The corresponding N-modified photocatalyst presented considerable activity in H2 evolution under pure visible light irradiation, while it improved activity under UV-Vis irradiation compared with the non-doped TiO2 [22]. Black titania is another example of improving visible light response. Such structures are usually developed through a hydrogenation treatment at high temperatures that induces surface defects’ formation. Defect formation may present strong morphology dependence as demonstrated in TiO2 nanosheets with {001} exposed facets [23]. Titanate 2D structures are another interesting class of materials applied in photo-triggered reactions. Rh doping in exfoliated protonated titanate nanosheets induced new mid-gap states and H2 production under UV-light irradiation was controlled by the dopant content [24]. The activity of the layered titanate nanosheets is critically affected by the chemical exfoliation as well as by the co-catalyst and the nanosheets have been proven active even using non-precious CdS quantum dots as co-catalysts [25].
In addition to attempts to control the properties of single-phase 2D TiO2 nanostructures, considerable attention has been paid to the construction of composite materials based on TiO2 sheets. She et al. coupled TiO2 nanosheets with 2D SnS2 [26]. The 2D/2D SnS2/TiO2 heterojunction was applied in CO2 reduction and exhibited a 10-times increased CH4 production compared with bare TiO2 nanosheets. The authors suggested that the observed significant enhancement of the activity originated from the reduced recombination rates of the charges due to the increment in the contact between the two parts. The enhancement of the photocatalytic H2 production from H2O/methanol solution was attributed to the same reason in composites of TiO2 with exposed {001} facets coupled with layered MoS2 [27]. Most importantly, the 2D/2D TiO2/MoS2 junction was more active than the TiO2 nanosheets when Pt was used as a co-catalyst. We obtained similar results in a 2D/2D heterojunction composed of TiO2 and carbon nitride (CN) nanosheets [14]. By applying a facile one-pot synthesis strategy, TiO2 nanocrystals with a specific morphology were deposited on ultrathin CN nanosheets (Figure 4). Composites of CN with TiO2 nanocrystals with a bipyramidal or nanosheet morphology were developed in the presence or the absence of a capping agent. Bare TiO2 nanosheets presented a significantly lower activity than the non-shaped TiO2 in gas-phase CO2 reduction. When both TiO2 structures were coupled with 2D CN nanosheets, they exhibited improvement in photoactivity compared with their individual counterparts. Surprisingly, the 2D/2D CN/TiO2 junction was more active than the corresponding composite bearing unshaped TiO2 (i.e., 2D/3D junction). Charge dynamic analysis using transient absorption spectroscopy revealed that interfacial charge transfer was favored in the 2D/2D structure, reducing charge recombination and increasing the availability of photo-induced electrons. This is attributed to the increased contact between the two-dimensional phases of the composite. This is a clear example of the benefits in charge separation efficiency in heterojunctions developed through the coupling of 2D structures.
In addition to metal oxides, metal sulfides and phosphides have also been coupled with 2D CN structures [28,29]. CN decorated with SnS2 particles presented an impressive 6305.18 μmol h−1 g−1 without the use of expensive noble metals as co-catalysts under pure visible light irradiation [28]. Three-dimensional NiCoP coupled with two-dimensional CN nanosheets presented improved H2 evolution rates compared with the corresponding CN photocatalyst bearing Pt as a co-catalyst. More importantly, the NiCoP/CN heterostructure presented a stable H2 production rate for up to 60 h [29]. The strong interaction between the two phases (CN and NiCoP) was ascribed to the formation of Co–N and Ni–N bonding states and was identified as the key parameter for the improved stability and activity. Two-dimensional heterojunctions composed of CN and BiOCl provided structural defects at the interface enhancing CO2 reduction [30]. Once again, efficient charge separation, which increased the availability of the charges, was the main reason for the improved photocatalytic activity.
Perovskite-based materials are another interesting family of photoactive materials that can possess a 2D structure. Inorganic as well as hybrid organic/inorganic perovskites have been developed and applied in photocatalysis. HCa2Nb3O10 nanosheet structures obtained via chemical exfoliation presented an impressive 16-fold enhanced H2 photocatalytic production compared to the bulk structure [31]. The enhancement was even greater in Rh-doped KCa2Nb3O10 nanosheets compared with the bulk layered oxide [32]. In these materials, the chemical composition controls the actual band structure, which may have a great effect in activity [33]. Shifting the CB to more positive levels results in reduced activity due to the less energetic CB e. Two-dimensional organic–inorganic hybrid perovskites were also proven to be active in H2 [34]. Recently, phenylmethylammonium lead-iodide 2D perovskites achieved a record-high solar-to-chemical conversion efficiency (H2 production) for hybrid perovskites [35]. The length of the perovskites’ organic cation determined the photocatalytic activity. Two-dimensional perovskites with the shortest organic cation exhibited a faster charge separation and transportation, producing 333 μmol h−1 of H2 using Pt as a co-catalyst in aqueous solutions under visible light irradiation.
Layered double hydroxides (LDHs) are another class of inorganic 2D materials that have been applied in CO2 photocatalytic reduction and H2 production. The composition of these layered materials is tunable, and they possess ion exchange abilities, a property that can be used to anchor chemical moieties and provide specific properties with applications in photocatalysis. Cu2O particles formed in situ from Cu-Zn-Cr LDHs were proven to be more active than the pristine material in the CO2 reduction to CO [36]. Cu2O nanoparticles promoted charge separation and acted as active centers for CO2 reduction. The suppression of the charge recombination rate was also suggested for the improved photoactivity in MTi-LDH (M:Ni, Zn, Mg) by anchoring TiO6 octahedra units on the surface of the LDH structure, a process that also modified the band structure [37]. In such systems, the content of the immobilized heterostructure plays a crucial role in the activity as demonstrated in CdSe nanocrystals deposited on ZnCr-LDH forming a heterojunction (Figure 5) [38].
Metal chalcogenides have been widely applied as photocatalysts in many reactions, including solar fuels production. One of their main advantages is the short band gap energy that allows the utilization of the visible part of solar irradiation. Among the different materials in this family, CdS has been extensively studied since it is an exceptional photocatalyst. Nevertheless, stability issues due to photocorrosion and high toxicity are the top drawbacks that need to be addressed. Over the last decade, molybdenum dichalcogenide has also been proven suitable for solar fuel production [39]. MoS2 has been used as a co-catalyst in H2 evolution by CdS [40]. A correlation between H2 production and MoS2 thickness was revealed. Photoactivity was improved via decreasing the MoS2 thickness. Such studies are very important since expensive noble metals are usually used as co-catalysts in H2 evolution reaction. The development of inexpensive co-catalysts, such as MoS2 or SnS2 [13], would be a significant step forward in the field. In a similar study, Yin et al. systematically studied the thickness and lateral size of MoS2 as co-catalysts deposited on CdS in H2 production [41]. The thickness was tuned using a facile exfoliation prosses followed by centrifugation. In addition to thickness, lateral size also affected activity. The decrease in lateral size increased H2 evolution. Improvements in activity were attributed to the shorter distance from the basal plane to the active sites at the edges of the MoS2 sheet. The MoS2 edges are considered highly active for many critical reactions, including CO2 reduction [42]. Hierarchical CdS-Au/MoS2 core/shell nanostructures composed of CdS nanorods and MoS2 nanosheets (Figure 6) presented excellent photocatalytic H2 production [43]. Improvements in visible light absorption, charge separation through the formation of a heterojunction and the increased number of active sites at the edges of the MoS2 structure due to the 2D morphology were responsible for the observed enhanced H2 production. The coupling of MoS2 with other nanosheets has also been performed. MoS2 nanosheets were deposited on 2D Cu2S (Figure 7) [44]. The morphology of both nanostructures played a crucial role in the activity, regulating the light trapping and scattering ability, charge separation and specific surface area. In this type of materials, exchange reactions may be applied to change the composition of the structure due to the presence of chalcogens and transition metals. However, such approaches are challenging. For example, CdIn2S4 nanosheets have been successfully supported on Co3O4, forming a p–n heterojunction [45]. The in situ synthesis process allowed the formation of a tide interface, improving stability and the selective reduction of CO2 to CO (5300 μmol h−1 g−1) under visible light irradiation due the efficient charge separation.
In addition to the nature of the coupled materials and the composition of the final heterojunctions, precise materials engineering may control the actual coupling of two SC and provide advanced systems. This is highlighted in the work of Hu et al., in which the effect of the selective deposition of 2D MoS2 on specific facets of TiO2 nanosheets was reported [46]. Photocatalytic H2 production was drastically improved when MoS2 were deposited on the {101} facets of TiO2. Through this approach, the {001} facets were exposed, while charge separation occurred due to the formation of a surface junction on TiO2 and the formation of a heterojunction between MoS2 and TiO2 (Figure 8). The coupling of MoS2 nanosheets with anatase/rutile mixed phase TiO2 nanofibers has also been prepared and tested in CO2 reduction (Figure 9) [47]. Two types of junctions were suggested by the authors: the formation of a heterojunction between TiO2 and MoS2 and a homojunction between the two phases of TiO2 (anatase and rutile). A 2D/2D heterojunction was also formed by growing MoS2 on TiO2 nanosheets [48]. The presence of MoS2 nanosheets as a co-catalyst presented a higher activity, but, most importantly, a higher selectivity for CO2 conversion to CH3OH than Pt, Au and Ag. The improved charge separation and the presence of highly active sites at the edges of MoS2 were responsible for the enhanced activity. MoS2 has also been effectively coupled with other SCs [49,50], while other 2D chalcogenides, such as Cu2S and ZnS [51,52,53], have also been proven suitable for solar fuel production.
Carbon nitride (CN) is another interesting material that can possess a 2D structure under certain circumstances [54]. It is a promising SC, suitable for both photocatalytic H2 production and CO2 reduction due to its highly negative CB level. The low-cost, facile preparation protocols, visible light response and the organic character that allows easy modification sparked intensive research in the development of novel CN structures since as-synthesized CN usually possesses a bulk structure, but 2D morphology can be obtained using post-synthetic methodologies, such as sonication [14], thermal oxidation etching [55] and the intercalation of salts [56]. Thermal conditions, including the atmosphere of the thermal treatment, may regulate the properties providing defects and shape control [57,58]. Ou et al. reported an interesting study, in which a 2D/2D CN/MoS2 layered nanojunction was more active in H2 production than using Pt as a co-catalyst [59]. Depending on the catalytic application, CN may be optimized towards enhancing activity. We have recently shown that, while H2 evolution is controlled by the available CB e linked with light absorption and charge separation, CO2 reduction is mostly regulated by CO2 adsorption on the surface of CN [57]. Charge dynamics analysis using transient absorption spectroscopy (TAS) under real catalytic conditions revealed that the temperature of the thermal treatment has a crucial effect on the reactivity with a sacrificial agent that enables higher H2 evolution rates. In addition, the thermal post-treatment of the preformed CN may result in the formation of intermediate mid-gap states, improving charge formation (Figure 10). Monometallic as well as bimetallic alloy nanoparticles (NPs) have also been efficiently coupled with 2D CN. In such heterostructures, of particular importance is the utilization of low-cost Cu NPs. AuCu alloy NPs on ultrathin porous CN nanosheets have proven particularly efficient in the photothermal CO2 reduction to ethanol [58]. The strong coupling of the two phases is a prerequisite to allow the migration of photogenerated charges.
In addition to the formation of junctions via the coupling of CN with other nanomaterials, CN-based homojunctions may also be formed. This originates from the organic character of the material that allows easy modification [60]. p–n homojunctions may be formed through the introduction of electron-acceptor groups into the CN structure [61]. This allows efficient charge separation, improving solar energy conversion.
Metal–organic frameworks (MOFs), a relatively new class of well crystalline and porous materials, are formed through the link between metal clusters and organic molecules. Structures containing MOFs have been applied for solar fuel production [14,18]. Their application in CO2 photocatalytic reduction reactions offers the advantage of increasing the CO2 concentration close to the active catalytic sites [18], eliminating one of the main rate-limiting steps of the process. Two-dimensional MOF structures have been also developed and applied in solar fuel production [62], despite such structures being difficult to obtained due to stability issues. A 2D/2D MOF/CN composite was developed by Liang et al. [63], improving H2 production. In this study, the MOF structure was involved in the electron transfer from the catalyst to the Pt co-catalyst. A 2D Ni-based MOF structure was coupled with reduced graphene oxide (rGO) through via Coulomb interaction forming a 2D/2D heterojunction [64]. A ruthenium complex (Ru(bpy)3) was used as light sensitizer and an electrostatic charge transfer from rGO to the MOF structure was suggested, promoting CO2 adsorption. Another interesting approach of MOFs in the field is their use as precursors for the development of novel photocatalysts. Co3O4 2D nanoparticles were developed using a modified version of ZIF-67 as a precursor (Figure 11) [65]. The obtained Co3O4 nanosheets were proven to be more active than the bulk Co3O4 in CO2 photocatalytic reduction under visible light. However, a light sensitizer was used in these tests. Improvements in activity were attributed to the presence of defective oxygen in the structure, increasing the ability to capture CO2.

4. Concluding Remarks and Directions

Both H2O and CO2 could be used as cheap and largely available feedstocks for chemical energy storage. The use of solar energy for fuel production through artificial photosynthesis is a simple process to obtain energy reach chemicals. The cheap photocatalytic conversion of CO2 could be also coupled with the Carbon Capture Utilization and Storage technology (CCUS), a process that must be included in any future carbon management approach. However, many improvements are required in order to meet industrial needs for large-scale applications. In such systems, efficiency is defined by activity, stability and reproducibility as well as selectivity. All these critical parameters are not related to the fundamentals of the process itself, but rather with the properties of the material. Two-dimensional materials have proven to be more efficient than their bulk counterparts, and a variety of materials and protocols have been developed. Precise engineering allowed the perfect control of the composition, size and morphology that regulate their electronic and textural properties.
For any photocatalytic application, research must focus on the development of solar-light-activated catalysts. In the specific case of H2 evolution reaction, future work should focus on improving light absorption properties and increase the availability of photogenerated e. These could be achieved by optimizing single-phase materials or by the development of multi-component systems through the synthesis of heterojunctions and heterostructures. In the specific case of composites, special attention must be paid to the development of the interface between the different parts, allowing their intimate coupling and interaction. On the other hand, CO2 photocatalytic reduction is a more challenging. In addition to improvements in light absorption and charge availability, issues related to mass transfer as well as selectivity towards the production of a specific product must be solved. Therefore, materials with a high selectivity and ability to absorb higher amounts of CO2 must be developed. In addition, the developed catalyst must be always fully characterized, including structural, morphological, compositional and electronic characterization as well as the utilization of techniques able to discriminate differences at the sub-nanometer scale at the bulk, but, most importantly, on the surface of the catalyst. Time-resolved in situ methodologies [13] and advanced operando [66] characterization techniques, both under real catalytic or synthesis conditions, could also significantly help in the identification of the reaction mechanism or understanding the development of the catalyst. This will give the opportunity to identify key material parameters and link them with catalytic activity. Based on the published contributions in the field of solar fuels in recent years, it seems that there is room for improvement in terms of efficiency and selectivity. Two-dimensional materials are expected to contribute in this direction given the documented advantages from recent years.

Author Contributions

Conceptualization, K.C.C.; methodology, T.I. and K.C.C.; formal analysis, T.I., M.A. and K.C.C.; investigation, T.I. and M.A.; resources, K.C.C.; writing—original draft preparation, T.I., M.A. and K.C.C.; writing—review and editing, T.I., M.A. and K.C.C.; visualization, T.I., M.A. and K.C.C.; supervision, K.C.C.; project administration, K.C.C.; funding acquisition, K.C.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work benefited from state aid managed by the French National Research Agency (ANR) under the Investments for the Future (program “Make Our Planet Great Again”) with reference ANR-18-MOPGA-0014. The research was partially funded by Democritus University of Thrace (project 82268).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Vasiliadou, I.A.; Berná, A.; Manchon, C.; Melero, J.A.; Martinez, F.; Esteve-Núñez, A.; Puyol, D. Biological and Bioelectrochemical Systems for Hydrogen Production and Carbon Fixation Using Purple Phototrophic Bacteria. Front. Energy Res. 2018, 6, 107. [Google Scholar] [CrossRef] [Green Version]
  2. Vasiliadou, I.A.; Molina, R.; Martinez, F.; Stathopoulou, P.M.; Tsiamis, G. Toxicity assessment of pharmaceutical compounds on mixed culture from activated sludge using respirometric technique: The role of microbial community structure. Sci. Total Environ. 2018, 630, 809–819. [Google Scholar] [CrossRef]
  3. Vasiliadou, I.A.; Sanchez-Vazquez, R.; Molina, R.; Martínez, F.; Melero, J.A.; Bautista, L.F.; Iglesias, J.; Morales, G. Biological removal of pharmaceutical compounds using white-rot fungi with concomitant FAME production of the residual biomass. J. Environ. Manag. 2016, 180, 228–237. [Google Scholar] [CrossRef] [PubMed]
  4. Christoforidis, K.C.; Pantazis, D.A.; Bonilla, L.L.; Bletsa, E.; Louloudi, M.; Deligiannakis, Y. Axial ligand effect on the catalytic activity of biomimetic Fe-porphyrin catalyst: An experimental and DFT study. J. Catal. 2016, 344, 768–777. [Google Scholar] [CrossRef]
  5. Ong, W.-J.; Tan, L.L.; Ng, Y.H.; Yong, S.T.; Chai, S.-P. Graphitic Carbon Nitride (g-C3N4)-Based Photocatalysts for Artificial Photosynthesis and Environmental Remediation: Are We a Step Closer to Achieving Sustainability? Chem. Rev. 2016, 116, 7159–7329. [Google Scholar] [CrossRef]
  6. Tan, C.; Cao, X.; Wu, X.J.; He, Q.; Yang, J.; Zhang, X.; Chen, J.; Zhao, W.; Han, S.; Nam, G.H.; et al. Recent Advances in Ultrathin Two-Dimensional Nanomaterials. Chem. Rev. 2017, 117, 6225–6331. [Google Scholar] [CrossRef]
  7. Ganguly, P.; Byrne, C.; Breen, A.; Pillai, S.C. Antimicrobial Activity of Photocatalysts: Fundamentals, Mechanisms, Kinetics and Recent Advances. Appl. Catal. B Env. 2018, 225, 51–75. [Google Scholar] [CrossRef]
  8. Ganguly, P.; Harb, M.; Cao, Z.; Cavallo, L.; Breen, A.; Dervin, S.; Dionysiou, D.D.; Pillai, S.C. 2D Nanomaterials for Photocatalytic Hydrogen Production. ACS Energy Lett. 2019, 4, 1687–1709. [Google Scholar] [CrossRef] [Green Version]
  9. Esswein, A.J.; Nocera, D.G. Hydrogen Production by Molecular Photocatalysis. Chem. Rev. 2007, 107, 4022–4047. [Google Scholar] [CrossRef]
  10. Christoforidis, K.C.; Fornasiero, P. Photocatalysis for Hydrogen Production and CO2 reduction: The case of copper-catalysts. ChemCatChem 2018, 11, 368–382. [Google Scholar] [CrossRef]
  11. Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef]
  12. Inoue, T.; Fujishima, A.; Konishi, S.; Honda, K. Photoelectrocatalytic reduction of carbon dioxide in aqueous suspensions of semiconductor powders. Nature 1979, 277, 637–638. [Google Scholar] [CrossRef]
  13. Barba-Nieto, I.; Christoforidis, K.C.; Fernández-García, M.; Kubacka, A. Promoting H2 photoproduction of TiO2-based materials by surface decoration with Pt nanoparticles and SnS2 nanoplatelets. Appl. Catal. B Env. 2020, 277, 119246. [Google Scholar] [CrossRef]
  14. Crake, A.; Christoforidis, K.C.; Godin, R.; Moss, B.; Kafizas, A.; Zafeiratos, S.; Durrant, J.R.; Petit, C. Titanium dioxide/carbon nitride nanosheet nanocomposites for gas phase CO2 photoreduction under UV-visible irradiation. Appl. Catal. B Environ. 2019, 242, 369–378. [Google Scholar] [CrossRef]
  15. Christoforidis, K.C. g-C3N4/Ag3PO4 based binary and ternary heterojunction for improved photocatalytic removal of organic pollutants. Int. J. Environ. Anal. Chem. 2021, 1–16. [Google Scholar] [CrossRef]
  16. Su, T.; Shao, Q.; Qin, Z.; Guo, Z.; Wu, Z. Role of Interfaces in Two-Dimensional Photocatalyst for Water Splitting. ACS Catal. 2018, 8, 2253–2276. [Google Scholar] [CrossRef]
  17. Yang, H.G.; Sun, C.H.; Qiao, S.Z.; Zou, J.; Liu, G.; Smith, S.C.; Cheng, H.M.; Lu, G.Q. Anatase TiO2 single crystals with a large percentage of reactive facets. Nature 2008, 453, 638–641. [Google Scholar] [CrossRef] [Green Version]
  18. Crake, A.; Christoforidis, K.C.; Kafizas, A.; Zafeiratos, S.; Petit, C. CO2 capture and photocatalytic reduction using bifunctional TiO2/MOF nanocomposites under UV–vis irradiation. Appl. Catal. B Environ. 2017, 210, 131–140. [Google Scholar] [CrossRef] [Green Version]
  19. Pan, J.; Liu, G.; Lu, G.Q.; Cheng, H.-M. On the True Photoreactivity Order of {001}, {010}, and {101} Facets of Anatase TiO2 Crystals. Angew. Chem. Int. Ed. 2011, 50, 2133–2137. [Google Scholar] [CrossRef]
  20. Liu, G.; Sun, C.; Yang, H.G.; Smith, S.C.; Wang, L.; Lu, G.Q.; Cheng, H.-M. Nanosized anatase TiO2 single crystals for enhanced photocatalytic activity. Chem. Commun. 2010, 46, 755–757. [Google Scholar] [CrossRef]
  21. Yu, J.; Low, J.; Xiao, W.; Zhou, P.; Jaroniec, M. Enhanced Photocatalytic CO2-Reduction Activity of Anatase TiO2 by Coexposed {001} and {101} Facets. J. Am. Chem. Soc. 2014, 136, 8839–8842. [Google Scholar] [PubMed]
  22. Liu, G.; Yang, H.G.; Wang, X.; Cheng, L.; Pan, J.; Lu, G.Q.; Cheng, H.-M. Visible Light Responsive Nitrogen Doped Anatase TiO2 Sheets with Dominant {001} Facets Derived from TiN. J. Am. Chem. Soc. 2009, 131, 12868–12869. [Google Scholar] [CrossRef] [PubMed]
  23. Chen, S.; Li, D.; Liu, Y.; Huang, W. Morphology-dependent defect structures and photocatalytic performance of hydrogenated anatase TiO2 nanocrystals. J. Catal. 2016, 341, 126–135. [Google Scholar]
  24. Soontornchaiyakul, W.; Fujimura, T.; Yano, N.; Kataoka, Y.; Sasai, R. Photocatalytic Hydrogen Evolution over Exfoliated Rh-Doped Titanate Nanosheets. ACS Omega 2020, 5, 9929–9936. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Kim, H.N.; Kim, T.W.; Kim, I.Y.; Hwang, S.J. Cocatalyst-Free Photocatalysts for Efficient Visible-Light-Induced H2 Production: Porous Assemblies of CdS Quantum Dots and Layered Titanate Nanosheets. Adv. Funct. Mater. 2011, 21, 3111–3118. [Google Scholar] [CrossRef]
  26. She, H.; Zhou, H.; Li, L.; Zhao, Z.; Jiang, M.; Huang, J.; Wang, L.; Wang, Q. Construction of a Two-Dimensional Composite Derived from TiO2 and SnS2 for Enhanced Photocatalytic Reduction of CO2 into CH4. ACS Sustain. Chem. Eng. 2019, 7, 650–659. [Google Scholar] [CrossRef]
  27. Yuan, Y.-J.; Ye, Z.-J.; Lu, H.-W.; Hu, B.; Li, Y.-H.; Chen, D.-Q.; Zhong, J.-S.; Yu, Z.-T.; Zou, Z.-G. Constructing Anatase TiO2 Nanosheets with Exposed (001) Facets/Layered MoS2 Two-Dimensional Nanojunctions for Enhanced Solar Hydrogen Generation. ACS Catal. 2016, 6, 532–541. [Google Scholar] [CrossRef]
  28. Jing, L.Q.; Chen, Z.G.; He, M.Q.; Xie, M.; Liu, J.; Xu, H.; Huang, S.; Huang, S. Different morphologies of SnS2 supported on 2D g-C3N4 for excellent and stable visible light photocatalytic hydrogen generation. ACS Sustain. Chem. Eng. 2018, 6, 5132–5141. [Google Scholar] [CrossRef]
  29. Li, C.; Wu, H.; Du, Y.; Xi, S.; Dong, H.; Wang, S.; Wang, Y. Mesoporous 3D/2D NiCoP/g-C3N4 heterostructure with Dual Co-N and Ni-N bonding states for boosting photocatalytic H2 production activity and stability. ACS Sustain. Chem. Eng. 2020, 8, 12934–12943. [Google Scholar] [CrossRef]
  30. Chen, Y.; Cao, Y.; Zhang, F.; Zou, Y.; Huang, Z.; Ye, L.; Zhou, Y. Interfacial oxygen vacancy engineered two-dimensional g-C3N4/BiOCl heterostructures with boosted photocatalytic conversion of CO2. ACS Appl. Energy Mater. 2020, 3, 4610–4618. [Google Scholar] [CrossRef]
  31. Ebina, Y.; Sasaki, T.; Harada, M.; Watanabe, M. Restacked Perovskite Nanosheets and Their Pt-Loaded Materials as Photocatalysts. Chem. Mater. 2002, 14, 4390–4395. [Google Scholar] [CrossRef]
  32. Okamoto, Y.; Ida, S.; Hyodo, J.; Hagiwara, H.; Ishihara, T. Synthesis and Photocatalytic Activity of Rhodium-Doped Calcium Niobate Nanosheets for Hydrogen Production from a Water/Methanol System without Cocatalyst Loading. J. Am. Chem. Soc. 2011, 133, 18034–18037. [Google Scholar] [CrossRef] [PubMed]
  33. Maeda, K.; Eguchi, M.; Oshima, T. Perovskite oxide nanosheets with tunable band-edge potentials and high photocatalytic hydrogen-evolution activity. Angew. Chem. Int. Ed. 2014, 53, 13164–13168. [Google Scholar] [CrossRef] [PubMed]
  34. Li, Y.; Li, L.; She, S.; Chen, S.; Liu, Y.; Xu, B.; Lee, F.-W.; Zhu, X. Multidimensional Perovskite for Visible Light Driven Hydrogen Production in Aqueous HI Solution ACS Appl. Energy Mater. 2022, 5, 207–213. [Google Scholar]
  35. Wang, H.; Zhang, H.; Wang, J.; Gao, Y.; Fan, F.; Wu, K.; Zong, X.; Li, C. Mechanistic Understanding of Efficient Photocatalytic H2 Evolution on Two-Dimensional Layered Lead Iodide Hybrid Perovskites. Angew. Chem. Int. Ed. 2021, 60, 7376–7381. [Google Scholar] [CrossRef]
  36. Jiang, H.; Katsumata, K.; Hong, J.; Yamaguchi, A.; Nakata, K.; Terashima, C.; Matsushita, N.; Miyauchi, M.; Fujishima, A. Photocatalytic reduction of CO2 on Cu2O-loaded Zn-Cr layered double hydroxides. Appl. Catal. B Environ. 2018, 224, 783–790. [Google Scholar] [CrossRef]
  37. Zhao, Y.; Chen, P.; Zhang, B.; Su, D.S.; Zhang, S.; Tian, L.; Lu, J.; Li, Z.; Cao, X.; Wang, B.; et al. Highly dispersed TiO6 units in a layered double hydroxide for water splitting. Chemistry 2012, 18, 11949–11958. [Google Scholar] [CrossRef]
  38. Zhang, G.; Lin, B.; Qiu, Y.; He, L.; Chen, Y.; Gao, B. Highly efficient visible-light-driven photocatalytic hydrogen generation by immobilizing CdSe nanocrystals on ZnCr-layered double hydroxide nanosheets. Int. J. Hydrog. Energy 2015, 40, 4758–4765. [Google Scholar]
  39. Singh, S.; Modak, A.; Pant, K.K.; Sinhamahapatra, A.; Biswas, P. MoS2–Nanosheets-Based Catalysts for Photocatalytic CO2 Reduction: A Review. ACS Appl. Nano Mater. 2021, 4, 8644–8667. [Google Scholar] [CrossRef]
  40. Chang, K.; Li, M.; Ouyang, W.T.S.; Li, P.; Liu, L.; Ye, J. Drastic Layer-Number-Dependent Activity Enhancement in Photocatalytic H2 Evolution over nMoS2/CdS (n ≥ 1) Under Visible Light. Adv. Energy Mater. 2015, 5, 1402279. [Google Scholar] [CrossRef]
  41. Yin, L.; Hai, X.; Chang, K.; Ichihara, F.; Ye., J. Synergetic Exfoliation and Lateral Size Engineering of MoS2 for Enhanced Photocatalytic Hydrogen Generation. Small 2018, 14, 1704153. [Google Scholar] [CrossRef] [PubMed]
  42. Asadi, M.; Kumar, B.; Behranginia, A.; Rosen, B.A.; Baskin, A.; Repnin, N.; Pisasale, D.; Phillips, P.; Zhu, W.; Haasch, R.; et al. Robust carbon dioxide reduction on molybdenum disulphide edges. Nat. Commun. 2014, 5, 4470. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Chava, R.; Do, J.; Kang, M. Smart Hybridization of Au Coupled CdS Nanorods with Few Layered MoS2 Nanosheets for High Performance Photocatalytic Hydrogen Evolution Reaction. ACS Sustain. Chem. Eng. 2018, 6, 6445–6457. [Google Scholar] [CrossRef]
  44. Zhang, X.; Guo, Y.; Tian, J.; Sun, B.; Liang, Z.; Xu, X.; Cui, H. Controllable growth of MoS2 nanosheets on novel Cu2S snowflakes with high photocatalytic activity. Appl. Catal. B Environ. 2018, 232, 355–364. [Google Scholar] [CrossRef]
  45. Huang, L.; Li, B.; Su, B.; Xiong, Z.; Zhang, C.; Hou, Y.; Ding, Z.; Wang., S. Fabrication of hierarchical Co3O4@CdIn2S4 p–n heterojunction photocatalysts for improved CO2 reduction with visible light. J. Mater. Chem. A 2020, 8, 7177–7183. [Google Scholar] [CrossRef]
  46. Hu, X.; Lu, S.; Tian, J.; Wei, N.; Song, X.; Wang, X.; Cui, H. The selective deposition of MoS2 nanosheets onto (101) facets of TiO2 nanosheets with exposed (001) facets and their enhanced photocatalytic H2 production. Appl. Catal. B Environ. 2019, 241, 329–337. [Google Scholar] [CrossRef]
  47. Xu, F.; Zhu, B.; Cheng, B.; Yu, J.; Xu, J. 1D/2D TiO2/MoS2 Hybrid Nanostructures for Enhanced Photocatalytic CO2 Reduction. Adv. Opt. Mater. 2018, 6, 1800911. [Google Scholar] [CrossRef] [Green Version]
  48. Tu, W.; Li, Y.; Kuai, L.; Zhou, Y.; Xu, Q.; Li, H.; Wang, X.; Xiao, M.; Zou, Z. Construction of unique two-dimensional MoS2-TiO2 hybrid nanojunctions: MoS2 as a promising cost-effective cocatalyst toward improved photocatalytic reduction of CO2 to methanol. Nanoscale 2017, 9, 9065–9070. [Google Scholar] [CrossRef]
  49. Kong, L.; Yan, J.; Liu, S. Carbonyl Linked Carbon Nitride Loading Few Layered MoS2 for Boosting Photocatalytic Hydrogen Generation. ACS Sustain. Chem. Eng. 2019, 7, 1389–1398. [Google Scholar] [CrossRef]
  50. Zhang, Z.; Huang, L.; Zhang, J.; Wang, F.; Xie, Y.; Shang, X.; Gu, Y.; Zhao, H.; Wang, X. In situ constructing interfacial contact MoS2/ZnIn2S4 heterostructure for enhancing solar photocatalytic hydrogen evolution. Appl. Catal. B Environ. 2018, 233, 112–119. [Google Scholar] [CrossRef]
  51. Zhou, Y.; Lei, Y.; Wang, D.; Chen, C.; Peng, Q.; Li, Y. Ultra-thin Cu2S nanosheets: Effective cocatalysts for photocatalytic hydrogen production. Chem. Commun. 2015, 51, 13305–13308. [Google Scholar] [CrossRef] [PubMed]
  52. Zhang, J.; Yu, J.; Zhang, Y.; Li, Q.; Gong, J.R. Visible Light Photocatalytic H2-Production Activity of CuS/ZnS Porous Nanosheets Based on Photoinduced Interfacial Charge Transfer. Nano Lett. 2011, 11, 4774–4779. [Google Scholar] [CrossRef] [PubMed]
  53. 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]
  54. Christoforidis, K.C.; Montini, T.; Fittipaldi, M.; Jaén, J.J.D.; Fornasiero, P. Photocatalytic Hydrogen Production by Boron Modified TiO 2 /Carbon Nitride Heterojunctions. ChemCatChem 2019, 11, 6408–6416. [Google Scholar] [CrossRef]
  55. Niu, P.; Zhang, L.; Liu, G.; Cheng, H.-M. Graphene-Like Carbon Nitride Nanosheets for Improved Photocatalytic Activities. Adv. Funct. Mater. 2012, 22, 4763–4770. [Google Scholar] [CrossRef]
  56. Ma, L.; Fan, H.; Wang, J.; Zhao, Y.; Tian, H.; Dong, G. Water-assisted ions in situ intercalation for porous polymeric graphitic carbon nitride nanosheets with superior photocatalytic hydrogen evolution performance. Appl. Catal. B Env. 2016, 190, 93–102. [Google Scholar] [CrossRef]
  57. Dias, E.M.; Christoforidis, K.C.; Francàs, L.; Petit, C. Tuning Thermally Treated Graphitic Carbon Nitride for H2 Evolution and CO2 Photoreduction: The Effects of Material Properties and Mid-Gap States. ACS Appl. Energy Mater. 2018, 1, 6524–6534. [Google Scholar] [CrossRef]
  58. Li, P.; Liu, L.; An, W.; Wang, H.; Guo, H.; Liang, Y.; Cui., W. Ultrathin porous g-C3N4 nanosheets modified with AuCu alloy nanoparticles and C-C coupling photothermal catalytic reduction of CO2 to ethanol. Appl. Catal. B Env. 2020, 266, 118618. [Google Scholar] [CrossRef]
  59. Chen, Y.; Wang, B.; Lin, S.; Zhang, Y.; Wang, X. Activation of n → π* Transitions in Two-Dimensional Conjugated Polymers for Visible Light Photocatalysis. J. Phys. Chem. C 2014, 118, 29981–29989. [Google Scholar] [CrossRef]
  60. Ou, H.; Chen, X.; Lin, L.; Fang, Y.; Wang, X. Biomimetic Donor-Acceptor Motifs in Conjugated Polymers for Promoting Exciton Splitting and Charge Separation. Angew. Chem. Int. Ed. 2018, 57, 8729–8733. [Google Scholar] [CrossRef]
  61. Liu, G.; Zhao, G.; Zhou, W.; Liu, Y.; Pang, H.; Zhang, H.; Hao, D.; Meng, X.; Li, P.; Kako, T.; et al. In Situ Bond Modulation of Graphitic Carbon Nitride to Construct p-n Homojunctions for Enhanced Photocatalytic Hydrogen Production. Adv. Funct. Mater. 2016, 26, 6822–6829. [Google Scholar] [CrossRef]
  62. Ran, J.; Qu, J.; Zhang, H.; Wen, T.; Wang, H.; Chen, S.; Song, L.; Zhang, X.; Jing, L.; Zheng, R.; et al. 2D Metal Organic Framework Nanosheet: A Universal Platform Promoting Highly Efficient Visible-Light-Induced Hydrogen Production. Adv. Energy Mater. 2019, 9, 1803402. [Google Scholar] [CrossRef]
  63. Liang, Y.; Shang, R.; Lu, J.; An, W.; Hu, J.; Liu, L.; Cui, W. 2D MOFs enriched g-C3N4 nanosheets for highly efficient charge separation and photocatalytic hydrogen evolution from water. Int. J. Hydrog. Energy 2019, 44, 2797–2810. [Google Scholar] [CrossRef]
  64. Mu, Q.; Zhu, W.; Li, X.; Zhang, C.; Su, Y.; Lian, Y.; Qi, P.; Deng, Z.; Zhang, D.; Wang, S.; et al. Electrostatic charge transfer for boosting the photocatalytic CO2 reduction on metal centers of 2D MOF/rGO heterostructure. Appl. Catal. B Environ. 2020, 262, 118144. [Google Scholar] [CrossRef]
  65. Chena, W.; Han, B.; Tian, C.; Liu, X.; Liang, S.; Deng, H.; Lin, Z. MOFs-derived ultrathin holey Co3O4 nanosheets for enhanced visible light CO2 reduction. Appl. Catal. B Environ. 2019, 244, 996–1003. [Google Scholar] [CrossRef]
  66. Christoforidis, K.C.; Iglesias-Juez, A.; Figueroa, S.J.A.; Di Michiel, M.; Newton, M.A.; Fernandez-Garcia., M. Structure and activity of iron-doped TiO2-anatase nanomaterials for gas-phase toluene photo-oxidation. Catal. Sci. Technol. 2013, 3, 626–634. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of e/h+ pairs formation upon the irradiation of a SC with the energy of incident light greater than the band gap energy and their participation in redox reactions for solar fuel production. Adopted from Ref. [10].
Figure 1. Schematic illustration of e/h+ pairs formation upon the irradiation of a SC with the energy of incident light greater than the band gap energy and their participation in redox reactions for solar fuel production. Adopted from Ref. [10].
Solar 02 00017 g001
Figure 2. Charge transfer in a type-II (left side) and a Z-scheme (right side) heterojunction. Adapted from Ref. [10].
Figure 2. Charge transfer in a type-II (left side) and a Z-scheme (right side) heterojunction. Adapted from Ref. [10].
Solar 02 00017 g002
Figure 3. Spatial separation of charges in anatase TiO2 of varying {001} and {101} facets ratio. Reprinted with permission from Ref. [21]. Copyright 2014, American Chemical Society.
Figure 3. Spatial separation of charges in anatase TiO2 of varying {001} and {101} facets ratio. Reprinted with permission from Ref. [21]. Copyright 2014, American Chemical Society.
Solar 02 00017 g003
Figure 4. Schematic representation of the synthesis protocol for the coupling of ultrathin CN structures (CNNS) with TiO2 nanosheets (Ti-NS) or unshaped TiO2 nanoparticles (Ti-ISO) (top). CO2 adsorption (25 °C and 1 bar) and photocatalytic CO production under the UV-Vis irradiation of pure TiO2 and the corresponding composites with CNNS (bottom). Reprinted with permission from Ref. [14]. Copyright 2019, Elsevier.
Figure 4. Schematic representation of the synthesis protocol for the coupling of ultrathin CN structures (CNNS) with TiO2 nanosheets (Ti-NS) or unshaped TiO2 nanoparticles (Ti-ISO) (top). CO2 adsorption (25 °C and 1 bar) and photocatalytic CO production under the UV-Vis irradiation of pure TiO2 and the corresponding composites with CNNS (bottom). Reprinted with permission from Ref. [14]. Copyright 2019, Elsevier.
Solar 02 00017 g004
Figure 5. Charge separation in the CdSe ZnCr-LDH heterojunction (left panel) and H2 evolution (right panel) activity of the ZnCr-LDH having different CdSe contents under visible light (λ > 420 nm). Reprinted with permission from Ref. [38]. Copyright 2015, Elsevier.
Figure 5. Charge separation in the CdSe ZnCr-LDH heterojunction (left panel) and H2 evolution (right panel) activity of the ZnCr-LDH having different CdSe contents under visible light (λ > 420 nm). Reprinted with permission from Ref. [38]. Copyright 2015, Elsevier.
Solar 02 00017 g005
Figure 6. Hierarchical CdS-Au core/shell nanostructures and H2 production on MoS2. Reproduced with permission of Ref. [43]. Copyright 2018, American Chemical Society.
Figure 6. Hierarchical CdS-Au core/shell nanostructures and H2 production on MoS2. Reproduced with permission of Ref. [43]. Copyright 2018, American Chemical Society.
Solar 02 00017 g006
Figure 7. Formation of MoS2 nanosheets/Cu2S snowflake nanocomposite for efficient H2 production. Reprinted with permission from Ref. [44]. Copyright 2018, Elsevier.
Figure 7. Formation of MoS2 nanosheets/Cu2S snowflake nanocomposite for efficient H2 production. Reprinted with permission from Ref. [44]. Copyright 2018, Elsevier.
Solar 02 00017 g007
Figure 8. Rate of photocatalytic H2 production by a TiO2 NSs and MoS2@TiO2 composite (selective and random deposition). Reprinted with permission from Ref. [46]. Copyright 2019, Elsevier.
Figure 8. Rate of photocatalytic H2 production by a TiO2 NSs and MoS2@TiO2 composite (selective and random deposition). Reprinted with permission from Ref. [46]. Copyright 2019, Elsevier.
Solar 02 00017 g008
Figure 9. (a) Schematic representation of the TiO2/MoS2 heterojunction, charge separation under UV-Vis light irradiation and CO2 reduction into CH4 and CH3OH. (b) The charge transfer mechanism in the homojunction formed between anatase and rutile. Reprinted with permission from Ref. [47]. Copyright 2018, John Wiley and Sons.
Figure 9. (a) Schematic representation of the TiO2/MoS2 heterojunction, charge separation under UV-Vis light irradiation and CO2 reduction into CH4 and CH3OH. (b) The charge transfer mechanism in the homojunction formed between anatase and rutile. Reprinted with permission from Ref. [47]. Copyright 2018, John Wiley and Sons.
Solar 02 00017 g009
Figure 10. Photocatalytic H2 production under UV-Vis and visible light irradiation (top) and mid-gap state formation (bottom) in the thermally treated CN. Reprinted with permission from Ref. [57]. Copyright 2018, American Chemical Society.
Figure 10. Photocatalytic H2 production under UV-Vis and visible light irradiation (top) and mid-gap state formation (bottom) in the thermally treated CN. Reprinted with permission from Ref. [57]. Copyright 2018, American Chemical Society.
Solar 02 00017 g010
Figure 11. Schematic illustration for the development of ultrathin Co3O4 nanosheets using ZIF-67 as a precursor. Reprinted with permission from Ref. [65]. Copyright 2019, Elsevier.
Figure 11. Schematic illustration for the development of ultrathin Co3O4 nanosheets using ZIF-67 as a precursor. Reprinted with permission from Ref. [65]. Copyright 2019, Elsevier.
Solar 02 00017 g011
Table 1. Required redox potential of selected reactions [10].
Table 1. Required redox potential of selected reactions [10].
ReactionsE0 (SHE, pH 7)
H 2 O + 4 h + 4 H + + 2 O 2 + 4 e +0.81
H 2 + 2 e H 2 −0.42
CO 2 + e CO 2 −1.90
CO 2 + 2 H + + 2 e HCOOH −0.61
CO 2 + 2 H + + 2 e CO + H 2 O −0.53
CO 2 + 4 H + + 4 e HCHO + H 2 O −0.48
CO 2 + 6 H + + 6 e CH 3 OH + H 2 O −0.38
CO 2 + 8 H + + 8 e CH 4 + 2 H 2 O −0.24
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ioannidou, T.; Anagnostopoulou, M.; Christoforidis, K.C. Two-Dimensional Photocatalysts for Energy and Environmental Applications. Solar 2022, 2, 305-320. https://doi.org/10.3390/solar2020017

AMA Style

Ioannidou T, Anagnostopoulou M, Christoforidis KC. Two-Dimensional Photocatalysts for Energy and Environmental Applications. Solar. 2022; 2(2):305-320. https://doi.org/10.3390/solar2020017

Chicago/Turabian Style

Ioannidou, Thaleia, Maria Anagnostopoulou, and Konstantinos C. Christoforidis. 2022. "Two-Dimensional Photocatalysts for Energy and Environmental Applications" Solar 2, no. 2: 305-320. https://doi.org/10.3390/solar2020017

Article Metrics

Back to TopTop